16
Optimizing QAOA: Success Probability and Runtime Dependence on Circuit Depth Murphy Yuezhen Niu, 1, 2, 3, * Sirui Lu, 4 and Isaac L. Chuang 1, 2 1 Research Laboratory of Electronics, Massachusetts Institute of Technology, Cambridge, Massachusetts, 02139, USA 2 Department of Physics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA 3 Google Inc., 340 Main Street, Venice, CA 90291 4 Department of Physics, Tsinghua University, Beijing, 100084, China (Dated: May 30, 2019) The quantum approximate optimization algorithm (QAOA) first proposed by Farhi et al. promises near-term applications based on its simplicity, universality, and provable optimality. A depth-p QAOA consists of p interleaved unitary transformations induced by two mutually non-commuting Hamiltonians. A long-standing question concerning the performance of QAOA is the dependence of its success probability as a function of circuit depth p. We make initial progress by analyzing the success probability of QAOA for realizing state transfer in a one-dimensional qubit chain using two- qubit XY Hamiltonians and single-qubit Hamiltonians. We provide analytic state transfer success probability dependencies on p in both low and large p limits by leveraging the unique spectral property of the XY Hamiltonian. We support our proof under a given QAOA ansatz with numerical optimizations of QAOA for up to N =20 qubits. We show that the optimized QAOA can achieve the well-known quadratic speedup, Grover speedup, over the classical alternatives. Treating the QAOA optimization as a quantum control problem, we also provide numerical evidence of how the circuit depth determines the controllability of the QAOA ansatz. I. INTRODUCTION Quantum approximate optimization algo- rithm (QAOA) promises near-term applications for quantum devices given its simplicity, universality and provable optimality. In contrast to quantum adiabatic algorithms [10, 11], QAOA adopts abrupt switching between two different Hamiltonian evolutions [12]. This simple strategy can reduce the complexity of Hamil- tonian controls by obviating the need to continuously vary Hamiltonian in time. Fine-tuned controls over Hamiltonian trajectories are otherwise necessary for the traditional quantum adiabatic algorithms. Despite its simplicity, QAOA is computationally universal [22]. It also has important implications in computational complexity: the efficient classical sampling of a depth-1 QAOA will collapse the polynomial hierarchy to the third level [9]. Lastly, from Pontryagin’s maximum principle [31], QAOA is optimal for solving variational problems whose cost function is a linear function of the system Hamiltonian [35]. Despite these attractive properties, to be suitable for near-term quantum devices, a long-standing question re- mains to be addressed: how does the success probability of QAOA depend on its circuit depth? Near-term quan- tum device’s computation time is limited by noise and decoherence. This in turn limits the realizable quantum algorithms to relatively short circuit depth. To under- stand QAOA’s potential or limitations for near-term ap- plications, it is therefore critical to understand its per- formance when fixing the upper bound on the depth of the QAOA circuit. * email: [email protected] It is exceedingly hard to study the QAOA performance without choosing the QAOA Hamiltonian and optimiza- tion problem. This is mainly due to the lack of a sufficient condition for QAOA to achieve optimality in a generic scenario. Since the QAOA success probability directly depends on its optimality, a bound on QAOA success probability scaling usually requires problem-specific nu- merical optimizations. Recently, specific properties of the chosen optimization problem and QAOA Hamiltonians are utilizied to design protocols that imitate the Grover search algorithm [17, 30], or to prepare highly entangled quantum states [15, 16]. These encouraging results spot- light the importance of the problem and hardware spe- cific information such as the controllable system Hamil- tonians in designing QAOA algorithms for improving its performance guarantee. In this work, we make initial progress towards answer- ing this question by analyzing the performance of QAOA for state transfer in a one-dimensional qubit chain using the XY Hamiltonians and single-qubit Hamiltonians as QAOA ansatzes. We choose quantum state transfer as our QAOA optimization task considering its simplicity and wide applications [1, 2, 5–7, 28, 36]. State transfer is a preliminary requirement for realizing quantum net- works which are necessary for connecting quantum com- puters to form large-scale computation network [8, 18]. We choose the QAOA Hamiltonian ansatz based on the experimental availability, where XY couplings are avail- able in existing superconducting qubit device [14]. More- over, the XY Hamiltonian’s particle number conserving nature makes it suitable for realizing state transfer within a given particle number subspace, which can mitigate unwanted information leakage into the higher excitation subspace. We harness the analytic spectral features of the XY Hamiltonians to derive explicit success probability P succ arXiv:1905.12134v1 [quant-ph] 28 May 2019

Optimizing QAOA: Success Probability and Runtime ... · Optimizing QAOA: Success Probability and Runtime Dependence on Circuit Depth Murphy Yuezhen Niu,1,2,3, Sirui Lu,4 and Isaac

  • Upload
    others

  • View
    10

  • Download
    0

Embed Size (px)

Citation preview

  • Optimizing QAOA: Success Probability and Runtime Dependence on Circuit Depth

    Murphy Yuezhen Niu,1, 2, 3, ∗ Sirui Lu,4 and Isaac L. Chuang1, 2

    1Research Laboratory of Electronics, Massachusetts Institute of Technology, Cambridge, Massachusetts, 02139, USA2Department of Physics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA

    3Google Inc., 340 Main Street, Venice, CA 902914Department of Physics, Tsinghua University, Beijing, 100084, China

    (Dated: May 30, 2019)

    The quantum approximate optimization algorithm (QAOA) first proposed by Farhi et al. promisesnear-term applications based on its simplicity, universality, and provable optimality. A depth-pQAOA consists of p interleaved unitary transformations induced by two mutually non-commutingHamiltonians. A long-standing question concerning the performance of QAOA is the dependence ofits success probability as a function of circuit depth p. We make initial progress by analyzing thesuccess probability of QAOA for realizing state transfer in a one-dimensional qubit chain using two-qubit XY Hamiltonians and single-qubit Hamiltonians. We provide analytic state transfer successprobability dependencies on p in both low and large p limits by leveraging the unique spectralproperty of the XY Hamiltonian. We support our proof under a given QAOA ansatz with numericaloptimizations of QAOA for up to N=20 qubits. We show that the optimized QAOA can achieve thewell-known quadratic speedup, Grover speedup, over the classical alternatives. Treating the QAOAoptimization as a quantum control problem, we also provide numerical evidence of how the circuitdepth determines the controllability of the QAOA ansatz.

    I. INTRODUCTION

    Quantum approximate optimization algo-rithm (QAOA) promises near-term applications forquantum devices given its simplicity, universality andprovable optimality. In contrast to quantum adiabaticalgorithms [10, 11], QAOA adopts abrupt switchingbetween two different Hamiltonian evolutions [12]. Thissimple strategy can reduce the complexity of Hamil-tonian controls by obviating the need to continuouslyvary Hamiltonian in time. Fine-tuned controls overHamiltonian trajectories are otherwise necessary forthe traditional quantum adiabatic algorithms. Despiteits simplicity, QAOA is computationally universal [22].It also has important implications in computationalcomplexity: the efficient classical sampling of a depth-1QAOA will collapse the polynomial hierarchy to thethird level [9]. Lastly, from Pontryagin’s maximumprinciple [31], QAOA is optimal for solving variationalproblems whose cost function is a linear function of thesystem Hamiltonian [35].

    Despite these attractive properties, to be suitable fornear-term quantum devices, a long-standing question re-mains to be addressed: how does the success probabilityof QAOA depend on its circuit depth? Near-term quan-tum device’s computation time is limited by noise anddecoherence. This in turn limits the realizable quantumalgorithms to relatively short circuit depth. To under-stand QAOA’s potential or limitations for near-term ap-plications, it is therefore critical to understand its per-formance when fixing the upper bound on the depth ofthe QAOA circuit.

    ∗ email: [email protected]

    It is exceedingly hard to study the QAOA performancewithout choosing the QAOA Hamiltonian and optimiza-tion problem. This is mainly due to the lack of a sufficientcondition for QAOA to achieve optimality in a genericscenario. Since the QAOA success probability directlydepends on its optimality, a bound on QAOA successprobability scaling usually requires problem-specific nu-merical optimizations. Recently, specific properties of thechosen optimization problem and QAOA Hamiltoniansare utilizied to design protocols that imitate the Groversearch algorithm [17, 30], or to prepare highly entangledquantum states [15, 16]. These encouraging results spot-light the importance of the problem and hardware spe-cific information such as the controllable system Hamil-tonians in designing QAOA algorithms for improving itsperformance guarantee.

    In this work, we make initial progress towards answer-ing this question by analyzing the performance of QAOAfor state transfer in a one-dimensional qubit chain usingthe XY Hamiltonians and single-qubit Hamiltonians asQAOA ansatzes. We choose quantum state transfer asour QAOA optimization task considering its simplicityand wide applications [1, 2, 5–7, 28, 36]. State transferis a preliminary requirement for realizing quantum net-works which are necessary for connecting quantum com-puters to form large-scale computation network [8, 18].We choose the QAOA Hamiltonian ansatz based on theexperimental availability, where XY couplings are avail-able in existing superconducting qubit device [14]. More-over, the XY Hamiltonian’s particle number conservingnature makes it suitable for realizing state transfer withina given particle number subspace, which can mitigateunwanted information leakage into the higher excitationsubspace.

    We harness the analytic spectral features of the XYHamiltonians to derive explicit success probability Psucc

    arX

    iv:1

    905.

    1213

    4v1

    [qu

    ant-

    ph]

    28

    May

    201

    9

    mailto:email: [email protected]

  • 2

    dependencies on circuit depth p for state transfer in twodifferent limits. In the low circuit depth and short QAOAduration limits we have limp→0,δ→0 Psucc ∝ p2; and in thelarge circuit depth limit we have limp→∞,δ→0 Psucc ∝ p4p.Our proof reconfirms the achievable Grover speedup withQAOA ansatz [17]. As a compliment to the system-specific analysis, we apply the existing results of theLieb-Robinson bound to the QAOA success probabil-ity with any 2-local bounded-norm qubit Hamiltoniansfor one-dimensional state transfer. To verify the opti-mality of our scaling proof, we numerically optimize theassociated QAOA ansatz for different circuit depth andoverall runtime. Our numerical results confirm the ex-pected quadratic Grover-like scaling. We also demon-strate an interesting connection between the achievablesuccess probability and its controllability dependency onthe circuit depth: once the circuit depth is too low, theQAOA is no longer controllable when the control land-scape is full of local optima that are not globally optimal.

    The structure of the paper is as follows: in Sec. II weintroduce the basic setup of the QAOA for state trans-fer using the XY Hamiltonian; in Sec. III we derive theQAOA success probability as a function of circuit depthin both low and large circuit depth limit; we discuss theassociated quantum speed limit in Sec. IV; we summarizethe performance of the QAOA in regard to circuit depth,runtime and number of qubits in Sec. V, and conclude inSec. VI.

    II. QAOA FOR STATE TRANSFER

    We introduce in this section the basic concept of quan-tum state transfer and its realization through QAOA.Quantum state transfer has been proposed in both quan-tum optical systems and condensed matter system [2, 36].Different state transfer schemes include the use of quan-tum disorder [5, 6], optimal control [37], long-range in-teraction [28] and adiabatic evolution under a movingpotential [1]. Unlike the majority of these existing ap-proaches, QAOA resorts to a discrete set of operations ofa size given by the QAOA circuit depth. Such a circuit-based model is naturally suitable for near-term quantumdevices such as superconducting qubits and ion traps, butmore flexible than the traditional circuit-based model us-ing gates only from a predefined universal gate set.

    The state transfer problem of interest is defined ina one-dimensional qubit chain of length N . We use|n〉 to represent a product state of a positive eigen-state of the local Pauli-z operator at the nth site andthe negative eigenstates of the local Pauli-z of othersites: σzn|n〉 = |n〉, σzi,i 6=n|n〉 = −|n〉, e.g. |n〉 =|0〉1|0〉2 · · · |0〉n−1|1〉n|0〉n+1 · · · |0〉N . We denote |0〉 asthe product state of the negative eigenvalue eigenstate ofthe local Pauli-z operators: |0〉 = |0〉1|0〉2 · · · |0〉N . If wetreat qubits as spins, and negative eigenvalue of Pauli-z operator as an excitation of the spin state, the statetransfer problem we will solve lies in the span of zero and

    single excitation subspaces. In this subspace, a quantumstate with boundary excitation is represented as

    |ψi〉 = α|1〉+ β|0〉, (1)

    with |α|2+|β|2 = 1. Starting from the state |ψi〉, the taskof state transfer is therefore to realize a unitary transfor-mation U such that:

    |ψf 〉 = U |ψi〉 = α|N〉+ β|0〉. (2)

    We choose the two Hamiltonians used for QAOA iter-ation to be

    ĤC = |N〉〈N | =1

    2(σzN + IN ), (3)

    ĤB =

    N∑i=1

    (σxi σxi+1 + σ

    yi σ

    yi+1). (4)

    The reasons for our choice of QAOA Hamiltonians aretwo-fold: (1) ĤC ’s plus one eigenstate is our transferredtarget state |N〉 and can thus serve as a Grover-like or-acle by assigning a phase to the target state; (2) ĤB isoff-diagonal and induces a swap operation between neigh-boring qubits, and thus can move the excitation aroundfor the purpose of state transfer.

    Since the total qubit-z operator Sz =∑Ni=1 σ

    zi com-

    mutes with both ĤC and ĤB and |1〉 is an eigenstateof the total qubit-z operator: Sz|1〉 = (1 − N)|1〉, thetotal excitation is conserved throughout the QAOA sim-ulation. We can therefore solve the quantum dynamicsin the subspace spanned by {|0〉, |1〉, |2〉, · · · , |N〉}.

    Denoting the unitary evolution under ĤC for a timeduration t as U(ĤC , t) and the unitary evolution under

    ĤB for time duration t′ as U(ĤB , t

    ′), a depth p QAOArealizes the following unitary transformation:

    Up =

    p∏k=1

    U(ĤC , δCk )U(ĤB , δ

    Bk ), (5)

    where the durations of evolutions under given Hamiltoni-ans are represented by δBk and δ

    Ck . Here, each kth QAOA

    iteration consists of a unitary evolution under ĤB fortime δBk then followed by a unitary evolution under ĤCfor time δCk . Since the eigenvalues of ĤB are not rational,

    unlike the original QAOA, the rotation angle δBk of ĤBis not restricted to (0, 2π).

    Since the zero excitation state of the system is invariantunder the unitary evolution of both Hamiltonians, thestate transfer task is equivalent to realizing the unitarytransformation: |1〉 → |N〉. Thus, we can quantify thefidelity of state transfer by the fidelity between Up|1〉 and|N〉:

    F = |〈N |Up|1〉|2 = 〈1|U†pĤCUp|1〉. (6)

    This is equivalent to the success probability, so we willuse them interchangeably henceforth. We can then treat

  • 3

    state transfer as a special kind of maximization satis-faction problem, except that the cost function F herecontains only one clause as opposed to many clauses intraditional QAOAs [12].

    It is shown in [7] that if we have complete control overthe amplitudes of XY coupling at different sites, perfectstate transfer can be realized through a single unitaryevolution under the XY Hamiltonian. This is, however,unrealistic for near-term devices, where the system cali-bration for such fine-tuned interactions is costly and themaximum interaction strength is limited.

    III. SUCCESS PROBABILITY SCALING AS AFUNCTION OF CIRCUIT DEPTH

    In this section, we derive the success probability scal-ing as a function of circuit depth. Our analysis is based

    on an iterative procedure using the knowledge from thespectrum of the QAOA ansatz Hamiltonian.

    To simplify our analysis, we adopt the following QAOAansatz: the duration under the evolution of dispersionHamiltonian ĤB , δ, is short and the same for differentiterations, and the evolution under the diagonal Hamil-tonian ĤC is of angle π, resembling a Grover oracle:

    Up =(e−iπ|N〉〈N |U(ĤB , δ)

    )p. With this ansatz, our re-

    sult can be connected to the scaling analysis in conven-tional Grover search and thus serves as a lower boundon the success probability for the optimized QAOA to bediscussed in the subsequent sections.

    We first diagonalize the dispersion Hamiltonian ĤB inthe single excitation subspace to obtain its kth eigen-state:

    |φk〉 =1√N/2

    N/2∑n=1

    (sin

    [knπ

    N/4 + 1

    ]|2n〉+ sin

    [k(n+ 1/2)π

    N/4 + 1

    ]|2n− 1〉

    ), (7)

    with the kth eigenvalue being Ek = 2 cos[kπ

    N/2+1 ].

    Given the initial state of the system as |ψ0〉 = |1〉, thesystem evolves to |ψ1〉 = e−iĤBδ|1〉 after a depth oneQAOA. Then the success probability of transferring theexcitation to the other end of chain after a depth oneQAOA is

    Psucc(1) = 〈1|eiĤBδ|N〉〈N |e−iĤBδ|1〉 = |fN1 (δ)|2, (8)

    where we use fN1 (δ) = 〈N |e−iĤBδ|1〉 to represent theamplitude of target state. Now we apply another QAOA

    iteration to update the quantum system to

    |ψ2〉 = U(ĤB , δ)U(ĤC , π)|ψ1〉= e−i2ĤBδ|1〉 − 2e−iĤBδ|N〉〈N |e−iĤBδ|1〉

    (9)

    This gives the success probability of the state transferafter a depth two QAOA as:

    Psucc(2) = 〈ψ2|N〉〈N |ψ2〉 = |fN1 (2δ)− 2fN1 (δ)fNN (δ)|2,(10)

    where fNN (δ) = 〈N |e−iĤBδ|N〉 denotes the amplitude ofthe state |N〉 remaining in state |N〉 after Hamiltonianevolution under ĤB for time δ. Similarly, we continue theiteration to obtain the success probability after a depththree and depth four QAOA:

    Psucc(3) = fN1 (3δ)− 2fN1 (2δ)fNN (δ)− 2fN1 (δ)fNN (2δ) + 4

    (fNN (δ)

    )2fN1 (δ), (11)

    Psucc(4) = fN1 (4δ)− 2fN1 (3δ)fNN (δ)− 2fN1 (2δ)fNN (2δ)− 2fN1 (δ)fNN (3δ) + 4

    (fNN (δ)

    )2fN1 (2δ)

    + 4(fNN (δ)

    )2fNN (2δ) + 4f

    NN (2δ)f

    NN (δ)f

    N1 (δ)− 8

    (fNN (δ)

    )3fN1 (δ).

    (12)

    Now we provide expression for transition amplitude used above given the exact eigenstates of the dispersion Hamil-tonian ĤB in Eq. 7:

  • 4

    fN1 (δ) =

    N/2∑k=1

    〈N |φk〉〈φk|1〉e−iEkδ,

    ≈ − iδ2√N + 1

    (cos

    (2π(N/2)2 + 4πN/2 + π

    2(N/4 + 1)

    )csc

    (πN/2 + 2π

    2(N/4 + 1)

    )+ cos

    2(N/4 + 1)

    )csc

    (πN/2 + 2π

    2(N/4 + 1)

    ))= −iF (N)δ,

    (13)

    fNN (δ) =

    N/2∑k=1

    〈N |φk〉〈φk|N〉e−iEkδ,

    ≈ iδ2√N + 1

    {1

    2

    [− cos

    (4πN2 + πN − π

    2(N + 1)

    )csc

    (πN

    N + 1

    )+ 2N + cos

    (π(N − 1)2(N + 1)

    )csc

    (πN

    N + 1

    )]×[cos

    (πN

    2(N + 1)

    )csc

    2(N + 1)

    )+ cos

    (π(N + 2)

    2(N + 1)

    )csc

    2(N + 1)

    )]−[cos

    (3πN

    2(N + 1)

    )csc

    (π − 2π

    2(N + 1)

    )− cos

    (−4πN2 − πN2(N + 1)

    )csc

    (π − 2πN2(N + 1)

    )]−[cos

    (πN

    2(N + 1)

    )csc

    (2πN + π

    2(N + 1)

    )− cos

    (4πN2 + 3πN

    2(N + 1)

    )csc

    (2πN + π

    2(N + 1)

    )]}=− iG(N)δ

    (14)

    where the approximation is made to include only theterms that are of either zero or first order in δ, underthe limit δ → 0. And both F (N) and G(N) are real-valued. For a QAOA of depth p, we deduce the successprobability dependency on transition probabilities fN1 (δ)and fNN (δ) as

    Psucc(p) =

    p∑j=1

    (−1)j∑~vj∈Vj

    fN1 (~vj(1)δ)

    j−1∏k=1

    fNN (~vj(k + 1)δ)

    (15)where ~vj is a vector with each element representing thevalue of a j partition of p that belongs to the set Vj ={~vj |

    ∑jk=1 ~vj(k) = p}. The success probability can be in

    turn be expressed as

    Psucc(p) ≈

    ∣∣∣∣∣∣p+1∑j=1

    Ajδj

    ∣∣∣∣∣∣2

    , (16)

    with the each amplitude Aj given by

    A1 = −ipF (N), (17)

    A2 = −F (N)G(N)p(p+ 1)(p+ 2)

    3, (18)

    limn→∞

    An ≈ −F (N)G(N)np2n−1. (19)

    Eq. (19) is derived from the asymptotic value of the prod-uct of all possible values of n integers p1, p2, . . . , pn whosesum equals p:

    ∑ni=1 pi = p.

    So far we have only kept the leading order in O(δ)

    together with all orders of O((p2δ)n)

    which will be non-

    negligible when p2δ ∼ 1 or p2δ � 1. For the scalinganalysis, we neglect constant terms in the sum and findthe success amplitude to be

  • 5

    p+1∑j=1

    Ajδj = −i(p+ 1)F (N)δ −

    [F (N)G(N)p3δ2 + · · ·+ F (N)G(N)pp2p+1δp+1

    ]= −i(p+ 1)F (N)δ − F (N)G(N)p

    3δ2[(G(N)p2δ)p − 1

    ](G(N)p2δ)2 − 1 .

    (20)

    Since the amplitude is composed of imaginary and real parts, the success probability of a depth-p QAOA is thusfound to be

    Psucc(p) ≈ (p+ 1)2F (N)2δ2 +F (N)2G(N)2p6δ4

    [(G(N)p2δ)p − 1

    ]2[(G(N)p2δ)2 − 1]2

    . (21)

    This success probability dependence can be used as alower bound on the QAOA performance after optimiza-tion where the duration of each evolution can be of flex-ible value. In the large depth limit, the term with thelargest power of p dominates:

    limp→∞

    Psucc(p) ∝ p4p+2. (22)

    This exponential growth in success probability is basedon the assumption that δ is a small constant (which doesnot change with the circuit depth p), and the dominantcontribution to the transition amplitudes is of the lowestorder in δ. Such exponential dependence is also found byincreasing the speed of adiabatic Hamiltonian evolution,see [27]. We also observe such exponential growth in ournumerically optimized QAOA (see Sec. III).

    In the low-depth limit, only the lowest order of δ termsdominates:

    limp→1

    Psucc(p) ∝ F (N)2δp2. (23)

    Then the total number of steps required to achieve thetarget state is of order O(1/(δF (N))) = O(

    √N). This

    quadratic Grover-like dependence on circuit depth can beunderstood by mapping a Grover algorithm to a QAOAroutine. The dispersion step of Grover iteration, whichis a rotation of angle π around the equal superpositionstate:

    Us = H⊗N (−2|0〉〈0|+ I)H⊗N , (24)

    with H representing the Hadamard gate, can be gener-alized to a rotation around any state that is not parallelto the target state |ψ〉 [23]:

    Us = −2|ψ〉〈ψ|+ I. (25)

    If we choose |ψ〉 = e−iδĤB |1〉, the corresponding pthGrover iteration for searching the transferred state |N〉starting from the initial state |1〉 can then be representedby the unitary realized by a depth p QAOA circuit as

    UpGrover =(e−iĤ

    1cπeiĤBδ1e−iĤ

    2cπe−iĤBδ1

    )p, (26)

    where Ĥ1c =12 (σ

    zN + IN ) and Ĥ

    2c =

    12 (σ

    z1 + I1).

    IV. QUANTUM SPEED LIMIT

    As a supplement to the success probability scalinganalysis presented in the previous section that is lim-ited to our specific choices of the QAOA Hamiltoni-ans, we review in this section the general constraints onthe QAOA performance using spatially local Hamilto-nians imposed by the Lieb-Robinson bound. The Lieb-Robinson bound [21] is a powerful tool to study the prop-agation of quantum correlation and thus quantum infor-mation in many-body quantum systems [13]. It serves asa lower bound on the success probability for the QAOAperformance of the same total runtime. Although sucha bound is not tight, nor does it directly depends on thecircuit depth of QAOA, it provides a basic reference ofthe optimality of QAOA in regard to its success proba-bility scaling as a function of physical time. And in fact,the theoretical insights of Lieb-Robinson bound help usto understand the performance of numerically optimizedQAOAs in the next section.

    We rewrite our QAOA iterations as an evolution underthe time-dependent Schródinger equation with the time-dependent Hamiltonian:

    Ĥ(t) = s(t)ĤC + [1− s(t)]ĤB , (27)

    where s(t) is the time varying control parameter that canonly take on the values zero and one. Thus it realizes thebang-bang form of QAOA iterations.

    Note Ĥ(t) can be written as the sum of nearest-

    neighbor interacion terms: Ĥ(t) =∑i hi,i+1(t). Thus,

    by Lieb-Robinson bound, the maximum speed of quan-tum information propagation in this system is bounded.This speed determines how fast operations on the firstqubit can affect observables on the last qubit at somelater moment of time, and thus upper bounds the speedof state transfer. For convenience, let us call the firstqubit A, the last qubit B, and the rest part C (see

  • 6

    Fig. 1). The Lieb-Robinson bound determines the maxi-mum operator norm of the commutator between any op-erators OA and OB on the first and the last qubit ata later time t. More specifically, denoting L = N − 1as the distance between the first and the last qubitand J = maxi maxt ‖hi,i+1(t)‖ as the maximum inter-action strength, the Lieb-Robinson bound for a nearest-neighbour Hamiltonian on aD-dimensional square lattice[13, 28] is given by

    ‖[OA(t), OB(0)]‖ ≤ 2‖OA‖‖OB‖∞∑k=L

    (2Jt(4D − 1))kk!

    .

    (28)In the one-dimensional system, the above bound simpli-fies to [28]:

    ‖[OA(t), OB(0)]‖ ≤ 2‖OA‖‖OB‖∞∑k=L

    (6Jt)k

    k!(29)

    ≤ 2‖OA‖‖OB‖ exp(6eJt− L) (30)= 2‖OA‖‖OB‖ exp (vt− L), (31)

    where the Lieb-Robinson velocity v = 6eJ is approxi-mately 32.616 because J = ‖σxi σxi + σzi σzi ‖ = 2. Conse-quently, 1/v ≈ 0.03.

    FIG. 1. The emergent light cone in state transfer problem,where the quantum state localized at A (the first qubit) istransfered to B (the last qubit) through the quantum channelC consisting of qubits in the middle. In the short-range two-local system, a non-relativistic light-cone x = vt emerges.The amount of information that can be transferred outside ofthe lightcone is exponentially small.

    Let U0A = IA and U1A = σ

    xA. The uni-

    tary transformation of the whole system is inducedby the time-dependent Hamiltonian as UABC(t) =

    T exp(∫ t0−iĤ(t′)dt′). Specifically, with initial state

    of the system given by ρ0 = |0〉〈0|, we can inter-pret this procedure as a quantum channel where the

    input state is ρkABC = UkAρ0U

    k†A , and the output

    state, the reduced density matrix of B, is σkB(t) =

    TrAC(UABC(t)ρkABCU

    †ABC(t)).

    Ref. [4] shows that σkB(t) depend on k as follows. Forany observable OB and associated time evolved operator

    OB(t) = U†ABC(t)OBUABC(t), we have:

    TrB[OB(σ

    0B(t)− σ1B(t))

    ]≤ �‖OB(t)‖, (32)

    where � = 2 exp(vt − L) is given by the Lieb-Robinsonbound. Here, we have used U1†A U

    1A = I, the definition of

    operator norm, and the Lieb-Robinson bound.Therefore σ1B(t) and σ

    0B(t) are � close in the trace norm:

    ‖σ1B(t) − σ0B(t)‖1 ≤ �. By using the following inequalitybetween the trace distance and the fidelity, F (ρ, σ) ≥1− 12‖ρ−σ‖1, we obtain a bound on the fidelity betweenσ1B(t) and σ

    0B(t):

    F (σ1B(t), σ0B(t)) ≥ 1−

    1

    2�. (33)

    Now we would like to bound the success probability P (t)of a excitation of qubit state transferred from site 1 tothe site N after the evolution under Ĥ(t) for time t whichdepends on the fidelity defined above as:

    1− P (t) = F 2(σ1B(t), σ0B(t)) ≥ (1−1

    2�)2. (34)

    Rearranging, we obtain an upper bound for the successprobability of the QAOA as a function of time and thelength of the qubit chain:

    P (t) ≤ �− 14�2, (35)

    From the above expression, we can identify three dif-ferent regions of temporal dynamics. At the early timewhen t � L/v, we have � = 2 exp(vt − L)) � 1 and theprobability of success is nearly zero. In this first region,the success probability is exponentially suppressed andremains almost zero in time. When t ≈ L/v we have� = c exp(vt − L) < 1 and the first term of the right-hand side of Eq. (35) dominates, which gives rise to anexponentially growing success probability. Finally whent > L/v, the second term of Eq. (35) starts balancing outthe first term, and gives rise to a steady growing region.A rough estimation of perfect state transfer time can begiven by setting �− 14�2 = 1⇒ � = 2, which gives

    t ≈ L/v. (36)

    The main weakness of the Lieb-Robinson method is thelack of dependence on the specific form of Hamiltonianand the circuit depth. Nevertheless, it offers useful in-sights into the difficulty of state transfer problems. Inthe later numerical section, we confirm the existences ofthe exponentially suppressed region, the exponentiallygrowing region (see Fig. 9), the steady growing region,and the linear dependence between tf required for statetransfer and the number of qubits N (see Fig. 11).

    V. NUMERICAL OPTIMIZATION OF THEQAOA

    Our analytic success probability versus circuit depthscaling analyses so far do not assume the optimality ofthe QAOA solution. To verify the tightness of these re-sults for optimized QAOAs, we explore in this section

  • 7

    the numerically optimized QAOA performance in regardto its success probability scaling as a function of the cir-cuit depth and the physical runtime. We start by intro-ducing briefly our numerical optimization method, andthen describe and analyze the optimized QAOA perfor-mance obtained from our numerical method. We showthat the quadratic Grover-like speedup shown in our an-alytic spectral analysis is also present in the numericallyoptimized QAOA solutions. We also show that whenthe circuit depth is too low, the given QAOA protocolbecomes uncontrollable: its control landscape no longerpossesses only global optimal but also many local op-timum points. Consequently, the optimized QAOA nolonger necessarily guarantees the existence of a high fi-delity state transfer scheme. This finding unveils the re-lation between the F − p scaling and the controllabilityof the underlying physical system.

    We optimize QAOA parameters under two differentconstraints, one with limited physical run time and theother one with unlimited physical run time. Both situa-tions are experimentally relevant. If the coherence timeis sufficiently large, we may want to achieve state trans-fer with a minimum number of switches between ĤB andĤC . In contrast, if the switching operation is easy, thetotal physical run time should be minimized. The prac-tical implementation of QAOAs on near-term quantumcomputing hardware is outside the scope of this article.

    A. Numerical methods

    We use a gradient descent method to numerically de-termine the maximum achievable fidelity given tf andp. We choose the optimization parameters as the dura-tion for each QAOA Hamiltonian evolution δkB and δ

    kC

    of each kth iteration with k ∈ [p] for a depth-p QAOA.The total physical runtime is the sum of all QAOA it-erations: tf =

    ∑k(δ

    kB + δ

    kC). We chose the parameter

    ranges for the physical run time tf and the circuit depthp by preliminary numerical experiments. Table II and?? summarize the parameters we performed grid searchon. To increase the reliability of the total number ofrandom restarts for gradient descent iterations, we ranpreliminary experiments with a varying size of randomuniform (over 1-simplex) initial conditions, and discov-ered that 200 random initial conditions were enough forN = 2→ 15 qubits cases to find approximate global opti-mal solutions. For N = 16→ 20 qubits, we increased thenumber of random restarts to 400. We use L-BFGS Mat-lab toolbox for the QAOA optimization, which can beefficiently parallelized for large-scale experiments. Thesenumerical results to be discussed below are summarizedin Table III.

    B. Numerical results for unlimited tf

    We start with the optimized QAOA performance whenthe physical run time tf is not fixed. The analytic re-sult at low circuit depth limit (Eq. (23)) coincides withour numerically result in Section V B 1. And the connec-tion between QAOA circuit depth and the controllabilityis demonstrated in the numerical results on the controllandscape in Section V B 2.

    1. Maximum achievable fidelity versus circuit depth p

    Figure 2 shows maximum achievable fidelity F as afunction of circuit depth p with no constraints on tf forN = 2→ 20 qubits. The circuit depth dependence of suc-cess probability in Fig. 2 agrees with our analytic result:at the very beginning the quadratic dependence domi-nates (Eq. (23)), and a while later the exponential slowdown dominates (Eq. (35)). The time duration ansatzused in our analytical result in Sec. III is also supportedin our numerically optimized solution; see Fig. 4 for ex-ample, where the intervals for s(t) = 1 corresponding to

    the evolution under ĤC is shorter on average than theduration in which s(t) = 0.

    2. Control landscape

    The optimization of QAOA can be regarded as a quan-tum control problem, where the durations of differentQAOA Hamiltonian evolutions are the control parame-ters, and the fidelity of the state transfer is the controlcost function to be maximized. Under this analogy, whenthe system is controllable, the landscape of the controlcost function over parameter space generically has onlyglobal minima [26, 29, 34]. When the system is uncontrol-lable, the quantum control landscape admits many localminima [33]. In our case, if we allow the circuit depthto be infinite, the system is controllable; but the con-trollability for an intermediate number of circuit depthsdemands further investigation. As a simple example, weplot the control landscape for N = 3 qubits in Fig. 5with p = 2, 4. Fig. 5 shows that the QAOA ansatz withp = 2 is uncontrollable, as there are many local minima.In contrast, the case p = 4 admits only global minima(at least for the plotted part); thus more controllable.We review the control viewpoint of optimizing QAOA inAppendix A.

    C. Numerical results for fixed tf : optimized QAOA

    We now discuss the optimized QAOA performancewith a fixed physical run time tf . In Section V C 1, weinvestigate the controllability dependence on tf . Second,in Section V C 2, we discuss the Lieb-Robinson type scal-ing emerged in the optimized QAOA performance.

  • 8

    1 2 3 4 5 6

    Circuit Depth p

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    1.2M

    axim

    um

    ach

    ieva

    ble

    fid

    elit

    yF

    Maximum achievable fidelity F versus circuit depth pfor N=2-10 qubits

    N=2

    N=3

    N=4

    N=5

    N=6

    N=7

    N=8

    N=9

    N=10

    (a) N = 2 → 10 qubits

    1 2 3 4 5 6 7 8 9 10 11

    Circuit Depth p

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    Max

    imu

    mac

    hie

    vab

    lefi

    del

    ityF

    Maximum achievable fidelity F versus circuit depth pfor N=11-20 qubits

    N=11

    N=12

    N=13

    N=14

    N=15

    N=16

    N=17

    N=18

    N=19

    N=20

    (b) N = 11 → 20 qubits

    FIG. 2. Maximum achievable fidelity F using QAOAs as afunction of the circuit depth p with no constraints on tf forN = 2 → 20 qubits. The dots are numerical points. We fitthe results with quadratic function F (p) = ap2 + bp + c, asrepresented by the lines. We observe that the fidelity growsat low circuit depth, and then slowly converges to 1.0.

    1. Maximum achievable fidelity F versus circuit depth p

    In this section, we numerically study the maximumachievable fidelity F versus the circuit depth p for a fixedtf in Figs. 6 to 7. Generally speaking, the larger circuitdepth QAOA should always perform better than lowerdepth ones. However, if p is too large, the difficulty ofthe QAOA optimization increases and the optimizationcan get stuck in local optima. This results in a non-monotonic behavior in numerically optimized fidelity asa function of circuit depth. For fixed tf , there is a circuitdepth p beyond which fidelity can no longer be improved.As shown in Fig. 6, for tf = 6 , the maximum achievablefidelity does not increase for circuit depth larger thanp = 3, and for tf = 13, the maximum achievable fidelitydoes not increase for circuit depth larger than p = 4.This observation is also intimately related to the control-lability of the QAOA: for the fixed run time, there existsa threshold circuit depth below which the QAOA is no

    1 2 3 4 5 6Circuit depth p

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    Max

    imum

    achi

    evab

    lefi

    delit

    yF N=2

    N=3

    N=4

    N=5

    N=6

    N=7

    N=8

    N=9

    N=10

    (a) N = 2 → 10 qubits

    1 2 3 4 5 6 7 8 9 10 11Circuit depth p

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    Max

    imum

    achi

    evab

    lefi

    delit

    yF N=11

    N=12

    N=13

    N=14

    N=15

    N=16

    N=17

    N=18

    N=19

    N=20

    (b) N = 11 → 20 qubits

    FIG. 3. Maximum achievable fidelity F as a function of thecircuit depth p with no constraints on tf for N = 2 → 20qubits.The dots are numerical points. We fit the results withinverted exponential function F (p) = 1− exp(−a(p− b)). Wecan observe fidelity grows rapidly at low circuit depth, andthen the fidelity slowly converge to unity.

    longer controllable.

    2. Maximum achievable fidelity F versus physical run timetf

    In this subsection, we investigate the performances ofthe QAOA with a fixed physical runtime in Figure 8. Weidentify three different temporal dependencies of fidelityas predicted by the Lieb-Robinson bound, as depictedin Fig. 9: exponentially suppressed region; exponentiallygrowing region; and steady growing region. We find thatthe longer the physical run time tf is, the better achiev-able fidelity will be under the condition that the circuitdepth p is sufficiently large and tf is outside of the highlysuppressed region. For a low depth circuit, the oscillat-ing in success probability occurs (Fig. 8(a)), which is asign of uncontrollability. Such oscillation disappears forsufficiently large p, see Fig. 8(b). The three regions of thegrowth region become more apparent as circuit depth in-

  • 9

    0 20 40 60time t

    0

    1

    s(t)

    FIG. 4. An optimal bang-bang solution obtained throughnumerical optimizations for N = 10 qubits and circuit depthp = 15. In this case, the optimal solution favors much shorterduration for the evolution under ĤB than that under ĤC .This is in accord with our analytical result in Sec. III.

    creases, see in Fig. 10.In Fig. 11, we fit the minimum required run time tf for

    achieving fidelity F = 0.99 as a function of the numberof qubits (N = 2→ 19). The linear dependence from theLieb-Robinson bound Eq. (36) is seen with tf ∼ 2.439N .Given the same amount of run time, we show in Fig. 12that the QAOA with a higher circuit depth necessarilyachieves higher success probability.

    The Lieb-Robinson bound gives a prediction aboutthe size of the exponentially suppression region: ts ∼N/(6eJ) = 0.03N . Practically, we define the exponen-tially suppressed time as the time needed to make fidelityhigher than 0.01. To see if it agrees with our numericalresult, we plot the exponentially suppressed time as afunction of the number of qubits in Fig. 13. Our nu-merical result is ts ∼ 0.246N with a coefficient of deter-mination r2 = 0.997. We remark that the discrepancybetween 0.246N and 0.03N is because our QAOAs onlyoperate in the span of zero and single excitation sub-spaces, while the Lieb-Robinson bound considered thefull N -qubit Hilbert space.

    VI. SUMMARY

    We study the QAOA’s success probability scaling asa function of circuit depth and the physical runtime forimplementing state transfer problems. By carefully uti-lizing the spectral properties of the QAOA Hamiltonians,we obtain analytic expressions for the success probabilityscaling as a function of the circuit depth. At the low-circuit-depth and short-physical-duration limit, our ana-lytic results reproduce the Grover-like quadratic speedup.We further study the success probability scaling in nu-merically optimized QAOAs for a chain of up to N = 20qubits (limited by computational resources). These nu-merical experiments confirm the quadratic speed up andmatch with the Lieb-Robinson analysis of quantum speedlimit, i.e., when the circuit depth p is sufficiently large,

    (a)

    (b)

    FIG. 5. A comparison of the control landscapes of the QAOAwith low circuit depth (p = 2) and of that with a larger circuitdepth (p = 4) for a three-qubit system. (a) The control land-scape for two chosen variables of a depth-2 QAOA, whichadmits a maximum achievable fidelity of 0.787. As we ob-served many local minima, the system is uncontrollable. (b)The control landscape for two chosen variables of a depth-4QAOA, which admits a maximum achievable fidelity of 1.000.Since all local minima are global minima, the system is con-trollable.

    with the increase of tf , there are three different scenar-ios of success probability scaling: (1) exponentially sup-pressed region, (2) exponentially growing region and (3)steadily growing region. Treating QAOA optimizationas a quantum control problem, we demonstrate the re-lation between the circuit depth and the controllabilityof QAOA: when the circuit depth is too low for a fixeddistance state transfer, the control landscape possessesmany locally optimal solutions that are not globally op-timal and the QAOA becomes uncontrollable. Althoughthe state transfer problem we considered here is relativelysimple, , our results offer valuable insights into the per-formance of QAOAs by connecting its optimality to the

  • 10

    2 4 6 8Circuit depth p

    0.00

    0.25

    0.50

    0.75

    1.00M

    axim

    umac

    hiev

    able

    fidel

    ityF

    N=5

    N=10

    N=15

    N=20

    (a) Plots for a small tf = 6.

    1 2 3 4 5 6 7 8 9

    Circuit depth p

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    Max

    imum

    achi

    evab

    lefid

    elit

    yF

    N=5

    N=10

    N=15

    N=20

    (b) Plots for a medium tf = 13.

    FIG. 6. Maximum achievable fidelity F versus circuit depth pwith the same fixed tf for N = 5, 10, 15, 19 qubits. For fixedtf , we observed that there exists a circuit depth p beyondwhich there will be no improvement of fidelity. We find adepth-3 circuit is sufficient for tf = 6 while a depth-4 circuitis needed for tf = 13.

    Grover speedup and its success probability dependenceon circuit-depth to the controllability the QAOA ansatz.To fully explore the application of QAOA, however, morework remains to be done to study the effect of realisticquantum noise on QAOA implementations.

    [1] V. Balachandran and J. Gong. Adiabatic quantum trans-port in a spin chain with a moving potential. Phys. Rev.A, 77:012303, Jan 2008.

    [2] S. Bose. Quantum communication through an unmodu-lated spin chain. Phys. Rev. Lett., 91:207901, Nov 2003.

    [3] F. G. Brandao, M. Broughton, E. Farhi, S. Gutmann,and H. Neven. For fixed control parameters the quantumapproximate optimization algorithm’s objective functionvalue concentrates for typical instances. arXiv preprintarXiv:1812.04170, 2018.

    [4] S. Bravyi, M. B. Hastings, and F. Verstraete. Lieb-robinson bounds and the generation of correlations andtopological quantum order. Phys. Rev. Lett., 97:050401,Jul 2006.

    [5] C. K. Burrell and T. J. Osborne. Bounds on the speedof information propagation in disordered quantum spinchains. Phys. Rev. Lett., 99:167201, Oct 2007.

    [6] C. K. Burrell, J. Eisert, and T. J. Osborne. Informa-tion propagation through quantum chains with fluctuat-

    ing disorder. Phys. Rev. A, 80:052319, Nov 2009.[7] M. Christandl, N. Datta, T. C. Dorlas, A. Ekert, A. Kay,

    and A. J. Landahl. Perfect transfer of arbitrary states inquantum spin networks. Phys. Rev. A, 71:032312, Mar2005.

    [8] J. I. Cirac, P. Zoller, H. J. Kimble, and H. Mabuchi.Quantum state transfer and entanglement distributionamong distant nodes in a quantum network. Phys. Rev.Lett., 78:3221–3224, Apr 1997.

    [9] E. Farhi and A. W. Harrow. Quantum supremacythrough the quantum approximate optimization algo-rithm. arXiv preprint arXiv:1602.07674, 2016.

    [10] E. Farhi, J. Goldstone, S. Gutmann, and M. Sipser.Quantum computation by adiabatic evolution. arXivpreprint quant-ph/0001106, 2000.

    [11] E. Farhi, J. Goldstone, S. Gutmann, J. Lapan, A. Lund-gren, and D. Preda. A quantum adiabatic evolution al-gorithm applied to random instances of an np-completeproblem. Science, 292(5516):472–475, 2001.

    http://arxiv.org/abs/1812.04170http://arxiv.org/abs/1602.07674http://arxiv.org/abs/quant-ph/0001106

  • 11

    1 2 3 4 5 6

    Circuit Depth p

    0.2

    0.4

    0.6

    0.8

    1.0M

    axim

    um

    ach

    ieva

    ble

    fid

    elit

    yF

    Maximum achievable fidelity F versus circuit depth pfor 2, 4, 6, 8 qubits

    N=2, tf=8

    N=4, tf=12

    N=6, tf=16

    N=8, tf=20

    (a) N = 2, 4, 6, 8 qubits.

    2 4 6 8 10 12 14

    Circuit depth p

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    Max

    imum

    achi

    evab

    lefid

    elit

    yF

    N=10, tf=25

    N=12, tf=30

    N=14, tf=34

    N=16, tf=39

    (b) N = 10, 12, 14, 16 qubits.

    FIG. 7. Maximum achievable fidelity F versus circuit depthp with fixed tf . The physical run time tf is chosen through ahyper-parameter grid search separately.

    [12] E. Farhi, J. Goldstone, and S. Gutmann. A quan-tum approximate optimization algorithm. arXiv preprintarXiv:1411.4028, 2014.

    [13] M. B. Hastings. Locality in quantum systems, volume 95.Oxford University Press, 2010.

    [14] U. L. Heras, A. Mezzacapo, L. Lamata, S. Filipp,A. Wallraff, and E. Solano. Digital quantum simulationof spin systems in superconducting circuits. Phys. Rev.Lett., 112:200501, May 2014.

    [15] W. W. Ho and T. H. Hsieh. Efficient variational simula-tion of non-trivial quantum states. SciPost Phys., 6:29,2019.

    [16] W. W. Ho, C. Jonay, and T. H. Hsieh. Ultrafast statepreparation via the quantum approximate optimizationalgorithm with long range interactions. arXiv preprintarXiv:1810.04817, 2018.

    [17] Z. Jiang, E. G. Rieffel, and Z. Wang. Near-optimal quan-tum circuit for grover’s unstructured search using a trans-verse field. Phys. Rev. A, 95:062317, Jun 2017.

    [18] H. J. Kimble. The quantum internet. Nature, 453(7198):1023, 2008.

    [19] V. Krotov. Global methods in optimal control theory, vol-ume 195. CRC Press, 1995.

    0 10 20 30 40Total run time tf

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    Max

    imu

    mac

    hie

    vab

    lefi

    del

    ityF

    Maximum achievable fidelity F versus total run time tffor 5, 10, 15, 19 qubits and p =2

    N=5

    N=10

    N=15

    N=19

    (a) Plots for p = 2 (uncontrollable).

    0 10 20 30 40

    Total run time tf

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    Max

    imum

    achi

    evab

    lefid

    elit

    yF

    N=5

    N=10

    N=15

    N=19

    (b) Plots for p = 9 (controllable).

    FIG. 8. Maximum achievable fidelity F versus physical runtime tf with a fixed circuit depth p for N = 5, 10, 15, 19qubits. The oscillating behavior is due to the low circuitdepth; such behaviour disappears when the circuit depth pgrows larger (equivalently, the system gets more controllable).

    [20] X. Li et al. Perfect quantum state transfer in a super-conducting qubit chain with parametrically tunable cou-plings. Phys. Rev. Applied, 10:054009, Nov 2018.

    [21] E. H. Lieb and D. W. Robinson. The finite group velocityof quantum spin systems. In Statistical Mechanics, pages425–431. Springer, 1972.

    [22] S. Lloyd. Quantum approximate optimization is com-putationally universal. arXiv preprint arXiv:1812.11075,2018.

    [23] M. A. Nielsen and I. Chuang. Quantum computation andquantum information. Cambridge University Press, 2011.

    [24] H. Pichler, S.-T. Wang, L. Zhou, S. Choi, and M. D.Lukin. Computational complexity of the rydberg block-ade in two dimensions. arXiv preprint arXiv:1809.04954,2018.

    [25] H. Pichler, S.-T. Wang, L. Zhou, S. Choi, and M. D.Lukin. Quantum optimization for maximum indepen-dent set using rydberg atom arrays. arXiv preprintarXiv:1808.10816, 2018.

    [26] H. A. Rabitz, M. M. Hsieh, and C. M. Rosenthal. Quan-tum optimally controlled transition landscapes. Science,303(5666):1998–2001, 2004.

    http://arxiv.org/abs/1411.4028http://arxiv.org/abs/1810.04817http://arxiv.org/abs/1812.11075http://arxiv.org/abs/1809.04954http://arxiv.org/abs/1808.10816

  • 12

    0 2 4 6 8

    Total run time tf

    0.0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6M

    axim

    umac

    hiev

    able

    fidel

    ity

    FN=4

    N=6

    N=8

    N=10

    N=12

    N=14

    N=16

    N=18

    N=20

    1. Exponentially suppressed 2. Exponential increasing

    3. Steady increasing

    FIG. 9. Maximum achievable fidelity F versus small phys-ical run time tf with a sufficiently large circuit depth p forN = 4, 6, 8, 10, 12, 14, 16, 18, 20 qubits. In general,we identify three different growing patterns: (1) exponen-tially suppressed region; (2) exponentially increasing region;(3) steady increasing region.

    [27] A. T. Rezakhani, A. K. Pimachev, and D. A. Lidar. Ac-curacy versus run time in an adiabatic quantum search.Phys. Rev. A, 82:052305, Nov 2010.

    [28] P. Richerme, Z.-X. Gong, A. Lee, C. Senko, J. Smith,M. Foss-Feig, S. Michalakis, A. V. Gorshkov, andC. Monroe. Non-local propagation of correlations inquantum systems with long-range interactions. Nature,511(7508):198–201, July 2014.

    [29] B. Russell, H. Rabitz, and R. Wu. Quantum controllandscapes are almost always trap free. arXiv preprintarXiv:1608.06198, 2016.

    [30] V. N. Smelyanskiy, K. Kechedzhi, S. Boixo, S. V.Isakov, H. Neven, and B. Altshuler. Non-ergodic delo-calized states for efficient population transfer within anarrow band of the energy landscape. arXiv preprintarXiv:1802.09542, 2018.

    [31] R. F. Stengel. Optimal control and estimation. CourierCorporation, 2012.

    [32] Z. Wang, S. Hadfield, Z. Jiang, and E. G. Rieffel. Quan-tum approximate optimization algorithm for maxcut: Afermionic view. Phys. Rev. A, 97:022304, Feb 2018.

    [33] R.-B. Wu, M. A. Hsieh, and H. Rabitz. Role of control-lability in optimizing quantum dynamics. Phys. Rev. A,83:062306, Jun 2011.

    [34] R.-B. Wu, R. Long, J. Dominy, T.-S. Ho, and H. Rabitz.Singularities of quantum control landscapes. Phys. Rev.A, 86:013405, Jul 2012.

    [35] Z.-C. Yang, A. Rahmani, A. Shabani, H. Neven, andC. Chamon. Optimizing variational quantum algorithmsusing pontryagin’s minimum principle. Phys. Rev. X, 7:021027, May 2017.

    [36] N. Y. Yao, L. Jiang, A. V. Gorshkov, Z.-X. Gong,A. Zhai, L.-M. Duan, and M. D. Lukin. Robust quantumstate transfer in random unpolarized spin chains. Phys.Rev. Lett., 106:040505, Jan 2011.

    [37] X.-P. Zhang, B. Shao, S. Hu, J. Zou, and L.-A. Wu. Opti-mal control of fast and high-fidelity quantum state trans-fer in spin-1/2 chains. Annals of Physics, 375:435–443,2016.

    2.5 5.0 7.5 10.0 12.5 15.0 17.5 20.0Total run time tf

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    Max

    imu

    mac

    hie

    vab

    lefi

    del

    ityF

    Maximum achievable fidelity F versus total run time tffor 2, 4, 6, 8 qubits

    N=2, p=7

    N=4, p=9

    N=6, p=11

    N=8, p=13

    (a) N = 2, 4, 6, 8 qubits.

    0 5 10 15 20 25 30 35 40Total run time tf

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    Max

    imu

    mac

    hie

    vab

    lefi

    del

    ityF

    Maximum achievable fidelity F versus total run time tffor 10, 12, 14, 16 qubits

    N=10, p=15

    N=12, p=17

    N=14, p=19

    N=16, p=23

    (b) N = 10, 12, 14, 16 qubits.

    FIG. 10. Maximum achievable fidelity F versus physical runtime tf with a sufficiently large circuit depth p.

    [38] L. Zhou, S.-T. Wang, S. Choi, H. Pichler, and M. D.Lukin. Quantum approximate optimization algorithm:Performance, mechanism, and implementation on near-term devices. arXiv preprint arXiv:1812.01041, 2018.

    http://arxiv.org/abs/1608.06198http://arxiv.org/abs/1802.09542http://arxiv.org/abs/1812.01041

  • 13

    2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19

    Number of qubits N

    0

    10

    20

    30

    40

    Min

    imu

    mt f

    Minimum tf required for achieveing F = 0.99 verse N

    Fit: tf = 2.439N − 4.661, r2=0.991data

    FIG. 11. Minimum required run time tf for achieving fidelityF = 0.99 versus the number of qubits (N = 2 → 19). TheLieb-Robinson bound gives a lower bound of approximatelytf = 0.03N + c.

    0 5 10 15 20 25Total run time tf

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    Max

    imu

    mac

    hie

    vab

    lefi

    del

    ityF

    Maximum achievable fidelity F versus circuit depth pfor 10 qubits and p=1, 4, 7, 10, 13

    N=10, p=1

    N=10, p=4

    N=10, p=7

    N=10, p=10

    N=10, p=13

    (a) N = 10 qubits.

    0 5 10 15 20 25 30Total run time tf

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    Max

    imu

    mac

    hie

    vab

    lefi

    del

    ityF

    Maximum achievable fidelity F versus circuit depth pfor 15 qubits and p=1, 4, 7, 10, 13, 16, 19

    N=15, p=1

    N=15, p=4

    N=15, p=7

    N=15, p=10

    N=15, p=13

    N=15, p=16

    N=15, p=19

    (b) N = 15 qubits.

    FIG. 12. Maximum achievable fidelity F versus physical runtime tf using QAOA as a function of different different circuitdepth p. The lines with oscillating behaviors are uncontrol-lable.

    2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    Number of qubits N

    0

    1

    2

    3

    4

    Exp

    onen

    tial

    lysu

    ppre

    ssed

    tim

    et s

    Fit: ts = 0.246N − 0.439, r2=0.997numerical results

    theoretical bound

    FIG. 13. Exponentially suppressed time tf (achieved fidelityF < 0.01) versus the number of qubits (N = 2 → 20). TheLieb-Robinson bound gives a bound approximately of tf ∼0.03N + c. The exponentially suppressed time is defined asthe time needed to make fidelity higher than 0.01.

  • 14

    Appendix A: Optimal control solution and Pontryagin’s maximum principle

    In this appendix, we first solve the dynamical equation of our system in the span of zero and single excitationsubspaces. Then we apply optimal control theory [19, 31] to help design our state transfer protocol.

    The dynamic of the system is governed by the Schrödinger’s equation (with ~ = 1):

    d|ψ(t)〉dt

    = −iĤ(t)|ψ(t)〉. (A1)

    Let |ψ(t)〉 = ∑Ni=1 Ci(t)|i〉, where Ci(t)s are the complex amplitudes of the wave function defined in computationalbasis. We choose ~c(t) = {C1(t), C2(t), · · · , CN (t)} as the dynamic variables for our problem. Substituting |ψ(t)〉 =∑Ni=1 Ci(t)|i〉 into Eq. (A1), we have

    N∑j=1

    d

    dtCj(t)|j〉 = −i

    {s(t)Ĥc + [1− s(t)]ĤB

    } N∑j=1

    Cj(t)|j〉

    = −i{s(t)|N〉〈N |+ [1− s(t)]

    N−1∑h=1

    [σxhσxh+1 + σ

    yhσ

    yh+1]

    }N∑j=1

    Cj(t)|j〉

    = −is(t)CN (t)|N〉+{−i[1− s(t)]

    N−1∑h=1

    [σxhσ

    xh+1 + σ

    yhσ

    yh+1

    ]} N∑j=1

    Cj(t)|j〉.

    (A2)

    Left multiplying both sides by 〈j|, we get

    d

    dtCj(t)〈j|j〉 = −is(t)CN 〈j|N〉+

    {−i[1− s(t)]〈j|

    N−1∑h=1

    [σxhσxh+1 + σ

    yhσ

    yh+1]

    }N∑k=1

    Ck(t)|k〉

    = −i{s(t)CN 〈j|N〉+ [1− s(t)]

    N∑k=1

    Bj,kCk

    },

    (A3)

    where

    Bj,k = 〈j|N−1∑h=1

    [σxhσxh+1 + σ

    yhσ

    yh+1]|k〉

    = 2δ(k − j − 1).(A4)

    Then we arrive in the dynamics equation in the form ~̇c(t) = f(~c(t), s(t)).

    d

    dtCj(t) = −i

    {s(t)CN 〈j|N〉+ [1− s(t)]

    N∑k=1

    Bj,kCk(t)

    }

    = −i{s(t)CNδjN + 2[1− s(t)]

    N∑k=1

    δ(k − j − 1)Ck}

    =

    N∑k=1

    Aj,kCk,

    (A5)

    where

    Aj,k = −i {s(t)δjN + 2[1− s(t)]δ(k − j − 1)} . (A6)The cost function (action) for our state transfer problem is given by

    J [~c(tf )] = −|CN (tf )|2, (A7)which only depends on the final state. Thus the problem we are solving is of Mayer type [31]. Then the controlHamiltonian is linearly dependent to the conjugate momentum ~p and control dynamics is

    Hcontrol = ~pT · f(~c(t), s(t)) = ~pT ·A · ~c, (A8)

  • 15

    such that the conjugate momentum is determined by the control Hamiltonian in the same way as that in the classicalmechanics:

    ~̇p = −∂~cHcontrol(~c, ~p, s). (A9)

    With the fixed total time tf , the boundary condition for conjugate variable is given by

    ~p(tf ) =∂J(t)

    ∂~c

    ∣∣∣∣t=tf

    . (A10)

    Denote the components of ~p to be Pi(t). Pi(t) should satisfy:

    d

    dtPj(t) = −is(t)PNδjN + 2[1− s(t)]Pj+1. (A11)

    Subsequently, the necessary and sufficient conditions for an optimal control s(t) is determined by:

    ∂Ĥ

    ∂s= 0,

    ∂2Ĥ

    ∂s2≥ 0. (A12)

    However, this cannot be applied to linear control problem where ∂Ĥ∂s is not a function of s. Pontryagin’s principlecomes to rescue, which replace two criteria with three new ones that are necessary and sufficient.

    Hcontrol(~c∗, ~p∗, s∗) ≤ Hcontrol(~c∗, ~p∗, s),∀t ∈ [0, tf ] (A13)

    ~p(tf ) = ∂~cJ [~c(tf )], (A14)

    ∂tJ [~c] +Hcontrol(~c∗, ~p∗, s∗)|t=tf = 0 (A15)

    Since the control Hamiltonian is a linear function of the control parameter s(t)(0 ≤ s(t) ≤ 1) , s(t) should bemaximized when ∂sHcontrol < 0 and should be minimized when ∂sHcontrol > 0. The optimal control for QAOA istherefore determined from Pontryagin’s theorem as follows:

    s(t) =

    {0 ∂Ĥcontrol∂s > 0

    1 ∂Ĥcontrol∂s < 0(A16)

    Then, the best control solution s(t) is of bang-bang form, which corresponds to switching between two constantcontrols for each time duration. The bang-bang form of control contains abrupt switch between two values of s(t) attime t0 determined by ∂sHcontrol|t=t0 = 0, or specifically as

    ~pT · F · ~c = 0, (A17)

    where the matrix elements of F are given by

    Fij = δiN − 2δ(j − i− 1). (A18)

    It is therefore trivial to verify whether a given control s(t) as a function of t is optimal or not. However, findingthe ‘optimal’ control is generally hard due to the mutual dependency of control and system dynamics. A brute forcesearch on switching time is already computationally formidable but is still unable to find the optimal control withoutspecifying the number of bangs.

  • 16

    Appendix B: Details of our numerical optimization

    In this appendix, we provide more details on our numerical optimizations. Table II summarizes the parameters tfand p we performed grid search on. In the first round of grid search (Run 1), we investigate the general performanceof optimized QAOAs with different fixed total run time tf and circuit depth p. To investigate the perfomance ofoptimized QAOA within or near the exponentially suppressed region, we performed the second round of grid searchwhich scanned more densely spaced total run time over a smaller interval. Table III overviews our numerical resultsand their implications.

    In these tables, we adopt the MATLAB style to represent arrays, for example, a ”0.2:0.2:1.6” corresponds to anarray of ”[0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6] and a ”1:5” corresponds an array of ”[1 2 3 4 5]”.

    TABLE I. Parameters of the first and the second round of grid searches over tf and p. The first round were used to determinethe general performance of optimized QAOAs. The second round were used to investigate the performance of optimized QAOAsnear or within the exponentially suppressed region that we have identified.

    Runs N 2 3 4 5 6 7 8 9 10 11

    Run 1: p 1:7 1:8 1:9 1:10 1:11 1:12 1:13 1:14 1:15 1:16

    tf 1:8 1:10 1:12 1:14 1:16 1:18 1:20 1:22 1:25 1:27

    Run 2: p 1:7 1:8 1:9 1:10 1:11 1:12 1:13 1:14 1:15 1:16

    tf 0.2:0.2:1.6 0.2:0.2:2.0 0.2:0.2:2.4 0.2:0.2:2.8 0.2:0.2:3.2 0.2:0.2:3.6 0.2:0.2:4.0 0.2:0.2:4.4 0.2:0.2:5.0 0.2:0.2:5.4

    N 12 13 14 15 16 17 18 19 20

    Run 1: p 1:17 1:18 1:19 1:20 1:2:23 1:2:25 1:2:27 1:2:29 1:2:31

    tf 1:30 1:32 1:34 1:36 1:2:39 1:2:41 1:2:43 1:2:45 1:2:47

    Run 2: p 1:17 1:18 1:19 1:20 1:2:23 1:2:25 1:2:27 1:2:29 1:2:31

    tf 0.2:0.2:6.0 0.2:0.2:6.4 0.2:0.2:6.8 0.2:0.2:7.2 0.2:0.4:7.8 0.2:0.4:8.2 0.2:0.4:8.6 0.2:0.4:9.0 0.2:0.4:9.4

    TABLE II. Parameters of the grid search over p with no constraints on tf . The obtained numerical results are presented andanalyzed in Section V B.

    N 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    p 1:7 1:8 1:9 1:10 1:11 1:12 1:13 1:14 1:15 1:16 1:17 1:18 1:19 1:20 1:22 1:24 1:26 1:28 1:30

    TABLE III. Overview of our numerical results. We consider the performance of optimized QAOAs with unlimited tf inSection V B and that with fixed tf in Section V C. In Section V C 1, we study the maximum achievable fidelity F as a functionof circuit depth p , and investigated the controllability of QAOAs. In Section V C 2, we study the maximum achievable fidelityF as a function of total run time tf , and investigate the Lieb-Robinson type quantum speed limit emerged in QAOAs.

    Figures Max achi. F Physical run time tf Circuit depth p Number of qubits N Goal

    Section V BFigure 2 Max achi. F unlimited tf (a): p =[1:6]; (b): p =[1:11] N = 2→ 20 QAOAFigure 4 N/A unlimited tf p = 15 N=10 bang-bang sol.

    unlimited tf Figure 5 z variable unlimited tf p = 2, 4 N = 3 Landscape

    Section V C 1

    Figure 6(a) y variable tf = 6 x variable: [1:9] [5,10,15,19] Controllability

    Figure 6(b) y variable tf = 13 x variable: [1:9] [5,10,15,19]

    Figure 7(a) y variable sufficiently large x variable: [1:6] [2,4,6,8]

    Fixed tf Figure 7(b) y variable sufficiently large x variable: [1:15] [10,12,14,16]

    Section V C 2

    Figure 8(a) y variable x variable: all p = 2 [5,10,15,19] LR-bound

    Figure 8(b) y variable x variable: all p = 9 [5,10,15,19]

    Figure 10(a) y variable x variable: all sufficiently large [2,4,6,8]

    Figure 10(b) y variable x variable: all sufficiently large [10,12,14,16]

    Fixed p Figure 12(a) y variable x variable: [1:25] sufficiently large N = 10

    Figure 12(b) y variable x variable: [1:30] sufficiently large N = 15

    Figure 11 F > 0.99 y variable sufficiently large x variable: [2:20]

    Figure 13 F < 0.01 y variable sufficiently large x variable: [2:20]

    Optimizing QAOA: Success Probability and Runtime Dependence on Circuit DepthAbstractI IntroductionII QAOA for State TransferIII Success Probability Scaling as a Function of Circuit DepthIV Quantum Speed LimitV Numerical Optimization of the QAOAA Numerical methodsB Numerical results for unlimited tf 1 Maximum achievable fidelity versus circuit depth p2 Control landscape

    C Numerical results for fixed tf: optimized QAOA1 Maximum achievable fidelity F versus circuit depth p2 Maximum achievable fidelity F versus physical run time tf

    VI Summary ReferencesA Optimal control solution and Pontryagin's maximum principleB Details of our numerical optimization