arxiv:1409.3204v2 [quant-ph] 19 feb 2015 · 2018. 10. 3. · arxiv:1409.3204v2 [quant-ph] 19 feb...

5
Solution to the quantum Zermelo navigation problem Dorje C. Brody 1,2 and David M. Meier 1 1 Department of Mathematics, Brunel University, Uxbridge UB8 3PH, UK 2 St Petersburg State University of Information Technologies, Mechanics and Optics, Kronwerkskii ave 49, St Petersburg 197101, Russia (Dated: September 11, 2014) The solution to the problem of finding a time-optimal control Hamiltonian to generate a given unitary gate, in an environment in which there exists an uncontrollable ambient Hamiltonian (e.g., a background field), is obtained. In the classical context, finding the time-optimal way to steer a ship in the presence of a background wind or current is known as the Zermelo navigation problem, whose solution can be obtained by working out geodesic curves on a space equipped with a Randers metric. The solution to the quantum Zermelo problem, which is shown here to take a remarkably simple form, is likewise obtained by finding explicit solutions to the geodesic equations of motion associated with a Randers metric on the space of unitary operators. The result reveals that the optimal control in a sense ‘goes along with the wind’. PACS numbers: 03.67.Ac, 42.50.Dv, 02.30.Xx The problem of finding the optimal Hamiltonian for processing a given quantum state, or implementing a quantum operation (gate), in shortest possible time sub- ject to certain constraints, has attracted considerable at- tention over the past decade [1–8]. Broadly speaking, the task can be classified into two closely-related categories: (a) transforming one quantum state into another; and (b) transforming one unitary operator into another, in the shortest possible time. If the constraint is concerned merely with a limit on energy resource, then the opti- mal Hamiltonian is time independent, and can be found easily by noting that under a unitary motion, the short- est path coincides with the path along which the speed of evolution is also maximised [9, 10]. If there are fur- ther constraints, for example, the choice of the Hamilto- nian is limited, then often a time-dependent Hamiltonian that minimises an action has to be determined by vari- ational approaches [11, 12]. Finding a solution to such a variational problem is in general difficult, however, an efficient numerical regularisation scheme to obtain an ap- proximate solution has been proposed more recently [13]. For many problems related to controlling quantum sys- tems considered in the literature, it is assumed that the experimentalist has full control over the allowable range of Hamiltonians within the constraint, whereas in a lab- oratory there can often be situations in which the sys- tem is immersed in an external field or potential that is beyond control (e.g., gravitational or electro-magnetic field), since a complete elimination of external fields in a laboratory can be prohibitively expensive for the given task. Evidently, this is a generic issue that needs to be addressed adequately to be able to accurately implement a rapid quantum processing. In the present paper we address this issue by finding the time-optimal control Hamiltonian ˆ H(t)= ˆ H 0 + ˆ H 1 (t) that generates a unitary motion to transform one uni- tary operator ˆ U I into another operator ˆ U F , subject to the constraints (i) that the background Hamiltonian ˆ H 0 cannot be controlled; (ii) that the control Hamiltonian fulfils the energy resource bound of the form tr(H 2 1 )=1 at all time; and (iii) that the background Hamiltonian is not dominating in the sense that tr( ˆ H 2 0 ) < tr( ˆ H 2 1 ). This is the quantum counterpart of a well-known classical navigation problem posed by Zermelo: given the present location in the ocean, with a given wind and/or current distribution characterised by a location-dependent vector field, one wishes to find the optimal control of the ves- sel so as to reach the destination in the shortest possible time [14, 15]. The vector field generated by the reference Hamiltonian ˆ H 0 can be thought of as representing the background wind or current, whereas ˆ H 1 determines the control. In the classical context, it was observed by Shen [16] that the solution to the Zermelo navigation problem can be obtained by finding the geodesic curves associated with a Randers metric on the configuration space. Mo- tivated by this result, more recently Russell & Stepney [17] introduced the quantum Zermelo navigation problem stated above, and analysed the shortest time required to realise the transformation ˆ U I ˆ U F . Their observation that quantum Zermelo navigation problems can be solved by finding the geodesics of Randers metrics opens up the possibility of addressing a wide range of realistic quan- tum control problems where the environmental influence cannot be eliminated. However, analyses involving Ran- ders spaces are generally difficult, and finding solutions to the geodesic equations is not straightforward [18, 19]. Indeed, the only examples considered in [17] concern the time-independent cases, while the optimal navigation is realised by a time-independent Hamiltonian only if the background Hamiltonian ˆ H 0 happens to be the one that realises the operation ˆ U I ˆ U F . But since ˆ U I and ˆ U F are arbitrary given unitary gates one wishes to implement, such a scenario will not prevail in real laboratories. arXiv:1409.3204v2 [quant-ph] 19 Feb 2015

Upload: others

Post on 04-Feb-2021

0 views

Category:

Documents


0 download

TRANSCRIPT

  • Solution to the quantum Zermelo navigation problem

    Dorje C. Brody1,2 and David M. Meier1

    1Department of Mathematics, Brunel University, Uxbridge UB8 3PH, UK2St Petersburg State University of Information Technologies, Mechanics and Optics,

    Kronwerkskii ave 49, St Petersburg 197101, Russia(Dated: September 11, 2014)

    The solution to the problem of finding a time-optimal control Hamiltonian to generate a givenunitary gate, in an environment in which there exists an uncontrollable ambient Hamiltonian (e.g.,a background field), is obtained. In the classical context, finding the time-optimal way to steer aship in the presence of a background wind or current is known as the Zermelo navigation problem,whose solution can be obtained by working out geodesic curves on a space equipped with a Randersmetric. The solution to the quantum Zermelo problem, which is shown here to take a remarkablysimple form, is likewise obtained by finding explicit solutions to the geodesic equations of motionassociated with a Randers metric on the space of unitary operators. The result reveals that theoptimal control in a sense ‘goes along with the wind’.

    PACS numbers: 03.67.Ac, 42.50.Dv, 02.30.Xx

    The problem of finding the optimal Hamiltonian forprocessing a given quantum state, or implementing aquantum operation (gate), in shortest possible time sub-ject to certain constraints, has attracted considerable at-tention over the past decade [1–8]. Broadly speaking, thetask can be classified into two closely-related categories:(a) transforming one quantum state into another; and(b) transforming one unitary operator into another, inthe shortest possible time. If the constraint is concernedmerely with a limit on energy resource, then the opti-mal Hamiltonian is time independent, and can be foundeasily by noting that under a unitary motion, the short-est path coincides with the path along which the speedof evolution is also maximised [9, 10]. If there are fur-ther constraints, for example, the choice of the Hamilto-nian is limited, then often a time-dependent Hamiltonianthat minimises an action has to be determined by vari-ational approaches [11, 12]. Finding a solution to sucha variational problem is in general difficult, however, anefficient numerical regularisation scheme to obtain an ap-proximate solution has been proposed more recently [13].

    For many problems related to controlling quantum sys-tems considered in the literature, it is assumed that theexperimentalist has full control over the allowable rangeof Hamiltonians within the constraint, whereas in a lab-oratory there can often be situations in which the sys-tem is immersed in an external field or potential thatis beyond control (e.g., gravitational or electro-magneticfield), since a complete elimination of external fields ina laboratory can be prohibitively expensive for the giventask. Evidently, this is a generic issue that needs to beaddressed adequately to be able to accurately implementa rapid quantum processing.

    In the present paper we address this issue by findingthe time-optimal control Hamiltonian Ĥ(t) = Ĥ0+Ĥ1(t)that generates a unitary motion to transform one uni-tary operator ÛI into another operator ÛF , subject to

    the constraints (i) that the background Hamiltonian Ĥ0cannot be controlled; (ii) that the control Hamiltonianfulfils the energy resource bound of the form tr(H21 ) = 1at all time; and (iii) that the background Hamiltonianis not dominating in the sense that tr(Ĥ20 ) < tr(Ĥ

    21 ).

    This is the quantum counterpart of a well-known classicalnavigation problem posed by Zermelo: given the presentlocation in the ocean, with a given wind and/or currentdistribution characterised by a location-dependent vectorfield, one wishes to find the optimal control of the ves-sel so as to reach the destination in the shortest possibletime [14, 15]. The vector field generated by the referenceHamiltonian Ĥ0 can be thought of as representing thebackground wind or current, whereas Ĥ1 determines thecontrol.

    In the classical context, it was observed by Shen [16]that the solution to the Zermelo navigation problem canbe obtained by finding the geodesic curves associatedwith a Randers metric on the configuration space. Mo-tivated by this result, more recently Russell & Stepney[17] introduced the quantum Zermelo navigation problemstated above, and analysed the shortest time required torealise the transformation ÛI → ÛF . Their observationthat quantum Zermelo navigation problems can be solvedby finding the geodesics of Randers metrics opens up thepossibility of addressing a wide range of realistic quan-tum control problems where the environmental influencecannot be eliminated. However, analyses involving Ran-ders spaces are generally difficult, and finding solutionsto the geodesic equations is not straightforward [18, 19].Indeed, the only examples considered in [17] concern thetime-independent cases, while the optimal navigation isrealised by a time-independent Hamiltonian only if thebackground Hamiltonian Ĥ0 happens to be the one thatrealises the operation ÛI → ÛF . But since ÛI and ÛF arearbitrary given unitary gates one wishes to implement,such a scenario will not prevail in real laboratories.

    arX

    iv:1

    409.

    3204

    v2 [

    quan

    t-ph

    ] 1

    9 Fe

    b 20

    15

  • 2

    Here we solve this problem by deriving the Euler-Poincaré equation of motion for the control HamiltonianĤ1(t), and obtain the solution in closed form. Remark-ably, we find that the solution to the quantum Zermelonavigation problem takes the simple form:

    Ĥ1(t) = e−iĤ0tĤ1(0)e

    iĤ0t, (1)

    where Ĥ1(0) is the initial condition such that the actiongenerated by the total Hamiltonian Ĥ(t) = Ĥ0 + Ĥ1(t)takes ÛI to ÛF in shortest possible time. Thus, the op-timal control is obtained by finding the initial directionH1(0) for the manoeuvre and drift along the ‘wind’ Ĥ0.We then provide a scheme for finding the initial con-dition Ĥ1(0). The results are illustrated in terms of aspin- 12 system. We shall also indicate how the analysispresented here can be applied to situations where thereare further constraints on the control Hamiltonian.

    Since the mathematical machinery required for solvingthe navigation problem is perhaps not widely accessibleto the broader physics community, we begin with a briefdiscussion of the background concepts before proceedingto derive (1). To address such navigation problems in thecalculus of variation, it is often the case that one requiresthe notion of a distance that depends not only on thelocation but also on the direction—a concept that goesoutside of the realm of Riemannian geometry. Specifi-cally, for a given curve xi(t) on the configuration spaceM, equipped with a Riemannian metric, we consider theintegral of the form

    T =

    ∫ t1t0

    F (xi, ẋi)dt (2)

    for some positive function F , which is assumed to behomogeneous of first degree in ẋi, F (xi, λẋi) = λF (xi, ẋi)for any λ > 0, so that T is independent of the choice ofthe parameter t along xi(t). Thus, F (xi, ẋi) defines, foreach fixed point xi ∈M, a distance on the tangent spaceof M. In particular, the level surface F (x, ẋ) = 1 on thetangent space of M at x defines the indicatrix [15]. Nowfor a fixed x and an arbitrary point ξ on the tangent space

    of M at x, the ray−→xξ clearly intersects the indicatrix at

    a point ρξ. Thus, conversely, for each point ξ if we definea function F according to F (ξ) = |ξ|/|ρξ|, where | · |denotes the Euclidean norm, then we can introduce ametric, known as the Minkowski metric [20], as follows:For ξ, ξ′ on the tangent space of M at x the distancebetween these points is defined by D(ξ, ξ′) = F (ξ − ξ′).In particular, the metric tensor defined on M induced bythe distance function D associated with the fundamentalfunction F can be expressed in the form:

    gij(x, ẋ) =1

    2

    ∂2

    ∂ẋi∂ẋjF 2(xi, ẋi), (3)

    and we have F 2 = gij ẋiẋj .

    In classical mechanics, often the fundamental functiontakes the form of the kinetic energy: F 2 = γij(x)ẋ

    iẋj .Thus, the resulting metric gij(x, ẋ) = γij(x) is indepen-dent of the direction ẋ, i.e. it defines a Riemannian met-ric on M, since the indicatrix is just a sphere. In manyproblems with engineering applications, such as a nav-igation problem, however, the relevant function takes adifferent form, and as such one is required to go beyondthe techniques of Riemannian geometry. Realising this,Carathéodory suggested to his then PhD student Finslerto investigate the geometry of spaces equipped with suchdirection-dependent metrics [21]. Subsequently, spacesendowed with locally Minkowski metrics were referred toas Finsler spaces [22].

    Let us now turn to the classical Zermelo navigationproblem of reaching a target on a manifold M equippedwith a Riemannian metric hij in the shortest possibletime, in the presence of background wind wi. The analy-sis of the problem simplifies if we observe that it sufficesto find the locally optimal solution on the tangent space[16]. Specifically, for any vector ~ξ on the tangent space

    to M at x we can regard |~ξ|h =√hijξiξj as represent-

    ing the time it takes to reach the endpoint of ~ξ. Nowsuppose that in the absence of wind the time it takes toreach the destination ~u at full throttle is 1 in a suitableunit (e.g., second), i.e. |~u|h = 1. In the presence of wind,with |~w|h < 1, however, after a journey of one second atfull throttle the vessel will reach the point ~v = ~u + ~w,instead of the destination ~u. In other words, the unitsphere |~u|h = 1 has been displaced by the wind, but since|~w|h < 1 by assumption, the centre point x remains inthe interior of the sphere. Therefore, for any vector ~ξ on

    the tangent space the ray−→xξ intersects the indicatrix at

    a point ρξ; working out the Euclidean norms of ~ξ and ~ρξand taking the ratio, a short calculation shows that thefundamental function takes the form (see also [23, 24]):

    F (x, ξ) =

    √〈~w, ~ξ〉2h + |~ξ|2h(1− |~w|2h)− 〈~w, ~ξ〉h

    1− |~w|2h, (4)

    where |~ξ|2h = hijξiξj and 〈~w, ~ξ〉h = hijwiξj . Makinguse of (3), an explicit form of the metric on M can beobtained. The calculation simplifies if one writes

    αij =hij

    1− |~w|2h+

    wiwj(1− |~w|2h)2

    , βi = −wi

    1− |~w|2h, (5)

    where wi = hijwj , so that we have F =

    √αijξiξj +βiξ

    i.The solution curves to the Zermelo navigation problemare then found by working out the geodesics of the metric.

    We remark that the metric of the type√αijξiξj +βiξ

    i

    was introduced by Randers in the context of a unified the-ory of gravitation and electromagnetism [25]. However,Randers was unaware of the Finslerian nature of the met-ric, and attempted to interpret it in the Riemannian sensein the context of a five-dimensional Kaluza-Klein theory.

  • 3

    Randers metrics are perhaps the most commonly inves-tigated Finsler metrics in physical applications such asthe electron microscope [26] and in propagation of soundand light rays in a moving medium [24, 27, 28].

    The relevance of Finsler geometry to problems in quan-tum control has been observed in [29, 30]. In the presenceof background fields, more recentl Russell & Stepney [17]proposed the technique of Shen [16] to be applied to themanifold M of special unitary matrices endowed with thebi-invariant trace norm. Specifically, working with the el-ements of the Lie algebra ξ̂, ξ̂′ ∈ su(N) we have

    〈ξ̂, ξ̂′〉h = tr(ξ̂†ξ̂′). (6)

    With this setup we wish to minimise the journey time(2) in the presence of ‘wind’ given in su(N) by −iĤ0,when ξ̂ = −iĤ(t) = −i(Ĥ0 + Ĥ1(t)). The fundamentalfunction (4) in this quantum context thus reads

    F (ξ̂) = i

    √[tr(ξ̂Ĥ0)]2 + tr(ξ̂2)(1− tr(Ĥ20 ))− tr(ξ̂Ĥ0)

    1− tr(Ĥ20 ),(7)

    which is just the Finslerian norm ‖ξ̂‖.To proceed we find it convenient to minimise the ki-

    netic energy 12∫F 2dt along the path, instead of

    ∫Fdt.

    It should be evident that the optimal path ξ̂(t) thatminimises the latter also minimises the former. WritingF 2(ξ̂) = ‖ξ̂‖2 we have

    δ‖ξ̂‖2 =

    〈δ‖ξ̂‖2

    δξ̂, δξ̂

    〉= 2‖ξ̂‖

    〈δ‖ξ̂‖δξ̂

    , δξ̂

    〉, (8)

    where we have written, for any ν̂ ∈ su(N) and any f(ξ̂),〈δf(ξ̂)

    δξ̂, ν̂

    〉=

    d

    d�f(ξ̂ + �ν̂)

    ∣∣∣∣�=0

    , (9)

    and on account of (7) we have〈δ‖ξ̂‖δξ̂

    , ν̂

    〉= −i tr(ν̂Ĥ0)

    1− tr(Ĥ20 )

    +itr(ξ̂Ĥ0)tr(ν̂Ĥ0) + (1− tr(Ĥ20 ))tr(ξ̂ν̂)

    (1− tr(Ĥ20 ))√

    [tr(ξ̂Ĥ0)]2 + tr(ξ̂2)(1− tr(Ĥ20 )). (10)

    Our aim is to solve

    0 = δ

    (1

    2

    ∫ 10

    ‖ξ̂‖2dt)

    =

    ∫ 10

    ‖ξ̂‖

    〈δ‖ξ̂‖δξ̂

    , δξ̂

    〉dt (11)

    with fixed end points of the curve on SU(N). The con-straints on the end points restricts admissible variationsδξ̂. In particular, a standard result of Euler–Poincaréreduction [31] asserts that

    δξ̂ = ˙̂η − [ξ̂, η̂], (12)

    where η̂ is an arbitrary curve in su(N) with η̂(0) = η̂(1) =0. Substituting (12) and (10) in (11) and rearrangingterms, we are thus led to the relation:

    0 = −∂t(‖ξ̂‖)Ĥ0 − ‖ξ̂‖[Ĥ0, ξ̂]

    +∂t

    (‖ξ̂‖√· · ·

    tr(ξ̂Ĥ0)

    )Ĥ0 +

    ‖ξ̂‖√· · ·

    tr(ξ̂Ĥ0)[Ĥ0, ξ̂]

    +(1− tr(Ĥ20 )) ∂t

    (‖ξ̂‖ξ̂√· · ·

    ), (13)

    where we have written√· · · for the square-root term ap-

    pearing in the numerator of (7). This result appears un-duly complicated, however, if we take note of the factthat we are interested in the quantum navigation at fullthrottle, i.e. ‖ξ̂‖ = 1, then by taking the relevant timederivatives in (13) we deduce the Euler-Poincaré equa-

    tion of the form:˙̂ξ+i[Ĥ0, ξ̂]−(iĤ0+ ξ̂)tr( ˙̂ξĤ0)/

    √· · · = 0.

    Substituting ξ̂ = −i(Ĥ0 + Ĥ1(t)) in here we thus obtainthe relevant equation of motion for the control Hamilto-

    nian Ĥ1(t): −i ˙̂H1 + [Ĥ0, Ĥ1] + Ĥ1tr(Ĥ0 ˙̂H1)/√· · · = 0. If

    we eliminate the square-root term using (7) along with

    F (ξ̂) = ‖ξ̂‖ = 1, which gives us i√· · · = 1 + tr(Ĥ0Ĥ1),

    then we deduce that

    ˙̂H1 + i[Ĥ0, Ĥ1]−

    Ĥ1

    1 + tr(Ĥ1Ĥ0)tr(Ĥ0

    ˙̂H1) = 0, (14)

    where we have made use of the constraint that tr(Ĥ21 ) =1. Multiplying (14) with Ĥ0 and taking the trace, we thus

    deduce that tr(Ĥ0˙̂H1) = 0. We therefore conclude from

    (14) that the quantum Zermelo–Euler–Poincaré equationtakes the simple form:

    ˙̂H1 + i[Ĥ0, Ĥ1] = 0. (15)

    This, however, is just the equation for a co-adjoint mo-tion, so it can be solved, with the solution (1).

    It is interesting to observe that, after some lengthybut straightforward algebra, we are led to a simple andintuitive solution to the quantum navigation problem,namely, that we must pick the initial direction Ĥ1(0)and let it be advected by the prevailing field Ĥ0. Tohit the right target ÛF starting from the initial pointÛI , however, the initial direction Ĥ1(0) has to be chosenappropriately. In what follows we shall derive an ordinarydifferential equation satisfied by the initial direction.

    We proceed by first solving the navigation problemin the absence of the wind: Ĥ0 = 0. In this case, theoptimal control Ĥ1 is time independent, and the initialcondition Ĥ1(0) can thus be obtained by taking the ma-trix logarithm of ÛF Û

    −1I . The idea behind our scheme

    is to gradually increase Ĥ0 from zero to the level speci-fied by the problem, while calculating, for each incrementof Ĥ0, the optimal control Hamiltonian that solves the

  • 4

    Zermelo problem with that wind. Clearly, as Ĥ0 is in-creased, Ĥ1(0) has to be adjusted as well, or else the tar-get gate will be missed. Moreover, the trajectory mighttake slightly more or slightly less time. Hence, the dura-tion of the trajectory needs also be adapted.

    With this in mind, let us calculate how the final gatevaries when the wind, the initial control, and the terminaltime are adjusted infinitesimally. Let Û(t) be a curve in

    SU(N) starting at ÛI satisfying ∂tÛ = ξ̂Û for some curve

    ξ̂ in su(N), and fix a time s. If ξ̂(t)+ �δξ̂(t) is a variation

    of ξ̂, then Û(s) varies as

    δÛ(s) = Û(s)

    ∫ s0

    Û(t)−1δξ̂(t)Û(t) dt. (16)

    This follows from adapting Lemma 2.4 of [32] to thepresent context. To proceed, let us write Ĥ1(0, λ) forthe optimal initial control and Tλ for the duration of thetrajectory when the wind is given by λĤ0, λ ∈ [0, 1]. Letus further denote by Ûλ(t) the corresponding geodesiccurve in SU(N). In what follows we shall write deriva-tives with respect to λ as T ′, Û ′λ, and so on. Notice

    that Ûλ(Tλ) equals the target gate ÛF for all λ. Hence,Û ′λ(Tλ) = 0. Using (16), we thus obtain

    0 = Û−1λ (Tλ)Û′λ(Tλ)

    =

    ∫ Tλ0

    Ûλ(t)−1ξ̂′λ(t)Ûλ(t) dt+ T

    ′λÛ−1λ (Tλ)ξ̂λ(Tλ)Ûλ(Tλ).

    (17)

    Recall that ξ̂λ(t) = −i(λĤ0 + Ĥ1(t, λ)), where Ĥ1(t, λ) isgiven by (1). Therefore, upon differentiation, ξ̂′λ(t) =

    −ie−iĤ0λt(Ĥ0 + it[Ĥ1(0, λ), Ĥ0] + Ĥ ′1(0, λ))eiĤ0λt, fromwhich it follows that (17) is a linear equation in T ′λ and

    Ĥ ′1(0, λ), admitting a unique solution for each λ once thelinear constraint tr(Ĥ1(0, λ)Ĥ

    ′1(0, λ)) = 0 is taken into

    account. Finally, T ′λ and Ĥ′1(0, λ) can be integrated up

    to λ = 1 starting from the wind-free solution λ = 0. Theoptimal initial control is then given by Ĥ1(0, 1), and thetrajectory is traversed in time T1.

    In summary, we have derived the Euler-Poincaré equa-tion (15) associated with the quantum Zermelo naviga-tion problem introduced in [17]. The equation of motionis surprisingly simple, and admits an elementary solution(1). We have provided a scheme which allows for the de-termination of the initial control Hamiltonian Ĥ1(0) re-quired to hit the correct target point ÛF . On account oflinearity, our scheme can easily be implemented in prac-tice. With the solution (1) at hand, optimal quantumcontrol with finite energy resources becomes feasible un-der the presence of external field or potential that mightbe difficult to eliminate in laboratories. As an illustra-tive example let us consider the control of a spin- 12 sys-

    tem, where the objective is to transform ÛI = 1 intoÛF = e

    −iπσ̂x/2 = −iσ̂x, under the influence of an external

    ÛI•ÛF•x

    z

    y

    FIG. 1: Optimal generation of target unitary gate. Thetime-optimal trajectories Û(t) are shown for various wind-strengths ω = 0, 0.25, 0.5, 0.75, 1, as curves in the rotationgroup using the standard covering map. The centre of thesphere corresponds to the initial gate ÛI = 1, while the ter-minal point that lies on the surface of the sphere is the targetgate ÛF = −iσ̂x. The direction of the vector joining the cen-tre 1 to a point Û(t) on a given curve represents the axis ofrotation, whereas the radius of the vector represents the angleof rotation. The sphere upon which the target gate lies thushas radius π.

    field Ĥ0 = −ωσ̂z, where σ̂x, σ̂y, σ̂z are the Pauli matrices.In this example a closed-form expression for the optimalinitial Hamiltonian Ĥ1(0) can be obtained on account of

    the relation (cf. [33, 34]) ÛF = −iσ̂x = eiωσ̂zT e−iĤ1(0)T ,which follows from (1). Specifically, a short calculationshows that Ĥ1(0) =

    1√2n · σ̂ and T = π/

    √2, where the

    unit vector n is given by n = (cos(ωT ), sin(ωT ), 0). Theresulting unitary orbit Û(t) is sketched in figure 1 for arange of values of ω.

    We conclude by remarking that in the presence of addi-tional constraints on the control Hamiltonian that limitthe implementability of (1), it suffices to include themin the maximisation of F 2 by use of Lagrange multipli-ers. It then follows that the solution (1) remains valid,except that the initial control H1(0) is replaced by atime-dependent one (cf. [34]). More precisely, what thesolution (1) shows is that it is possible to switch to aframe that moves in the counter direction to the wind sothat the analysis of constrained optimisation performed,for example in [11], with time-dependent constraints, be-comes applicable. In this manner the solution to theZermelo navigation problem presented here can be ex-tended straightforwardly to accommodate further con-straints that one might encounter for instance in systemsinvolving a large number of coupled spins where control-lable degrees of freedom are typically limited.

  • 5

    We thank Gary Gibbons for drawing our attention to[17, 24, 28].

    Note added: Russell & Stepney have independentlyobtained the solution (1) to the quantum Zermelo navi-gation problem [33], using a theorem of [35] on geodesicsof Randers spaces (rather than deriving and solving theEuler–Poincaré equation as we have done here).

    [1] Khaneja, N., Glaser, S. J. & Brockett, R. 2002 Sub-Riemannian geometry and time optimal control of threespin systems: Quantum gates and coherence transfer.Phys. Rev. A65, 032301.

    [2] Schulte-Herbrüggen, T., Spörl, A., Khaneja, N. &Glaser, S. J. 2005 Optimal control-based efficient synthe-sis of building blocks of quantum algorithms: A perspec-tive from network complexity towards time complexity.Phys. Rev. A72, 042331.

    [3] Rezakhani, A. T., Kuo, W. J., Hamma, A., Lidar, D. A.& Zanardi, P. 2009 Quantum adiabatic brachistochrone.Phys. Rev. Lett. 103, 080502.

    [4] Caneva, T., Murphy, M., Calarco, T., Fazio, R., Mon-tangero, S., Giovannetti, V. & Santoro, G. E. 2009 Opti-mal control at the quantum speed limit. Phys. Rev. Lett.103, 240501.

    [5] Rojo, A. G. & Bloch, A. M. 2010 The rolling sphere, thequantum spin and a simple view of the Landau-Zenerproblem. Ammer. J. Phys. 78, 1012.

    [6] Lee, K. Y. & Chau, H. F. 2013 Relation between quantumspeed limits and metrics on U(n). J. Phys. A46, 015305.

    [7] Hegerfeldt, G. C. 2013 Driving at the quantum speedlimit: Optimal control of a two-level system. Phys. Rev.Lett. 111, 260501.

    [8] Garon, A., Glaser, S. J. & Sugny, D. 2013 Time-optimalcontrol of SU(2) quantum operations. Phys. Rev. A88,043422.

    [9] Brody, D. C. 2003 Elementary derivation for passagetime. J. Phys. A36, 5587.

    [10] Brody, D. C. & Hook, D. W. 2006 On optimum Hamil-tonians for state transformations. J. Phys. A39, L167.

    [11] Carlini, A., Hosoya, A., Koike, T. & Okudaira, Y. 2006Time-optimal quantum evolution. Phys. Rev. Lett. 96,060503.

    [12] Carlini, A., Hosoya, A., Koike, T. & Okudaira, Y.2007 Time-optimal unitary operations. Phys. Rev. A75,042308.

    [13] Wang, X., Allegra, A., Jacobs, K., Lloyd, S., Lupo, C.& Mohseni, M. 2014 Quantum brachistochrone curves asgeodesics: obtaining accurate control protocols for time-optimal quantum gates. arXiv:1408.2465

    [14] Zermelo, E. 1931 Über das Navigationsproblem beiruhender oder veränderlicher Windverteilung. Ztschr. f.

    angew. Math. und Mech. 11, 114.[15] Carathéodory, C. 1935 Variationsrechnung und Par-

    tielle Differentialgleichungen erster Ordnung. (Berlin:B. G. Teubner).

    [16] Shen, Z. 2003 Finsler metrics with K=0 and S=0. Canad.J. Math. 55, 112-132.

    [17] Russell, B. & Stepney, S. 2014 Zermelo navigation anda speed limit to quantum information processing. Phys.

    Rev. A90, 012303.[18] Yasuda, H. & Shimada, H. 1977 On Randers spaces of

    scalar curvature. Rep. Math. Phys. 11, 347.[19] Bao, D., Chern, S. S. & Shen, Z. 2000 An Introduction

    to Riemann-Finsler Geometry. (Berlin: Springer).[20] Busemann, H. 1942 Metric Methods in Finsler Spaces and

    in the Foundations of Geometry. (Princeton: PrincetonUniversity Press).

    [21] Finsler, P. 1918 Über Kurven und Flächen in allgemeinenRäumen. PhD Dissertation, Göttingen.

    [22] Rund, H. 1959 The Differential Geometry of FinslerSpaces (Berlin: Springer).

    [23] Bao, D., Robles, C. & Shen, Z. 2004 Zermelo navigationon Riemannian manifolds. J. Diff. Geom. 66, 377.

    [24] Gibbons, G. W. & Warnick, C. M. 2011 The geometry ofsound rays in a wind. Contemp. Phys. 52, 197.

    [25] Randers, G. 1941 On an asymmetrical metric in the four-space of general relativity. Phys. Rev. 59, 195.

    [26] Ingarden, R. S. 1957 On the geometrically absolute opti-cal representation in the electron microscope. Trav. Soc.Sci. Lett. Wrocaw B45, 1.

    [27] Luneburg, R. K. 1964 Mathematical Theory of Optics.(Berkeley: University of California Press).

    [28] Gibbons, G. W., Herdeiro, C. A. R., Warnick, C. M.& Werner, M. C. 2009 Stationary metrics and opti-cal Zermelo-Randers-Finsler geometry. Phys. Rev. D79,044022.

    [29] Nielsen, M. A. 2006 A geometric approach to quantumcircuit lower bounds. Quant. Info. Comp. 6, 213.

    [30] Nielsen, M. A., Dowling, M. R., Gu, M. & Doherty, A. C.2006 Optimal control, geometry, and quantum comput-ing. Phys. Rev. A73, 062323.

    [31] Marsden, J. E. & Ratiu, T. S. 2003 Introduction to Me-chanics and Symmetry. (New York: Springer).

    [32] Bruveris, M., Gay-Balmaz, F., Holm, D. D. & Ratiu,T. S. 2011 The momentum map representation of images.J. Nonlinear Sci. 21, 115

    [33] Russell, B. & Stepney, S. 2014 Zermelo navigation inthe quantum brachistochrone. J. Phys. A48 (to appear;arXiv:1409.2055).

    [34] Brody, D. C., Gibbons, G. W. & Meier, D. M. 2015 Time-optimal navigation through quantum wind. New J. Phys.17 (to appear; arXiv:1410.6724).

    [35] Robles, C. 2006 Geodesics in Randers spaces of constantcurvature. Trans. Amer. Math. Soci. 359, 1633.

    http://arxiv.org/abs/1408.2465http://arxiv.org/abs/1409.2055http://arxiv.org/abs/1410.6724

    Acknowledgments References