Fast adiabatic evolution by oscillating initial Hamiltonians

Francesco Petiziol,1, 2, ∗ Benjamin Dive,3 Florian Mintert,3 and Sandro Wimberger1, 2 1Dipartimento di Scienze Matematiche, Fisiche e Informatiche, Universit`adi Parma, Parco Area delle Scienze 7/a, 43124 Parma, Italy 2INFN, Sezione di Milano Bicocca, Gruppo Collegato di Parma, Parco Area delle Scienze 7/a, 43124 Parma, Italy 3Department of Physics, Imperial College, SW7 2AZ London, UK We propose a method to produce fast transitionless dynamics for finite dimensional quantum sys- tems without requiring additional Hamiltonian components not included in the initial control setup, remaining close to the true adiabatic path at all times. The strategy is based on the introduction of an effective counterdiabatic scheme: a correcting Hamiltonian is constructed which approximatively cancels nonadiabatic effects, inducing an evolution tracking the adiabatic states closely. This can be absorbed into the initial Hamiltonian by adding a fast oscillation in the control parameters. We show that a consistent speed-up can be achieved without requiring strong control Hamiltonians, using it both as a standalone shortcut-to-adiabaticity and as a weak correcting field. A number of examples are treated, dealing with quantum state transfer in avoided-crossing problems and entanglement creation.

I. INTRODUCTION control of complex interactions, and more generally of Hamiltonians which do not belong to the available con- Quantum adiabatic processes [1] are ubiquitous in trol setup. As a result, the implementation of HCD(t) quantum science and they represent an important re- is often tricky if even possible [3, 5, 6]. For this reason, source for quantum control. The different STAs have been developed which connect adi- states that when the system Hamiltonian H(t) is var- abatic states at different desired times, but completely ied slowly enough in time from an initial configurations deviate from the adiabatic states at intermediate times [7–13], therefore giving up the benefits of true adiabatic- H(ti) to a final configuration H(tf ), then a state initially in an eigenstate of H(t) will remain so during the whole ity. evolution [1, 2]. This makes such protocols intrinsically Here, we propose a method for achieving fast adia- robust against experimental imperfections. On the other batic driving remaining close to the adiabatic path, with- hand, the necessity of very long timescales for their imple- out needing new unrealizable terms in the Hamiltonian. mentation dramatically limits the number of operations This works by modulating the original Hamiltonian of the which can be performed on the system within reasonable system in time such that it effectively replicates the dy- coherence times. namics induced by HCD without needing any additional Recently, much work has been done towards the design control Hamiltonian on the quantum system. of adiabatic-inspired control strategies which, on the one In order to do so, we first resort to control-theoretic hand, inherit the robustness of the adiabatic dynamics, techniques to study how the matrix structure of the cor- while, on the other, avoid the necessity of slow driving. recting field HCD is related to the initial set of control Among such so-called “shortcuts to adiabaticity” (STAs), Hamiltonians which constitute H. This shows that HCD a particularly promising method was introduced under can always be emulated, to arbitrary precision, by intro- the name of counterdiabatic (CD) [3] or transitionless [4] ducing a suitable (fast) time dependence in the control quantum driving. The basic idea which is put forward parameters. Second, we identify a class of Hamiltonians, is that it is always possible to reverse-engineer a correct- generalizing the set of real ones, for which the matrix ing Hamiltonian HCD(t) such that the total Hamiltonian structure of HCD can be discussed on general algebraic arXiv:1807.10227v2 [quant-ph] 8 Nov 2018 H(t) + HCD(t) keeps the system in the instantaneous grounds and always involves components which are not eigenvectors of H(t) without requiring it to change slowly directly controllable. They can, however, be simulated – that is, such that the adiabatic dynamics is an exact arbitrarily well using existing protocols. solution of the time-dependent Schr¨odingerequation. Building on these results, we describe the construction When facing a quantum control problem, one must of an effective counterdiabatic (E–CD) field HE(t), which take into account that only a fairly restricted number of is a time-dependent combination of the initially available Hamiltonians can be realized and controlled in practice. control Hamiltonians. The control functions in HE will The crucial drawback of the CD method is that, although be chosen to be oscillating fast with respect to the nat- a well-defined expression for computing HCD(t) exists, ural time dependence of H, and HE will be enforced to the correcting field typically requires time-dependent simulate the dynamics UCD induced by HCD. As a result, the E–CD evolution tracks the adiabatic path, being arbi- trarily close at a set of sampling time points while slightly deviating at intermediate ones. The general idea of pro- ∗ [email protected] ducing the UCD dynamics approximatively by working 2

−iϕn(t)/~ with a set of available control terms was also pursued, n e n(t) n(t0) , with ϕn arbitrary phases, be yet with a different strategy, in Ref. [14]. a unitary matrix| whichi h diagonalizes| H(t) at all times. One must take into account that the adiabatic require- PThat is, ment of slow drivings can be recast as the need of strong † fields. With this in mind, we characterize the efficiency of Ud(t)H(t)Ud(t) = diag E1(t),...,EN (t) . (1) the E–CD scheme in terms of final fidelity and duration { } for different strengths of the correcting Hamiltonian HE The corresponding reverse-engineered Hamiltonian † as compared with the uncorrected H. We show that the Hcorr = i~∂tUdUd reads E–CD field, when acting as an auxiliary weak term in H, realizes a consistent speed up of the adiabatic evolution. Hcorr(t) = ∂ ϕ (t) n(t) n(t) + HCD(t), t n | i h | The E–CD Hamiltonian also works as a stand-alone STA, n X giving even better results but at the price of completely losing true adiabaticity. The net result is the ability to where we have introduced the CD field [4] generate entanglement or perform state transfer signifi- HCD(t) = i~ ∂tn(t) n(t) . (2) cantly faster which, when taking into account noise from | i h | n an external environment, inexorably leads to higher fi- X delities. The Hamiltonian H drives the instantaneous eigen- After recalling the theory of CD driving in Sec. II, corr vectors of H(t) exactly, with relative phase factors ϕ (t), the control-theoretic setting in presented in Sec. III, to- n e−iϕn(t)/~ n(t) . In particular, one can choose ϕ (t) = 0 gether with the results on the general matrix structure n for all n,| whichi gives H = H . Therefore, imple- of the correcting Hamiltonian H . The effective CD corr CD CD menting just H is sufficient if one is only interested in method is introduced in Sec. IV, where the derivation of CD preserving the populations of the instantaneous eigenvec- the E–CD field is fully worked through for a single spin tors. Another interesting choice is ϕ (t) = t E (t0)dt0, system. The results are then exemplified via a number n n in which case ϕ are the adiabatic dynamic phase factors. of applications in Sec. V, the first being the Landau- n This gives H = H(t) + H , so that H R can be in- Zener-Majorana model [15–18]. We afterwards show that corr CD CD terpreted as a correcting field acting beside the initial a two-qubit entangled state can be prepared with 99.9% Hamiltonian H(t). fidelity ten times faster with respect to purely adiabatic Alternative expressions of H are useful in order to evolution, using an E–CD field with strength comparable CD highlight specific properties. For example, from to that of the original Hamiltonian. We further discuss the case of a three-level system dynamically undergoing N a sequence of avoided crossings in the energy spectrum m m ∂tH n n HCD = i~ | i h | | i h |, (3) [19]. In this section, the E–CD method is also bench- Em En m6=n n=1 − marked against traditional finite-time adiabatic driving, X X and the main advantages and limitations are discussed in one can see that HCD, after having absorbed the geo- detail, before proceeding to the conclusion in Sec. VI. metric phases into the ϕn [4], is purely off-diagonal in the basis of instantaneous eigenvectors (adiabatic basis). In terms of the Hilbert–Schmidt inner product Tr(A†B), II. COUNTERDIABATIC DRIVING this implies that HCD is orthogonal to H, and in general to all matrices commuting with H. It is also orthogonal In this section, we recall the central elements of the to ∂tH. Moreover, if we assume that the whole evolution theory of counterdiabatic fields [3, 4] which underpins takes place starting from an initial time ti for a total time our method. τ, one can rescale time according to s = (t ti)/τ and − A unitary evolution U(t) is always a solution of see that HCD scales like 1/τ. In other words, the faster the Schr¨odinger equation with Hamiltonian HU (t) = the process, the stronger the field HCD should be. This † i~∂tUU . Equivalently, the same HU (t) is a reverse- property will be important in our discussion and it is engineered Hamiltonian producing dynamics U(t) ex- particularly interesting in the context of quantum speed actly. Our aim is to find this HU (t) for the U(t) which limits [20, 21]. performs perfect adiabatic transfer. By construction, the By taking the derivative of Eq. (1), one can obtain the new Hamiltonian will give the desired dynamics in an ar- relation bitrarily short period of time. i Let the initial Hamiltonian of the physical system be ∂tH(t) = [H(t),HCD(t)] + ∂D(t), (4) ~ N where ∂D(t) = N ∂ E (t) n(t) n(t) . Equation (4) H(t) = En(t) n(t) n(t) , n=1 t n | i h | highlights the fact that ∂D(t)| generatesi h the| variation of n=1 P X the instantaneous eigenvalues of H(t), while HCD is re- having instantaneous eigenvalues En(t) and in- sponsible for the variation of eigenvectors. This is so stantaneous eigenvectors n(t) . Let Ud(t) = since [∂D(t),H(t)] = 0, and then ∂D makes H(t) “move” | i 3 inside the set of Hamiltonians which commute with H the set of reachable states, i.e., the set of unitary ma- at time t, without changing the eigenvectors. On the trices that can be obtained as solution of the controlled other hand, the term i[H,HCD]/~ determines the devia- Schr¨odingerequation for different choices of the control tion from zero of the off-diagonal elements of H(t + dt) functions, coincides with the connected Lie group gen- with respect to the instantaneous adiabatic basis at time erated by . An intuition for this can be given as fol- t. lows. If anL element iH belongs to , then the Lie group element e−iHt that− it generates atL time t can be re- alized, to arbitrary precision, by suitably concatenating III. CONTROL FRAMEWORK group elements generated by Hamiltonians in the initial set iH, which belong to the one-parameter subgroups −− − In order to realize H , it is of great interest to un- e iH1t, . . . , e iHM t . From a control-theoretic perspec- CD { } derstand how the structure (matrix components) of the tive, this means that the evolution produced by H in a Hamiltonian HCD is related to the structure of the initial time t can also be obtained by a sequence of evolutions Hamiltonian H and to the control resources. In this sec- governed by the Hamiltonians H, in general in a different tion we exploit techniques from control theory to show total time. that, assuming that HCD can be computed, its action Lie algebraic methods were also used in Ref.s [9, 10], can be approximated at all times, arbitrarily well, by a in the context of STA, for designing feasible shortcuts time-dependent tuning of the initial parameters in the connecting the same initial and final adiabatic states, but system Hamiltonian. This proves that the realization of following different paths in the Hilbert space. our E–CD scheme is actually possible, and the specific Assuming that all Hamiltonians involved are made construction will be discussed in the next section, Sec. traceless, can at most be su(N), the algebra of skew- L IV. Second, this study permits us to identify a class of Hermitian traceless matrices generating the group of spe- Hamiltonians, that generalizes the set of real Hamilto- cial unitary matrices SU(N), and has dimension N 2 1. − nians, for which all matrix elements of H are not di- When = su(N), the system is said to be (operator) CD L rectly accessible without implementing new terms, which controllable [22]. The Lie group generated by through L L are not present in the original Hamiltonian. the exponential map will be denoted by e . Let H(t), of finite dimension N, be realized by tun- Since iHCD(t) is a skew-Hermitian matrix, it must − ing some available time-independent control Hamiltoni- belong to su(N) at all times. Therefore, if = su(N) L ans H = H1,...,H via a set of continuous control then iHCD(t) is obviously in . The first result we show { M } − L functions u(t) = u (t), . . . , u (t) . That is, H(t) can is that HCD is a general property of HCD, even when 1 M ∈ L be expressed in the{ form } is not equal to su(N), but is rather a smaller subalgebra L su(N) – that is, even when the system is not fully M controllable.L ⊂ This is an interesting result for two reasons. H u(t) = u(t) H = uk(t)Hk. First of all, it allows us to restrict the class of Hamiltoni- { } · =1 kX ans needed to identify and realize HCD, with respect to the full set of traceless Hermitian matrices. Second, and Let iH be the vector of skew-Hermitian matrices more importantly in the present work, it means that the iH−,..., iH . The matrices iH generate the 1 M action of H can be always approximated, in the sense so-called{− dynamical− } Lie algebra of− the system [22]. CD of the action of the Lie group, by working only with the This is the smallest algebra whichL contains iH, all initially available Hamiltonians. This result is stated in possible commutators [ iH , iH ] of matrices− from j k the following theorem, whose proof is given in Appendix iH, all possible commutators− − of commutators, and A. In all the following results, we will generally assume so− on, considering all possibly nested commutators that H can be computed. This excludes, in general, [ iH , [..., [ iH , iH ]] ... ]. CD l j k cases in which evolving eigenvectors can be degenerate −A basis of− can− be constructed by calculating, as a [7]. first step, theL commutator of all possible pairs of matri- ces drawn from iH. Among the new obtained matrices, Theorem 1. Let the Hamiltonian of the system be ex- − one should select those which are linearly independent pressible, at all times and for all values of the control from themselves and the original set, and compute the functions u(t) = u1(t), . . . , uM (t) , as a linear combi- commutator of such new matrices with this original set. M{ } nation H(t) = uk(t)Hk of time independent control The procedure is repeated iteratively until no new lin- k=0 Hamiltonians H = H1,...,HM . Let be the Lie alge- early independent elements are produced. An explicit P { } L bra generated by the matrices iH. Assuming the HCD algorithm can be found in [23]. In our context, the linear − exists, then iHCD(t) belongs to , for all times t. span span H can be though of as the set of all pos- − L sible matrices{ } attainable by H(t) at different times, for When the Lie algebra has an additional structure, L different values of the control functions. the general form of HCD can be characterized more ac- Dynamical Lie algebras are of central importance in curately. In order to do so, let us introduce Cartan de- the study of the controllability of quantum systems compositions, which are important tools in the study of [22, 24, 25]. This has its origin in the fact [26] that controllability of quantum systems [22, 24, 27]. These 4

J3 among the three possible J1,J2,J3 . Explicit diagonal- { } ization and computation of HCD [3, 4] shows that then H~ CD(t) HCD is proportional to the third generator. This can be also seen from Eq. (4): interpreting H geometrically as a vector H~ (t) = H1,H2,H3 in the three-dimensional algebra, the matrix{ ∂D generates} the stretching of the J 2 Hamiltonian vector along in its own direction, while HCD generates rotations on the plane spanned by Ji and Jj; H~ (t) see Fig. 1. Therefore, recalling the properties of the ro- tations in 3D, HCD must lie in the direction of Jk. In this J case the Cartan decomposition is unitarily conjugate to 1 H~ (t + dt) ∂ H = i[H,H ] so(2) I. t ⊥ CD ⊕

∂tHk = ∂D IV. EFFECTIVE CD FIELD

FIG. 1. Sketch of the geometric interpretation of relation (4) From Theorem 1 one knows that (i) HCD can be ap- for the algebra su(2). The Hamiltonian H(t) = u1(t)J1 + proximated by a suitable choice of the control functions u2(t)J2 can be interpreted as a vector H~ = {u1, u2, 0}, whose magnitude pu2 + u2 is the largest eigenvalue, in the three- u(t) for all t and that (ii) HCD can be expressed as a lin- 1 2 ear combination of (nested) commutators of elements of dimensional space spanned by the su(2) generators J1,J2,J3. The term ∂D is responsible for variations of magnitude (eigen- H. Observation (i) suggests that the problem of finding k values), which we indicate with ∂tH . The term i[H,HCD] an effective CD field can be formulated as that of finding generates rotations of the vector, i.e., a change of eigenvec- a correcting Hamiltonian HE, of fully controllable form tors of H, and we denote this component of the variation ⊥ M ∂tH . HE(t) = c(t) H = c (t)H , (6) · k k =1 kX are decompositions of the algebra into the direct sum which emulates HCD. More precisely, the Hamiltonian form = h p which satisfy the commutationL relations L ⊕ HE should produce a dynamics UE emulating the one [h, h] h, [h, p] p, [p, p] h. (5) induced by HCD, i.e., UCD. While the algebraic struc- ⊆ ⊆ ⊆ ture suggests this can be done, it does not constructively We can then state the following result (some of the quan- specify what the ck(t) need to be. To find them, from ob- tities were defined in Sec II). servation (ii), we choose to adopt a representation of the propagators UE and UCD based on a Magnus expansion, Theorem 2. Given a Cartan decomposition = h p, if see Appendix B, where terms involving commutators of L ⊕ span iH p and i∂D p for all t, then iHCD(t) the Hamiltonians appear naturally. The first two terms {− } ∈ − ∈ − ∈ h for all t. are given for completeness,

Proof. If iH p, then also i∂tH p, since it can be t − ∈ − ∈ (1) i written as a linear combination i∂ H = i∂ u(t) H. M (t) = H(t1)dt1, t t − p − − · ~ 0 If also i∂D for all t, then from Eq. (4) it holds Z 2 − ∈ t t1 [H,H ] p. From commutation relations defining (2) i CD M (t) = − dt dt [H(t ),H(t )]. the Cartan∈ decomposition, one can then conclude that 1 2 1 2  ~  Z0 Z0 iHCD h for all t. − ∈ The first step for designing the effective field HE is to Let us clarify the result of Theorem 2 by means of two choose an ansatz for the control functions c(t) involving examples. The first one is the situation in which H in- a certain number of free parameters. The second, con- cludes exclusively real symmetric Hamiltonians. If this sidering a small evolution time t, is to ask the first terms is the case, H(t) can always be diagonalized by a (real) of the Magnus expansion of UCD and UE to coincide up † special orthogonal matrix O(t). Thus HCD(t) = i∂tOO to a desired order in t. This requirement will produce is necessarily a purely imaginary, skew-symmetric ma- constraint equations for the free parameters. Since UCD trix, and iHCD(t) has thus no component on iH for and UE cannot coincide at all times, one needs to choose − − all t. When the set of control matrices generates the full a discrete temporal grid tk for decomposing the whole algebra su(N), this corresponds to a Cartan decomposi- evolution, and enforce the{ approximate} equality at such tion of the form so(N) I, where so(N) is the algebra time points. As a result, the approximating dynamics ⊕ ⊥ spanned by the real matrices in su(N), while I = so(N) will match the true one at times tk , while deviating at is the set of purely imaginary matrices in su(N). intermediate times. { } The second example is the angular momentum algebra The matrix components of HCD which do not belong su(2) with H containing two arbitrary generators J ,J to span H will not appear in the first Magnus term of i j { } 5

H +HE, which is essentially the time average of H +HE. A discussion of the accuracy of the method is presented They will appear from the commutators in the following in Appendix B 2, where an estimate of the expected error terms. Therefore, if one wants HE to reproduce HCD in the probability of not being in the target state, at the effectively, it must hold that HE gives no contribution to end of one period, is derived as a function of the number the first term, i.e., has vanishing time average on each of solved constraint equations. Let us show the overall interval (tk, tk+1). Besides, since the mediated effect will procedure by working through a specific case. be of some order m greater than one in t, the magnitude of the control functions in HE must be proportional to t−X , for some X > 0 depending on m, in order to amplify A. Worked through case: single spin the mediated effect so that it acts at first order. Making these general ideas concrete, we choose the We concentrate here on the case of the Lie algebra control functions in Eq. (6) to have the form of a trun- su(2), which can physically describe, for example, a sin- cated Fourier series, gle spin driven by magnetic fields. A first application of STA methods via CD driving to this kind of system L was studied in Ref. [29]. Since for tracking the instan- X ck(t) = ω Ak,j sin(jωt) + Bk,j cos(jωt)], (7) taneous ground state it is sufficient to implement HCD j=1 without H, as discussed in Sec. II, and for simplicity X  of presentation, we treat the case in which one wants to involving at most L harmonics of the fundamental fre- approximate HCD alone by means of HE. The general quency ω. This way, the needed time grid is naturally case, H + HCD approximated by H + HE, will be dis- defined by the set of stroboscopic times tk = ti + kT , cussed at the end of this section. For clarity, let us work k N, where T = 2π/ω is the period of the control with the Pauli matrices σ , σ , σ , having commuta- ∈ { x y z} functions and ti the initial time. The constant ampli- tion relations [σi, σj] = 2iijkσk. We assume that the tudes A ,B are the free parameters used to enforce { k,j k,j} Hamiltonian of the system is controllable only along two the constraints on the Magnus expansion at the end of directions, say x and z, and can thus be written as: each period. As a result, the control functions so deter- mined will be discontinuous between different periods, H(t) = ux(t)σx + uz(t)σz. (8) due to the jumps in the values of the amplitudes. From a practical perspective, this is of course physically in- The field HCD, as discussed in Sec. III, will be directed convenient, and so one would like to come up with an along the unavailable direction. It can be calculated ex- interpolation providing continuous and possibly smooth plicitly [3, 4] and it has the form HCD(t) = fCD(t)σy, with: functions Ak,j(t),Bk,j(t). We will see that this can in- deed be done, and often in a natural way. 1 ux(t)∂tuz(t) ∂tux(t)uz(t) The smaller the period T , the better the target dy- fCD(t) = 2 − 2 . (9) −2 ux(t) + uz(t) namics UCD will be sampled, so in general HE will need to oscillate fast with respect to the time-dependence of To achieve compensation to first order in T at strobo- H. Raising the number of harmonics L permits us to scopic times, we apply the E–CD Hamiltonian HE = introduce more parameters, and thus to obtain and solve cx(t)σx + cz(t)σz, with control functions of the form (7). constraint equations produced by higher Magnus terms. The first non-zero term of the Magnus exponent pro- Ideally, one would like to find a good compromise be- duced by HE is the second one, and has the desired ma- tween high sampling rates of UCD, good approximation trix structure. Generalizing to the choice of two arbitrary at stroboscopic times, while keeping the fundamental fre- generators σi, σj , one has { } quency and the number of harmonics as small as possi- ble. This will be discussed in detail in Sec. V. A strategy 2π L 1 M (2)(T ) = i [A B B A ] σ . based on similar ideas was introduced in Ref. [28] for the E − ijk ω2−2X n i,n j,n − i,n j,n k n=1 purpose of producing effective time-independent Hamil- X (10) tonians via optimal control. From Eq. (10) one can see that the highest order in T is Let us now focus on the class of Hamiltonians identi- produced by the interplay of sin and cos components re- fied by Theorem 2. Due to the commutation relations lated to the same harmonic n, belonging to different con- from Eq. (5), the Magnus terms of UE involving an odd trol functions c and c . There is no mixing between dif- (2k) i j number of commutators, ME , belong to h. Those in- ferent harmonics, and no mixing between fields with the (2k+1) volving an even number of them, ME , belong to p same phase. This feature is due to the properties of the T t1 instead. Therefore, the general form of the constraint integrals of the form 0 dt1 sin(nωt1) 0 dt2 cos(kωt2). equations is Higher-order Magnus terms can be computed analyti- cally, but the expressionsR get longer andR longer with re- (2k) (k) ME = MCD, spect to Eq. (10). (2k+1) Now, let us find the constraint equation to first order ME = 0. in T . For this purpose, two amplitudes are sufficient and, 6

(2) since we want ME to be of order T , we choose X = 1/2 initial Hamiltonian H, and the method is highly nonadia- in Eq. (7). A possible choice of control functions is then batic. In fact, even if the instantaneous eigenvectors of H are tracked during the evolution, the system is not in an cz(t) = A√ω sin(ωt); cx(t) = B√ω cos(ωt). (11) instantaneous eigenstate of the total (corrected) Hamil- tonian which is actually applied, H + HCD (or H + HE). Using this functions, we can compute the first Magnus On the other hand, if the nonadiabatic effects are weak, terms. Let tn = ti + nT , n = 1, 2,... , and let us indi- then one can think of being close to true adiabaticity. + cate with T the integral over a full period tn T . The tn Depending on the problem, it might be convenient to exponents of the Magnus expansion for the CD and the move to the time-dependent interaction picture for com- effective dynamicsR at the end of each periodR nT of the puting the effective field. In general, though, it becomes control functions are, respectively, more complicated to compute the terms in the Magnus expansion. In any case, the treatment above can be re- T peated, and to order T 3 in infidelity, one still obtains the MCD(T ) = i fCD(t)dt σ , − y same effective field as in Eq. (15). Z ! 3 = i[fCD(t + T/2)T + o(T )]σ , (12) − n y 3/2 V. APPLICATIONS ME(T ) = iABT σ + o(T ). (13) − y The last equality in Eq. (12) is obtained by formally In this section, we discuss three applications of the Taylor-expanding fCD(t) around the midpoint of the in- method introduced in Sec. IV. The first one is the tegration interval, tn +T/2, and then integrating. Equat- Landau-Zener-Majorana (LZM) model [15–18]. This is ing order-T terms in Eqs. (12) and (13) one obtains the the standard milestone in the study of nonadiabatic ef- first constraint equation: AB = fCD(tn + T/2). This can fects and more generally of the physics near avoided be straightforwardly solved, and a possible solution for crossings in the energy spectrum. The second one re- the n-th period is gards the adiabatic preparation of a two-qubit entangled state, a task of central importance for quantum infor- A = fCD(tn + T/2) ,B = sgn fCD(tn + T/2) A, mation processing [30]. The last application generalizes | | (14) the first one and deals with a three-level system whose wherep sgn(x) indicates the sign of x and takes care of the natural evolution induces the formation of a sequence of case in which fCD(t), and thus the product AB, is nega- avoided crossings [19]. Such a model can describe local tive. Different solutions may be more suitable depending parts of many-body nontrivial energy spectra. on the structure of fCD. We also benchmark the efficiency of the E–CD method The fact that the above solution for A and B, Eq. against standard (finite time) adiabatic driving. This is (14), is a simple function of fCD evaluated at the mid- done with a focus on the LZM case, it being the simplest point of the period suggests a straightforward manner to nonadiabatic scenario. This analysis underpins the use- interpolate the solutions in all intervals, obtaining con- fulness of the method not only for theoretical purposes tinuous and smooth amplitude functions A(t) and B(t). but, more importantly, for practical applications in the It is indeed sufficient to replace fCD(tn + T/2) fCD(t) control of quantum systems. in Eq. (14) and the same accuracy is maintained.→ As a For setting up the comparison between the E–CD conclusion, the effective field can be written as and the adiabatic paradigms, a quantification of their respective performance is first needed. Since this is

HE(t) = fCD(t) ω sgn fCD(t) cos(ωt)σx+sin(ωt)σz . not an obvious task, let us here introduce the crite- | | { } ria which are adopted for this purpose. The princi- h (15)i p ple figure of merit we are interested in is the infidelity Two immediate observations can be made from Eq. (15): = 1 ψ(t ) gs(t ) 2, i.e., the probability that the First, H is proportional to √ω, so, as expected, a better F f f E finalI state− | hψ(t )| of thei | system is not in the instanta- sampling can be obtained at the cost of having a stronger f neous ground| statei gs(t ) at the end of the protocol. field. Second, due to the proportionality with f (t) , f CD We are also interested| in makingi the evolution as fast as the effective CD field inherits to some extent the| behav-| possible with as few resources as possible, but there is not ior of the exact one. In particular, it vanishesp when H CD a unique way to quantify this. The general idea is that does, that is, when the probability of nonadiabatic tran- one should not only take into account the full duration sitions is zero. (3) of the protocol, but also the intensity of the Hamiltonian The third constraint equation, M = 0, can be solved E which is applied to the system. Indeed, the adiabatic the- by adding a further term 4A sin(2ωt)σ to Eq. (15). z orem [1, 31] (see Appendix A) formally states a property We conclude this section− by discussing the more inter- of the solutions of the family of Schr¨odingerequations esting case in which H + HE approximates H + HCD. This scenario is particularly important in a quasiadia- i~∂sUτ (s) = τH(s)Uτ (s), (16) batic regime. When nonadiabatic effects are strong, the correcting fields become dominant with respect to the for varying τ. Therefore, the theorem can be interpreted, 7

1 1 0.8 0.8 0.6 0.6 0.4 0.4 0.2 0.2

Probability 0 Probability 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 s s ] 2 0.3 − 4.0 10

0.2 × 3.0 2.0 0.1 1.0 Infidelity 0 0.0

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 Infidelity [ 0 0.2 0.4 0.6 0.8 1 s s

FIG. 2. Evolution given by the LZM Schr¨odingerequation FIG. 3. Evolution given by the E–CD field, Eq. (20), with (18), with parameters ε = 20, τ = ε. Above, the populations parameters ε = 20, τ = 20, NT = 2τ. Above, the populations of the bare states are depicted. The dashed lines represent of the bare states are shown, with an inset zooming around the the true instantaneous eigenstates, while the solid blue and avoided crossing. The solid (orange/blue) lines represent the dot-dashed orange lines represent the LZM evolution. Below, E–CD dynamics, which oscillates around the true adiabatic the infidelity, i.e., the probability of nonadiabatic transition, one, represented by dashed lines. Below, the infidelity, i.e., is shown. In both plots, the thin gray line represents the the probability of nonadiabatic transition, is shown. prediction given by the LZM formula, Eq. (19).

~β/2 prevents a net crossing, and an avoided crossing is on one side, as describing a varying-duration behavior, if produced instead with minimal gap ~β. Assuming that Eq. (16) is obtained after a rescaling s = t/τ of the phys- the sweep spans an energy difference ~α in a time interval ical time for a fixed Hamiltonian. Alternatively, it can be ti t tf , and that the anticrossing takes place at the interpreted as describing an intensity-varying behavior, if ≤ ≤ intermediate time tc = (tf ti)/2, the Hamiltonian can Eq. (16) is obtained from considering the duration fixed be written in the form − and amplifying the Hamiltonian like H(t) τH(t). → These considerations lead us to study the infidelity F α t tc β I HLZM(t)/~ = − σz + σx. both as a function of the protocol duration and of the 2 tf ti 2 ratio between the strength of HE and that of the initial − Hamiltonian H(t). We formalize this by choosing, as a In order to study the adiabatic properties of the system, measure of strength ( ), the maximal Frobenius norm S · it is convenient to rescale the quantities in terms of β, over the whole evolution, which defines a fundamental frequency scale of the prob- lem. Furthermore, let us parametrize the time according (H) max H(t) . (17) t−ti t F to s = , with 0 s 1. In terms of the dimen- S ≡ k k tf −ti ≤ ≤ sionless quantities s, ε = α/β and τ = β(t t ), the † f − i The Frobenius norm is defined as A F = Tr(AA ) and Schr¨odingerequation becomes provides a quantification of the magnitudek k of the whole Hamiltonian matrix. Different choices are possible as cost ∂U(s) τ 1 functions depending on the resource one is interested in. i = ε s σz + σx U(s). (18) ∂s 2 − 2 For instance, one might be practically interested in the     maximal matrix element, or the maximal or average am- The parameter ε determines the distance, in units of β, plitude reached by the control functions. from the avoided crossing at the beginning and at the end With these tools at hand, let us now proceed to the of the protocol. It also rules the decay rate of the LZM discussion of the above mentioned applications. oscillations in the transition probability, i.e., of nonadi- abatic transition, after the anticrossing [32, 33] (see also A. Landau-Zener-Majorana model Appendix A in [34]). Assuming that the system starts in the ground state, the transition probability to the excited state (i.e., of a nonadiabatic transition) is shown in Fig. We demonstrate the general approach detailed in Sec. 2. The asymptotic value, for large times, is given by the IV A for the specific case of the Landau-Zener-Majorana Landau-Zener formula [16–18], model [15–18]. Such a model describes a linear sweep of the gap between the energy levels of a two-level system. πτ PLZ = exp . (19) As they approach each other, the presence of a coupling − 2ε   8

100 100 1 10− 2 10 2 10− − 3 10− 4 4 10− 10− 5 Adiabatic 10− LZ formula 6

Infidelity Infidelity 10 6 − 10− k =1 7 k =1/2 Adiabatic 10− 8 k =1/4 10− k =1 8 10− k =1/8 k =1/2 9 k =1/16 k =1/4 10− 100 101 102 100 101 102 103 τ τ

FIG. 4. Comparison between finite-time adiabatic dynamics FIG. 5. The results obtained with the procedure described in and effective CD driving, using only HE. Different (colored) Fig. 4, shown for the application of the total Hamiltonian H+ line styles represent the results for different ratios k in rela- HE. The solid line represents the adiabatic dynamics, while tion (21). The empty squared points represent the adiabatic different line styles represent the E–CD method for different dynamics, whose behavior is well predicted by the LZ for- values of the factor k in Eq. (21), i.e., for different strength mula of Eq. (19) (black solid line). For each τ, the number of of HE as compared to H. The infidelity oscillates fast for periods NT is determined as the largest integer smaller than different τ, as shown by the thin gray line for the case k = ωτ/2π, once ω is chosen to be the largest such that Eq. (21) 1/2, so the depicted points are representing an average of 20 is satisfied. The conversion to integers explains the steps in surrounding points. the colored curves. The beginning of the latter is determined by the condition NT > 1. after having selected the maximal frequency ω such that, for a certain factor k, the inequality The CD field for this problem, satisfying i∂sU = (HE) k (H), (21) τHCD(s)U, is HCD(s) = fCD(s)σy, with S ≤ S with defined in Eq. (17), holds. 1 ε S fCD(s) = 2 . Let us remark that nonadiabatic transitions are expo- −2τ ε2 s 1 + 1 − 2 nentially weak in the adiabatic parameter [36, 37]. On  the other hand, one expects that the efficiency of the This CD protocol was experimentally studied in Refs. [5, E–CD method increases polynomially when raising the 35]. From the results of the previous Sec. IV A, Eq. (15), frequency ω, since it is based on a perturbative argu- the dimensionless effective CD field, satisfying i∂sU = ment. Therefore, one can safely predict that the adia- τHE(s)U, is batic method, for very large τ and for a fixed maximal strength of the control fields, performs better. Nonethe- HE(s) = fCD(s) ω[ cos(ωsτ)σx + sin(ωsτ)σz], (20) less, what we are interested in is the intermediate range of | | − durations, very far from the asymptotic limit in τ, which p with ω = 2π/(βT ). The negative sign in front of the is the regime where practical protocols need to work. cosine is due to the fact that fCD is always negative. The results for the LMZ model are shown in Figs. 4 The dynamics given by H alone can be compared with and 5. In Fig. 4, the case in which only HE, as a stand- the one produced by HE from Figs. 2 and 3, for β = 1. As alone shortcut to adiabaticity, is applied is reported. As expected, the E–CD dynamics oscillates fast, remaining expected from the above discussion, the infidelity under close to the target one. In the vicinity of the anticross- effective CD driving scales like a power-law for large du- ing, where nonadiabatic effects are strong, the deviation rations τ, and eventually the adiabatic exponential decay between consecutive periods is larger. takes over. However, in the intermediate range of τ, the Let us now study the comparison between the E–CD effective method does permit us to achieve an important and finite-time adiabatic methods more thoroughly. In speed-up of the adiabatic process, for fixed infidelity. The particular, we show that the E–CD strategy can provide a range of improvement gets wider and wider as one allows consistent speed up of adiabatic protocols, characterizing a stronger correcting field, while still remaining below the important regimes of parameters in terms of duration the maximal norm of H. For example, one can reach a of the protocol and strength of the involved Hamiltoni- fidelity of 99.9% twenty times faster using an E–CD ans. field of the∼same strength of H(t)[k = 1 in Eq. (21)]. Specifically, we compute the infidelity obtained with Figure 5 shows the same results for the case in which the E–CD Hamiltonian for different total durations τ, the E–CD Hamiltonian is implemented beside H(t), and 9 thus acts as a weak correcting field to the major dynamics 100 induced by H(t). The final infidelity oscillates strongly 1 for different values of τ (thin gray line), so the points 10− 2 indicate an average of 20 surrounding points. The oscil- 10− 3 lations can bring the infidelity to much lower values, but 10− never much higher than those represented by the points. 4 10− This method performs worse in numbers than the previ- 5 ous case (Fig. 4) for small τ, but its main advantage is 10− Infidelity 6 that it allows one to preserve true adiabaticity to a cer- 10− 7 tain extent. That is, the driven states are close to being 10− instantaneous eigenstates of the full Hamiltonian H+HE, 8 10− E–CD since HE is weaker than H and averages to zero in time. 9 Adiabatic One can see that there is an initial regime in which the 10− 100 101 102 103 protocol is too fast for the correction to give a significant tf H(t) dt improvement, and a second regime in which the infidelity ti k kF follows essentially the same behavior as in Fig. 4. As an R example, an infidelity of at least 10−4 can be achieved FIG. 6. Comparison between finite-time adiabatic and E–CD with a speed up of around 6.5∼ times, using a maximal driving in terms of the infidelity as a function of the time in- strength of the correcting field∼ satisfying Eq. (21) with tegral of the Frobenius norm of the respective Hamiltonians k = 1/2. A speed up of 2.2 times is obtained in the case [Eq. (22)], for parameters ε = 40, β = 1. The adiabatic curve (upper line) is obtained by raising the amplifying factor (cor- k = 1/4 for the same infidelity. All these results confirm responding to the total duration) τ of the original Hamilto- that the E–CD method represents an efficient strategy nian, while the E–CD curve (lower line) is obtained by keeping for obtaining fast adiabatic evolution, without requiring the initial value of τ fixed, while raising the frequency ω. The strong drivings. quantity on the abscissa is dimensionless, having set ~ = 1. The second criterium for comparing adiabatic and ef- fective CD driving is the following: we consider the total duration to be fixed, 0 s 1, and we compare the dy- ≤ ≤ namics produced by the amplified Hamiltonians Hτ (s) with the one produced by HE for increasing values of the 1 frequency ω. Due to the intertwined relation between 0.9 duration and amplitude, we choose as a basis for com- 0.8 0.7 parison the integral of the norm over the whole evolution 0.6

tf Populations 0.5 0.7 0.75 0.8 0.85 0.9 0.95 1 dt HX(t) . (22) k kF Zti s The results are shown in Fig. 6. As for the previous 0.16 criterium adopted, they are rather promising, confirm- 0.12 ing the validity of the E–CD method. Indeed, the E–CD 0.08 Hamiltonian turns out to always produce a better infi- 0.04 delity, for a given total integral norm of the Hamiltonian, Infidelity 0 in the range of values studied. 0.7 0.75 0.8 0.85 0.9 0.95 1 The robustness of the method against static errors in s the parameters, namely amplitude and relative phase of the fields, is discussed in Appendix C. The results show FIG. 7. Preparation of a two-qubit entangled state via the that the method is not very sensitive to errors: it behaves E–CD method (solid blue), compared against standard adia- linearly with respect to amplitude noise and quadrati- batic (dashed red). The results are zoomed in the rescaled– cally with respect to phase noise. time interval 0.7 ≤ s ≤ 1 for better visibility and the pa- rameters are ε = 5, τ = 5, T = 1/2. Above, the evolution of the population of the bare level |00i is shown, with the B. Two-qubit entanglement creation black dashed line indicating√ the target dynamics. The two eigenstates (|01i ± |10i)/ 2 are never populated, while the state |00i is adiabatically converted into the entangled su- The ability to reliably produce and manipulate entan- √ perposition (|00i + |11i)/ 2. Below, the infidelity is repre- glement is at the core of quantum information processing sented, reaching a final value of ∼ 0.1 for adiabatic driving [30]. High-fidelity entanglement preparing protocols are and 1.7 × 10−3 for E–CD driving. The total speed up is of thus a central touchstone for the development of quan- ∼ 12 times using an E–CD field HE with maximal strength tum technologies. These are typically difficult to realize, 1 satisfying S(HE) = 2 S(H) [see Eq. (21)]. especially due to the necessity of strong interactions and the ability of controlling many of them. 10

Here, we use the E–CD method for the high-fidelity 1 adiabatic preparation of a two-qubit entangled Bell state, 0.9 requiring time-dependent control of local terms and one 0.8 two-body interaction. Specifically, let us consider the Hamiltonian 0.7 1 0.6 0.95 H(s) = B(s)H1 + g(s)H2, (23) 0.5 0.9 (1) (2) (1) (2) (1) (2) = B(s)(σz + σz ) g(s)(σx σx + σz σz ). 0.4 0.5 0.51 0.52 − − Probability The Hamiltonian H1 is local, while H2 has entangled 0.3 eigenstates. Varying the field B(s) from high values to 0.2 very small values, while keeping g fixed, the adiabatic 0.1 path of the ground state connects a separable state 00 1 | i 0 to an entangled Bell state ψ+ = √ ( 00 + 11 ). In 0 0.2 0.4 0.6 0.8 1 | i 2 | i | i particular, we fix g = 1, choose the simple time depen- s dence B(s) = ε(1 s), and consider a total duration τ of the protocol. − FIG. 8. Comparison of E–CD and adiabatic methods for the The dynamical Lie algebra of the system has dimen- three-level system with Hamiltonian of Eq. (25) and parame- sion four and is isomorphic toLu(2), so the system is not ters τ = 25, ε = 40, d = 5/2. The evolution of the populations fully controllable, since su(4). However, from Theo- of the bare states is shown in dotted-solid red for adiabatic rem 1, we already knowL that ⊂ the Hamiltonian H will driving, by the blue solid line for E–CD driving, in dashed CD black for true (target) adiabatic dynamics. The inset high- be inside , and will thus be a combination of the four L lights the oscillation of the E–CD dynamics around the target basis elements. Since we know from Sec. II that HCD one. is orthogonal, with respect to the Hilbert-Schmidt inner product, both to H(t) and its time derivative, we also know that it will not have components along H1 and H2. was discussed in detail in Ref. [19]. The Hamiltonian, in Indeed, HCD can be computed analytically (Appendix terms of dimensionless quantities, reads D) and turns out to have the form d + ε s 1 1 0 1 ε − 2 τHCD(s) = H3 = τfCD(s)H3, (24) H(s) = 1 2d 1 . (25) 2 2 2 4ε (1 s) + 1  0 − 1 d ε s 1  − − − 2 (1) (2) (1) (2)   where H3 = σx σy +σy σx is a hopping term. There- This Hamiltonian can physically describe an effective fore, HCD would require the time-dependent control of an spin-1. For example, it could describe a local part of the extra two-body interaction. energy spectrum of a molecular nanomagnet [38], which is In order to avoid the additional implementation of H3, subject to a constant magnetic field along the x direction, we can use the results of Sec. IV A, since H3 [H1,H2]. ∝ a linearly sweeping magnetic field along the z direction, We thus apply an E–CD field of the form HE(s) = and an axial zero-field splitting [19]. The Hamiltonian is c1(s)H1 + c2(s)H2, with control functions real, so we are in the situation of Theorem 2. The CD field HCD cannot be computed analytically for all times, c1(s) = fCD(s) ω cos(ωsτ), − | | but has the general form c2(s) = p fCD(s) ω sin(ωsτ). | | (12) (13) 0 ifCD ifCD The outcomes of a numericalp simulation with parame- (12) − − (23) HCD = if 0 if , (26) ters ε = 5, τ = 5, NT = ωτ/2π = 10 are reported  CD CD  (13) (23) − in Fig. 7. A final infidelity of 1.7 10−3 is produced if if 0 ×  CD CD  with a speed up of 12 times, for HE having maximal   1∼ (12) (23) (13) strength (HE) (H) [see Eq. (17)]. These results with f and f always negative, while f assumes S ≤ 2 S CD CD CD are extremely favorable, showing that the E–CD method also positive values. The effective field is chosen of the provides a concrete advantage as a quantum control tool form for a difficult and important task such as entanglement creation. c3(t) c1(t) 0 HE = √ω c1(t) 0 c2(t) . (27)   0 c2(t) c3(t) − C. Three-level system   The control functions c(t) are chosen such that the (1) (2) We apply now the E–CD scheme to the case of a three- constraint equations given by MCD = ME can be easily level system undergoing a sequence of LZM avoided cross- solved analytically, ensuring precision in infidelity at least ing. The calculation and application of the exact CD field to order T 3 at the end of one period. This is done by 11 recalling, as observed in Sec. IV A, that only sines and Effective counterdiabatic Hamiltonians have been com- cosines of the same harmonics contribute to the first set puted for the prototype applications studied, by choos- of constraint equations. They read ing control functions which produce simple constraint equations for the amplitudes; see Secs. IV A and V. Re- c1(t) =A cos(ωt) B cos(2ωt), markably, this allowed us to solve the equations analyt- − c2(t) =C sin(ωt) D cos(3ωt), ically, and to obtain satisfying results without resorting − to heavy numerical methods. Nonetheless, the structure c3(t) =B sin(2ωt) + D sin(3ωt). of the E–CD problem still leaves room for optimization of the parameters, so further improvement can be predicted by the hybridization with optimal control strategies [25]. The constraint equations are Moreover, one could combine the E–CD method with a 1 T preselection of an optimal adiabatic protocol. For in- TB2 = f (12)(t)dt, (28a) stance, one might first conceive a “local” adiabatic driv- 4 − CD Z ing [39], in which the evolution is accelerated where nona- diabatic effects are small and then slowed down in the T vicinity of avoided crossings. This would provide an a 1 (23) TD2 = f (t)dt, (28b) priori improvement of the basic adiabatic protocol with 6 − CD Z respect, for example, to a simple LZM sweep. Such a protocol could then be ulteriorly sped up by means of 1 T E–CD corrections. T AC = f (13)(t)dt. (28c) 2 − CD In particular, the simplicity and versatility of the E– Z CD method makes it a strong strategy for the control of Approximating the integrals with the value of the inte- adiabatic processes. Therefore, we expect it to be of prac- grated functions multiplied by T , as explained in detail tical interest for many applications, ranging from quan- in Sec. IV A, an interpolated solution for Eqs. (28) is tum state preparation to adiabatic quantum computing given by [40, 41], quantum thermodynamics [42, 43], and probing of quantum critical dynamics [44–47]. A(t) = sgn[f (13)] 2 f (13) ,B(t) = 2 f (12) , The method has been developed for finite-dimensional − CD CD CD r r quantum systems. Still, the scheme could in principle be (13) (23) applied to infinite-dimensional systems with discrete lev- C(t) = 2 f ,D (t) = 6 f . (29) CD CD els, or restricted to well-separated nonpathological parts r r of the spectrum. Pathological here refers to possible de- The results with parameters τ = 25, ε = 40, d = 5/2 are generacies or continuous spectra. The first case, which is shown in Fig. 8. A speed up of 2.5 times is obtained, for problematic also for finite dimensions, would require ad- equal strength (HE) = (H) of the Hamiltonians. This dressing the decoupling of instantaneous eigensubspaces, S S shows that the E–CD paradigm behaves well also in a rather then single instantaneous eigenvectors. On the situation where more complex nonadiabatic phenomena other hand, continuous spectra would require a revision are present in the dynamics of the quantum system. of the CD framework, starting from the adiabatic the- orem itself [48]. Concerning the more mathematical re- sults, noncompactness of the dynamical Lie group of the VI. CONCLUSION system would break some of the assumptions used in our control-theoretic setting [49], which should then be in- We have introduced and discussed a method for vestigated more thoroughly. speeding-up adiabatic quantum state transfer without re- quiring the introduction of new Hamiltonian components, while remaining always close to the true adiabatic path. This is achieved by introducing a control Hamiltonian ACKNOWLEDGMENTS which simulates in real time the effect of a counterdia- batic field [3, 4], but can be adsorbed into the initial Hamiltonian H(t). As a trade-off, complete control of F.P. is grateful for hospitality at Imperial College Lon- H(t) is required. However, the algebraic results of Sec. don and acknowledges partial funding from the Doc- III can be exploited to limit the number of independent toral International Mobility Program of the University control Hamiltonians needed in H(t) to implement the of Parma. B.D. acknowledges funding from the Engi- control protocol. For instance, we have shown in Sec. V neering and Physical Sciences Research Council (UK) ad- that, by encoding a two-level problem into a two-qubit ministered by Imperial College London via the Postdoc- framework, a two-qubit entangled state can be prepared toral Prize Fellowship program. The authors are grate- with high fidelity in short time, requiring control of a ful to Riccardo Mannella for a critical reading of the single two-body interaction and a local field. manuscript. 12

Appendix A: Proof of Theorem 1 with 1 t t1 Before proceeding with the proof of Theorem 1, and R(t) = dt1 dt2[Hcorr(t1),H(t2)] − 2 0 0 to fix notation, let us recall the basic formulation of the Z Z 1 t t1 adiabatic theorem. dt1 dt2[H(t1),Hcorr(t2)] + ... − 2 Considering the Schr¨odinger equation, i~∂tU(t) = Z0 Z0 ∗ H(t)U(t), from initial time ti, let us parametrize the By expanding in Taylor series around the midpoint t = physical time t according to t ti = sτ, 0 s 1, t/2 of the (small) integration interval (0, t), see Eq. (B6) − ≤ ≤ where τ = tf ti is the total evolution time. The equa- in Appendix B, one can write tion becomes − 1 R(t) = [H(1) ,H(0)] [H(0) ,H(1)] t3 + o(t4), −12 corr − corr i~∂sUτ (s) = τH(s)Uτ (s). (A1) where we have used the notation  Theorem 3. (Adiabatic theorem [1, 2]) Let Uτ (s) be the k (k) 1 ∂ HX(t) solution of (A1). Then, HX = k . k! ∂ t ∗ t i R s 0 0 − τ En(t(s ))ds −1 3 U (s) e ~ 0 n(t(s)) n(0) = O(τ ), Therefore R(t) = o(t ). τ − | i h | n Now, considering small times t, and using repeatedly X (A2) the Zassenhaus formula [50] to factorize the exponential exp MCD(t) , UCD can be decomposed like In the adiabatic limit τ , and for s = 1, it holds { } 1 3 → ∞ † − [MCD(t),M(t)] o(t ) exactly UCD(t) = Ud(t)U (t)e 2 e .

→∞ i R t 0 0 τ − t En(t )dt The unitary matrix Ud(t), introduced in Sec. II, is the Uτ (1) Uad(ti, t) e ~ i n(t) n(0) . −→ ≡ | i h | solution of i∂tUd = HcorrUd. Proceeding as above, one n 1 2 X can estimate [MCD(t),M(t)] = o(t ). The adiabatic (A3) − 2 Under the control-theoretic setup introduced in Sec. theorem, Eqs. (A2) and (A3), then states that Ud(t) III, we proceed with the proof of Theorem 1. can be approximated to arbitrary precision through an adiabatic process of long duration τ,

2 Proof. Let HˆCD = HCD( tˆ) be the time-independent ma- † o(t ) −1 UCD(t) = Uad(t)U (t)e + o(τ ). (A6) trix obtained by evaluating HCD(t) at the time instant t = tˆ. In order to prove that HCD , one can equiv- We are then ready to prove the theorem. First of all, let ∈ L alently show that either (i) HCD can be written at all us observe that, for sufficiently small t, one can rewrite times as a linear combination of elements of a basis of , ˆ ˆ L ∂ t 1 t+t or (ii) that the group elements generated by the matrices HˆCD = HCD(t1)dt1 = HCD(t1)dt1 + o(t). − ˆ ∂tˆ t ˆ Hˆ for all time instants tˆ, e iHCDt, can be written as 0 t CD Z Z (A7) a concatenation of elements of the group eL. We follow Applying the exponential map to Eq. (A7) and using Eq. route (ii). With the initial time t set to zero and set to i ~ (A6) we eventually obtain one for simplicity of notation, let UCD(t) be the solution 2 −iHˆCDt o(t ) of the Schr¨odingerequation e = UCD(t,ˆ tˆ+ t)e , ˆ ˆ † ˆ ˆ o(t2) −1 i∂ UCD(t) = HCD(t)UCD(t) = Hcorr(t) H(t) UCD(t), = Uad(t, t + t)U (t, t + t)e + o(τ ). (A8) t − Since U and U can be realized via time dependent con- where the Hamiltonian H was introduced in Sec. II. ad corr trol of H(t), they belong to the group eL. Equation Let us assume that a Magnus expansion (see Appendix ˆ MCD(t) (A8) then means that, for all times t, the group element B) UCD(t) = e can be formally written, where exp( iHˆCDt) generated by iHˆCD can be approximated − − L t by a concatenation of elements of the group e , arbi- MCD(t) = i (Hcorr(t1) H(t1))dt1 trarily well for sufficiently small t and large τ. That is, − 0 − ˆ Z HCD for all tˆ. 1 t t1 ∈ L dt1 dt2[HCD(t1),HCD(t2)] + ... − 2 0 0 Z Z (A4) Appendix B: Magnus expansion

Now, calling Mcorr(t) and M(t) the Magnus exponents 1. Basics and useful bounds generated by Hcorr and H respectively, the terms in (A4) can be rearranged as to give The Magnus expansion [50, 51] is a representation of the evolution operator U(t), solution of the time- MCD(t) = Mcorr(t) M(t) + R(t), (A5) − dependent Schr¨odingerequation i~∂tU(t) = H(t)U(t), 13 in exponential form U(t) = exp M(t) . The exponent the error in the infidelity at the end of one period T of M(t) is an infinite sum M(t) = M{ (1)(t}) + M (2)(t) + ... the oscillating control functions, for the problem of ap- whose first terms read proximating UCD by means of UE, and for the class of t Hamiltonians characterized by Theorem 2 (see the end (1) i M (t) = H(t1)dt1, (B1) of Sec. IV). We will assume that the system starts in − ~ 0 the ground state gs(ti) at the initial time ti. Writing Z 2 | i i t t1 the states in terms of the evolution operators, and denot- M (2)(t) = dt dt [H(t ),H(t )], (B2) − 1 2 1 2 ing with the expectation value gs(ti) gs(ti) , the ~ 0 0   Z Z infidelityh·i can be written in the formh | · | i 3 t t1 (3) i M (t) = − dt1 dt2 † 2 ~ 0 0 F = 1 U UE . (B7)   Z Z I − |h CD i| t3 dt3 [H(t1), [H(t2),H(t3)]] We then write the propagators in terms of their Magnus 0 † −MCD(T ) ME(T ) Z  expansions, UCD(T ) = e and UE(T ) = e , + [H(t3), [H(t2),H(t1)]] . (B3) and use the Baker-Campbell-Hausdorff formula for com- †  puting the product UCDUE. Assuming that one has (k) solved the constraint equations, discussed in Sec. IV, We will in general denote with MX the k-th Magnus term for the Hamiltonian HX. The behavior of the Mag- involving the first 2m Magnus terms of UE, then it holds nus terms for small integration time t can be conveniently m 2m studied by Taylor-expanding the Hamiltonian around the (k) (k) M = M . midpoint of the integration interval, and computing the CD E k=1 k=1 integrals afterwards [51]. It turns out that, due to the X X time symmetry of the expansion, all terms are odd func- We can then write tions of t. In particular, one has † (2m+1) (2m+2) (m+1) U UE = exp M + M M M (2m)(t) = o(t2m+1); M (2m+1) = o(t2m+3). (B4) h CD i E E − CD D 1 n M (1),M (2m+1) (B8) Due to its usefulness in Appendix A, for the proof of The- − 2 CD E orem 1, we do the explicit calculation for the following 1 h (m+1) (2) i “generalized” Magnus term, which involves the commu- M ,M + ... . (B9) −2 CD E tator of two different matrices, h i  Let us remark that, since the oscillations in H are 1 t t1 E Ω2(t) = dt1 dt2[A(t1),B(t2)]. (B5) much faster than the timescales of HCD, the behavior 2 0 0 (2m) Z Z in Eq. (B4) does hold for MCD(T ), that is MCD (T ) = ∗ 2m+1 (2m+1) 2m+3 Let us denote with t = t/2 the midpoint, and let us o(T ) and MCD (T ) = o(T ). For the dy- introduce the notation namics UE, though, the integration is over a full period and the same is thus not true. Recalling the prefac- 1 ∂kA(x) 1 ∂kB(x) tor √ω in H , the E–CD Magnus terms are of order ak = k , bk = k , E k! ∂ x ∗ k! ∂ x ∗ t t M (m)(T ) = o(T m/2). Equation (B9) can then be ex- E ∗ k pressed like Inserting the Taylor expansions A(t) = k ak(t t ) ∗ k − and B(t) = bk(t t ) into Eq. (B5) one obtains k − P † (2m+1) m+1 UCDUE = exp ME + o T . (B10) P0,∞ + +2 h i 1 1 1 ( 1)k n D n oE Ω2(t) = − −  2 n + 1 k + n + 2 One can now expand the matrix exponential in Taylor Xk,n n series and take the expectation value termwise. Taking ( 1)n+1[1 ( 1)k+1] t n+k+2 the modulus squared, for a generic Hermitian matrix X, − − − [a , b ] , − k + 1 k k 2 it holds o   1 2 1 3 iXt 2 2 2 2 3 = [a0, b0]t + ([a1, b0] [a0, b1]) t e = 1 + t Re X + X + o(t ), (B11) 4 24 − |h i| h i |h i|   1 4 5 + ([a0, b2] + [a2, b0]) t + o(t ). (B6) Applying the same reasoning to Eq. (B10) and inserting 48 into Eq. (B7) one finally obtains

2 2 2. Error in the infidelity for the E–CD method (2m+1) (2m+1) 2m+ 3 F(T ) = Re M M + o(T 2 ), I − E − E    D E (B12) It is useful in our study to estimate the error when the expansion is truncated. In particular, we will focus on = o(T 2m+1). (B13) 14

100 improves and a region for greater values where it per- forms worse instead. The behavior is symmetric with 2 10− respect to the sign of δ for phase noise.

4 | 10− 0 Appendix D: CD field for the two-qubit problem

6 F/F 10−

− In this appendix, we report the explicit computation 1

| 8 10− of the CD field for the two-qubit problem described in Sec. V B. The initial Hamiltonian in dimensionless units, 10 10 Amplitude (δ > 0) see Eq. (23), is − Amplitude (δ < 0) 12 Relative phase 10− (1) (2) (1) (2) (1) (2) 5 4 3 2 1 0 H(s)/~ = ε(1 s)[σz + σz ] [σx σx + σz σz ]. 10− 10− 10− 10− 10− 10 − − − δ | | As a first step, we need to diagonalize H(s). First of all, H(s) can be partially diagonalized by writing it in the FIG. 9. Robustness of the E–CD method, Eq. (20), against (time-independent) basis imperfections in amplitude and relative phase of the control fields as considered in Eq. (C1). For the amplitude, the solid 01 + 10 10 01 line indicates the results for relative error δ > 0, while the 1 = 00 ; | i | i; | i − | i; 11 . B | i √2 √2 | i dashed line is used for δ < 0. In the case of the relative   phase, the results are completely symmetric with respect to Let Q be the change-of-basis matrix from basis 0 = the sign of δ (the two lines are overlapping). The dependence B on δ is linear for the amplitude for |δ| < 10−2. It is quadratic 00 , 01 , 10 , 11 to 1, thus having the kets in 1 instead for the relative phase. {|as columns.i | i | Thei | Hamiltoniani} B reads B

1 + 2ε(s 1) 0 0 1 − − 0 0− 0 Appendix C: Robustness . (D1)  0 0 2 0  We test the robustness of the E–CD method against  1 0 0 1 2ε(s 1)  − − − −  possible experimental imperfection for the LZM case dis- cussed in Sec. V A. We take into account possible errors We thus see that two levels are actually decoupled among in the amplitude and relative phase of the driving fields. themselves and from the rest of the spectrum. What re- This is done by adding a static relative offset δ. More mains to be diagonalized is a two-by-two real symmetric specifically, with reference to Eq. (20), matrix (formed by the four corner elements). For this we can use the usual convenient trigonometric parametriza- sin(ωsτ) (1 + δ) sin(ωsτ), tion of two-by-two unitary matrices [3], and the matrix → diagonalizing (D1) is sin(ωsτ) sin(ωsτ + 2πδ), (C1) → sin[θ(s)] 0 0 cos[θ(s)] with 1/2 δ 1/2. The quantity F = − − ≤2 ≤ 0 1 0 0 gs(tf ) ψ(tf ) is the final fidelity between system state P (s) = , | h | i |  0 0 1 0  ψ(t ) and target (ground) state gs(t ) , while F0 is the | f i | f i fidelity in the absence of errors. The behavior of the rel-  cos[θ(s)] 0 0 sin[θ(s)]   ative error 1 F/F0 is shown in Fig. 9 as a function of the offset δ|for− the two| error situations described by Eq. 1 1 with θ(s) = 2 arctan 2ε(1−s) . Eventually, the full diag- (C1). Both positive and negative values of δ are consid- onalizing matrix is Uh(s) = P (is)Q, and thus the CD field ered. The E–CD method results in being mostly linearly can be readily computed, recalling that s = (t t )/τ, stable against amplitude errors, while it turns out to be − i quadratically sensitive to phase errors. It should be re- i ∂U(s) † marked that in some cases small errors in amplitude can HCD(s) = U(s) , incidentally lead to a small improvement of the method: τ ∂s 1 ε this can be seen from the two regimes for pos- = [σ(1)σ(2) + σ(1)σ(2)]. −1 2τ 1 + 4ε2(s 1)2 x y y x itive δ. The negative peak around 10 indicates the − transition between a region δ < 10−1 where the protocol

[1] A. Messiah, (Dover Publications, [2] T. Kato, Journal of the Physical Society of Japan 5, 435 New York, 1961). (1950). 15

[3] M. Demirplak and S. A. Rice, J. Phys. Chem. A 107, A 87, 012116 (2013). 9937 (2003). [36] A. M. Dykhne, Sov. Phys. JETP 14, 941 (1962). [4] M. V. Berry, J. Phys. A 42, 365303 (2009). [37] J. P. Davis and P. Pechukas, The Journal of Chemical [5] M. G. Bason, M. Viteau, N. Malossi, P. Huillery, E. Ari- Physics 64, 3129 (1976). mondo, D. Ciampini, R. Fazio, V. Giovannetti, R. Man- [38] D. Collison, M. Murrie, V. S. Oganesyan, S. Piligkos, nella, and O. Morsch, Nature Phys. 8 (2012). N. R. J. Poolton, G. Rajaraman, G. M. Smith, A. J. [6] C. Jarzynski, Phys. Rev. A 88, 040101 (2013). Thomson, G. A. Timko, W. Wernsdorfer, R. E. P. Win- [7] M. Demirplak and S. A. Rice, J. Chem. Phys. 129, penny, and E. J. L. McInnes, Inorg. Chem. 42, 5293 154111 (2008). (2003). [8] S. Ib´a˜nez,X. Chen, E. Torrontegui, J. G. Muga, and [39] J. Roland and N. J. Cerf, Phys. Rev. A 65, 042308 (2002). A. Ruschhaupt, Phys. Rev. Lett. 109, 100403 (2012). [40] E. Farhi, J. Goldstone, S. Gutmann and M. M. Sipser, [9] E. Torrontegui, S. Mart´ınez-Garaot, and J. G. Muga, arXiv:quant-ph/0001106v1 (2000). Phys. Rev. A 89, 043408 (2014). [41] T. Albash and D. A. Lidar, Rev. Mod. Phys. 90, 015002 [10] S. Mart´ınez-Garaot,E. Torrontegui, X. Chen, and J. G. (2018). Muga, Phys. Rev. A 89, 053408 (2014). [42] R. Alicki and R. Kosloff, arXiv:1801.08314v2 (2018). [11] S. Deffner, C. Jarzynski, and A. del Campo, Phys. Rev. [43] A. del Campo, J. Goold, and M. Paternostro, Scientific X 4, 021013 (2014). Reports 4, 6208 (2014). [12] Y.-C. Li and X. Chen, Phys. Rev. A 94, 063411 (2016). [44] M. Greiner, O. Mandel, T. Esslinger, T. W. H¨ansch, and [13] A. Baksic, H. Ribeiro, and A. A. Clerk, Phys. Rev. Lett. I. Bloch, Nature 415, 39 (2002). 116, 230503 (2016). [45] A. Polkovnikov, Phys. Rev. B 72, 161201 (2005). [14] T. Opatrn´yand K. Mølmer, New Journal of Physics 16, [46] A. del Campo, M. M. Rams, and W. H. Zurek, Phys. 015025 (2014). Rev. Lett. 109, 115703 (2012). [15] L. D. Landau, Phys. Z. Sowjetunion 2, 46 (1932). [47] H. Saberi, T. Opatrn´y,K. Mølmer, and A. del Campo, [16] C. Zener, Proc. R. Soc. A 137, 696 (1932). Phys. Rev. A 90, 060301 (2014). [17] E. Majorana, Nuovo Cimento 9, 43 (1932). [48] M. Maamache and Y. Saadi, Phys. Rev. Lett. 101, [18] S. Shevchenko, S. Ashhab, and F. Nori, Physics Reports 150407 (2008). 492, 1 (2010). [49] M. G. Genoni, A. Serafini, M. S. Kim, and D. Burgarth, [19] M. Theisen, F. Petiziol, S. Carretta, P. Santini, and Phys. Rev. Lett. 108, 150501 (2012). S. Wimberger, Phys. Rev. A 96, 013431 (2017). [50] W. Magnus, Communications on Pure and Applied [20] S. Campbell and S. Deffner, Phys. Rev. Lett. 118, 100601 Mathematics 7, 649 (1954). (2017). [51] S. Blanes, F. Casas, J. Oteo, and J. Ros, Physics Reports [21] S. Deffner and S. Campbell, Journal of Physics A: Math- 470, 151 (2009). ematical and Theoretical 50, 453001 (2017). [22] D. D’Alessandro, Introduction to Quantum Control and Dynamics, Chapman & Hall/CRC Applied Mathematics & Nonlinear Science (CRC Press, Boca Raton, 2007). [23] S. G. Schirmer, H. Fu, and A. I. Solomon, Phys. Rev. A 63, 063410 (2001). [24] G. Dirr and U. Helmke, GAMM-Mitteilungen 31, 59 (2008). [25] S. J. Glaser, U. Boscain, T. Calarco, C. P. Koch, W. K¨ockenberger, R. Kosloff, I. Kuprov, B. Luy, S. Schirmer, T. Schulte-Herbr¨uggen, D. Sugny, and F. K. Wilhelm, The European Physical Journal D 69, 279 (2015). [26] V. Jurdjevic and H. J. Sussmann, Journal of Differential Equations 12, 313 (1972). [27] N. Khaneja and S. J. Glaser, Chemical Physics 267, 11 (2001). [28] A. Verdeny, L. Rudnicki, C. A. M¨uller, and F. Mintert, Phys. Rev. Lett. 113, 010501 (2014). [29] X. Chen, I. Lizuain, A. Ruschhaupt, D. Gu´ery-Odelin, and J. G. Muga, Phys. Rev. Lett. 105, 123003 (2010). [30] M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information (Cambridge University Press, Cambridge, UK, 2010). [31] S. Jansen, M.-B. Ruskai, and R. Seiler, Journal of Math- ematical Physics 48, 102111 (2007). [32] N. V. Vitanov, Phys. Rev. A 59, 988 (1999). [33] N. V. Vitanov and B. M. Garraway, Phys. Rev. A 53, 4288 (1996). [34] R. Lim and M. V. Berry, J. Phys. A 24, 3255 (1991). [35] N. Malossi, M. G. Bason, M. Viteau, E. Arimondo, R. Mannella, O. Morsch, and D. Ciampini, Phys. Rev.