This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Optimal shortcut-to-adiabaticity quantum control

C. L. Latune, D. Sugny, S. Guérin Laboratoire Interdisciplinaire Carnot de Bourgogne (ICB), UMR 6303 CNRS-Université Bourgogne Europe, 9 Av. A. Savary, BP 47 870, F-21078 DIJON, France
Abstract

We introduce a new class of Shortcut-To-Adiabaticity (STA) protocols with minimal energy expenditure. The control process produces the same transformation as a counterdiabatic drive, but at the lowest possible energy cost. We apply optimal control theory to analytically design the latter for a qubit. We discuss the robustness of this control scheme with respect to a standard STA approach.

I Introduction

Control of quantum systems is at the core of quantum applications and quantum technologies [1, 2, 3, 4, 5, 6]. A particularly well-known and effective method for designing control protocols is the Shortcut-To-Adiabaticity approach (STA) [7, 8, 9, 10]. STA is a generic term for various techniques that aim to ensure that the quantum system of interest follows a given adiabatic trajectory at an arbitrary speed via a Hamiltonian transformation. A practical approach is to consider an additional term in the Hamiltonian system, called counterdiabatic driving, to cancel out the non-adiabatic losses despite the finite duration of the process. This technique, which was first introduced in [11], later in [12, 13, 14, 15] and then in a quantum thermodynamic context [16, 17, 18, 19, 20, 21, 22] to avoid quantum friction [23, 24, 25], has gained much importance in adiabatic quantum computing [26], experimental state engineering [27], and quantum information processing [28], to name a few. STA techniques do not provide a definite scheme for accelerating the dynamics, since there are in principle infinitely many ways of defining an adiabatic trajectory, and require an ansatz typically based on physical considerations. On the other hand, optimal control theory (OCT) is a general mathematical procedure whose goal is to find time-dependent control parameters while minimizing or maximizing a functional that can be the control time-length or the energy used by the control, to name a few [29, 30, 2, 31]. The mathematical construction of OCT is based on the Pontryagin’s Maximum Principle (PMP) which was established in the late 1950s [32, 33, 34, 35, 36]. Today, OCT has become a powerful tool to optimize a variety of operations in quantum technologies [1, 2, 37].

STA provides a simple answer to perform a Hamiltonian transformation from a given Hamiltonian HiH_{i} to a final Hamiltonian HfH_{f} while following the adiabatic trajectory defined by a known protocol H0(t)H_{0}(t) where H0(0)=HiH_{0}(0)=H_{i} and H(tf)=HfH(t_{f})=H_{f}. Such a protocol may be motivated, for example, by some thermodynamic protocols such as quantum Otto or Carnot cycles [38], or by other requirements in quantum annealing processes, qubits resets, or even in adiabatic Grover search algorithm [39]. In the situation in which the initial and the final Hamiltonians do not commute with each other, the protocol H0(t)H_{0}(t) does not commute with itself at different times, and it may induce some transitions between different energy levels, leading in particular to quantum friction [23, 24, 25] and larger work expenditure. However, there are two natural situations that avoid such extra work: (i) slow driving so that the induced dynamics satisfies the quantum adiabatic theorem [40, 41, 42], and (ii) some adequately chosen initial non-passive states [43]. STA techniques offer an alternative as they allow fast transformations while still preserving the adiabatic trajectory defined by H0(t)H_{0}(t), but at the cost of adding an extra driving term V(t)V(t). In this framework, one of the widely used STA technique is the counter-diabatic drive [12, 13, 14, 15], which can be expressed as

VCD(t)=in[|n˙(t)n(t)|n(t)|n˙(t)|n(t)n(t)|],V_{\text{CD}}(t)=i\hbar\sum_{n}\big{[}|\dot{n}(t)\rangle\langle n(t)|-\langle n(t)|\dot{n}(t)\rangle|n(t)\rangle\langle n(t)|\big{]}, (1)

where |n(t)|n(t)\rangle denotes the instantaneous eigenstates of the Hamiltonian H0(t)H_{0}(t), H0(t)=nen(t)|n(t)n(t)|H_{0}(t)=\sum_{n}e_{n}(t)|n(t)\rangle\langle n(t)|. The effect of VCD(t)V_{\text{CD}}(t) is actually to cancel the transitions between the energy levels of H0(t)H_{0}(t). However, the additional drive VCD(t)V_{\text{CD}}(t) has to come with an additional energy cost [44, 45]. Several arguments have been put forward such as the quantum speed limit [46], qualitative estimate of the power needed to generate the control fields [47], or a connection with the classical entropy production generated during the generation of the control signal [48].

In this work, we define a new class of STA, where the additional driving is uniquely designed from the optimization of its energy cost. All the above propositions suggest a figure of merit of the form Wcostωi2α0tf𝑑uVCD(u)αW_{\text{cost}}\propto\omega_{i}^{2-\alpha}\int_{0}^{t_{f}}du||V_{\text{CD}}(u)||^{\alpha}, with α=1\alpha=1 or 2 and ωi\omega_{i} a typical frequency of the Hamiltonian H0H_{0}. Note that this additional energy cost is related to the power consumption of the devices used to control the quantum system, and is therefore usually much higher than the work cost associated with the Hamiltonian transformation (which is a cost at the quantum level). In addition, unnecessarily large controls applied to the quantum system can lead to extra dissipation and heating of the quantum system, which in turn leads to additional energy costs to cool the experimental setup [49].

Then, a natural question, of growing importance due to the intense debate on the energy cost of quantum technologies [50, 51], concerns the design of a STA protocol with minimal energy consumption. In other words, the goal is to find the minimum amount of energy required to implement an STA protocol. In this paper, we propose to solve this problem by using OCT. In the case of qubits, we highlight some processes where the energetically optimized STA is much less energetically expensive than the counter-diabatic drive. As expected, in the situation of very slow initial drive H0(t)H_{0}(t) (in a sense that will be made explicit later), both optimal and counter-diabatic drive become equal.

The paper is organized as follows. The optimization problem is formulated in Sec. II. In Sec. III, we show how the Pontryagin Maximum Principle can be applied and we derive the optimal equations. Formal solutions of these equations are given. Sections IV and V are dedicated to two examples, namely the case of a two-level quantum system with a constant energy gap and the Landau-Zener model. We discuss the similarities and differences of the energetically-optimized protocol and of the counter-adiabatic driving. A systematic analysis of the robustness with respect to the different Hamiltonian parameters is carried out in the two examples. A conclusion and prospective views are given in Sec. VI. Additional results are provided in the Appendix A.

II The optimization problem

In this work we focus on qubit systems. We consider arbitrary initial and final Hamiltonians, denoted by HiH_{i} and HfH_{f}, respectively, for the desired transformation. The Hamiltonians are parameterized in units where =1\hbar=1 as

Hi:=ωik=x,y,zui,kσk=ωi(|eiei||gigi|),H_{i}:=\omega_{i}\sum_{k=x,y,z}u_{i,k}\sigma_{k}=\omega_{i}(|e_{i}\rangle\langle e_{i}|-|g_{i}\rangle\langle g_{i}|), (2)

and

Hf:=ωfk=x,y,zuf,kσk=ωf(|efef||gfgf|),H_{f}:=\omega_{f}\sum_{k=x,y,z}u_{f,k}\sigma_{k}=\omega_{f}(|e_{f}\rangle\langle e_{f}|-|g_{f}\rangle\langle g_{f}|), (3)

with σk\sigma_{k}, k=x,y,zk=x,y,z, are the Pauli matrices and ui=(ui,x,ui,y,ui,z)\vec{u}_{i}=(u_{i,x},u_{i,y},u_{i,z}), uf=(uf,x,uf,y,uf,z)\vec{u}_{f}=(u_{f,x},u_{f,y},u_{f,z}) are unit vectors characterizing the initial and final Hamitonians. We also denote by |ei|e_{i}\rangle and |gi|g_{i}\rangle (|ef|e_{f}\rangle and |gf|g_{f}\rangle) the initial (final) excited and ground states, respectively. We consider a protocol H0(t)H_{0}(t) that satisfies H0(0)=HiH_{0}(0)=H_{i} and H0(tf)=HfH_{0}(t_{f})=H_{f}, i.e. H0(t)H_{0}(t) realizes the above Hamiltonian transformation and is given by the physics of the problem (imposed, e.g., by a thermodynamic or algorithmic protocol on a certain platform).

In most of quantum control applications, one is only interested in the final state and not in all intermediate states. On this basis, we derive an optimal driving Vopt(t)V_{\text{opt}}(t) allowing the exact connection between the initial and target Hamiltonians as the adiabatic or STA processes, but with no constraint on the instantaneous trajectory followed by the system, other than minimizing the energetic cost. In particular, this new class of STA does not suffer from the high energy expenditure of standard STA protocols based on a counterdiabatic driving that have to compensate the adiabatic losses at all times.

More specifically, the goal is to find a control protocol of the form

Vopt(t)=ωiv(t).σ=ωik=x,y,zvk(t)σk,V_{\text{opt}}(t)=\omega_{i}\vec{v}(t).\vec{\sigma}=\omega_{i}\sum_{k=x,y,z}v_{k}(t)\sigma_{k}, (4)

such that the eigenstates of HiH_{i}, |ei|e_{i}\rangle and |gi|g_{i}\rangle are brought to the corresponding eigenstates of HfH_{f}, |ef|e_{f}\rangle and |gf|g_{f}\rangle, the dynamics being governed by H(t)=H0(t)+Vopt(t)H(t)=H_{0}(t)+V_{\text{opt}}(t). In optimal control terminology, the original Hamiltonian, H0(t)H_{0}(t), can be interpreted as a time-dependent drift term that cannot be modified. Note that the final eigenstates can be reached up to a global phase factor. Then, starting from the state |ψ(0)=|ei|\psi(0)\rangle=|e_{i}\rangle, the target state is

|ψtraget=eiξf|ef,|\psi_{\text{traget}}\rangle=e^{i\xi_{f}}|e_{f}\rangle, (5)

where ξf\xi_{f} is an unspecified global phase. Finding controls such that the generated dynamic U(tf)U(t_{f}) brings |ψ(0)=|ei|\psi(0)\rangle=|e_{i}\rangle to |ψtarget|\psi_{\text{target}}\rangle automatically implies that U(tf)U(t_{f}) brings |gi|g_{i}\rangle to |gf|g_{f}\rangle up to a global phase. This is because U(tf)|ei=eiξf|efU(t_{f})|e_{i}\rangle=e^{i\xi_{f}}|e_{f}\rangle leads to ef|U(tf)|gi=0\langle e_{f}|U(t_{f})|g_{i}\rangle=0.

In addition, for any state ρi\rho_{i} commuting with the initial Hamiltonian, and for any state ρf\rho_{f} commuting with the final Hamiltonian HfH_{f}, it is also interesting to require that

Tr[ρiHi]=Tr[ρiH(0)]=Tr[ρiHi]+ωiTr[ρiv(0).σ]{\rm Tr}[\rho_{i}H_{i}]={\rm Tr}[\rho_{i}H(0)]={\rm Tr}[\rho_{i}H_{i}]+\omega_{i}{\rm Tr}[\rho_{i}\vec{v}(0).\vec{\sigma}] (6)

and

Tr[ρfHf]=Tr[ρfH(tf)]=Tr[ρfHf]+ωiTr[ρfv(tf).σ],{\rm Tr}[\rho_{f}H_{f}]={\rm Tr}[\rho_{f}H(t_{f})]={\rm Tr}[\rho_{f}H_{f}]+\omega_{i}{\rm Tr}[\rho_{f}\vec{v}(t_{f}).\vec{\sigma}], (7)

which means that the additional drive ωiv(t).σ\omega_{i}\vec{v}(t).\vec{\sigma} does not affect the initial (final) energy of such states. In other words, this additional constraint guarantees that if the initial and final states are diagonal in the initial and final Hamiltonian bases, the protocol H(t)H(t) does not contribute to the initial and final energy of the qubit. This property can also be found in some standard Shortcut-to-Adiabaticity protocols [46]. The two conditions can be expressed in terms of the control functions as

v(0).ui=0,\displaystyle\vec{v}(0).\vec{u}_{i}=0,
v(tf).uf=0.\displaystyle\vec{v}(t_{f}).\vec{u}_{f}=0.

In the reminder of the paper, we will use {|ei,|gi}\{|e_{i}\rangle,|g_{i}\rangle\}, the eigenbasis of HiH_{i} as the reference basis.

III Application of the Pontryagin Maximum Principle

An optimal solution satisfies the boundary conditions described in Sec. II while minimizing a cost functional. Since the goal here is to design a low-energy STA-like protocol, it is natural to consider an energy cost in the optimization process. As mentioned in the introduction, there have been several proposals to evaluate such a cost associated with a quantum control scheme. In this paper, following [48], we choose an energy cost defined by α=2\alpha=2, which leads to the following cost functional

𝒞=ωi20tf𝑑uv2(u).{\cal C}=\frac{\omega_{i}}{2}\int_{0}^{t_{f}}du\vec{v}^{2}(u). (8)

The corresponding optimal control can be designed by applying the PMP. In this approach, the optimal control problem is transformed into a generalized Hamiltonian system with specific boundary conditions. The time evolution of the control parameters is found by maximizing this Hamiltonian over the allowed controls. We refer the interested reader to recent reviews on the subject for details [2, 1, 35, 34].

In our control problem, the Pontryagin Hamiltonian can be written in the normal case as [34]

Hp=[χ(t)|H(t)|ψ(t)]ωi2v2(t),H_{p}=\Im[\langle\chi(t)|H(t)|\psi(t)\rangle]-\frac{\omega_{i}}{2}\vec{v}^{2}(t), (9)

where χ(t)|\langle\chi(t)| is the adjoint state of |ψ(t)|\psi(t)\rangle. The PMP states that the state and the adjoint state are solutions of the Schrödinger equation given by

d|ψ(t)dt\displaystyle\frac{d|\psi(t)\rangle}{dt} =\displaystyle= iH(t)|ψ(t),\displaystyle-iH(t)|\psi(t)\rangle, (10)
d|χ(t)dt\displaystyle\frac{d|\chi(t)\rangle}{dt} =\displaystyle= iH(t)|χ(t).\displaystyle-iH(t)|\chi(t)\rangle. (11)

Since there is no additional constraint on the controls v\vec{v}, the maximization condition on HpH_{p} gives Hp/v=0\partial H_{p}/\partial\vec{v}=0. The optimal controls denoted by v\vec{v}^{*} can then be expressed as

vk(t)=[χ(t)|σk|ψ(t)].v_{k}^{*}(t)=\Im[\langle\chi(t)|\sigma_{k}|\psi(t)\rangle]. (12)

Plugging them into the Hamiltonian H(t)H(t), we obtain (see details in Appendix A),

H(t)\displaystyle H^{*}(t) =\displaystyle= H0(t)iωi(|ψ(t)χ(t)||χ(t)ψ(t)|)\displaystyle H_{0}(t)-i\omega_{i}\Big{(}|\psi(t)\rangle\langle\chi(t)|-|\chi(t)\rangle\langle\psi(t)|\Big{)} (13)
ωi[χ(t)|ψ(t)].\displaystyle-\omega_{i}\Im[\langle\chi(t)|\psi(t)\rangle].

Using Eq. (13), the dynamical system satisfied by |ψ(t)|\psi(t)\rangle and |χ(t)|\chi(t)\rangle can be written as

d|ψ(t)dt\displaystyle\frac{d|\psi(t)\rangle}{dt} =\displaystyle= iH0(t)|ψ(t)ωi(d|ψ(t)|χ(t)),\displaystyle-iH_{0}(t)|\psi(t)\rangle-\omega_{i}\Big{(}d|\psi(t)\rangle-|\chi(t)\rangle\Big{)},
d|χ(t)dt\displaystyle\frac{d|\chi(t)\rangle}{dt} =\displaystyle= iH0(t)|χ(t)ωi(c|ψ(t)d|χ(t)),\displaystyle-iH_{0}(t)|\chi(t)\rangle-\omega_{i}\Big{(}c|\psi(t)\rangle-d|\chi(t)\rangle\Big{)},

where d:=[χ(t)|ψ(t)]d:=\Re[\langle\chi(t)|\psi(t)\rangle] and c:=χ(t)|χ(t)0c:=\langle\chi(t)|\chi(t)\rangle\geq 0. Since |ψ(t)|\psi(t)\rangle and |χ(t)|\chi(t)\rangle follow unitary evolutions, we deduce that ψ(t)|χ(t)\langle\psi(t)|\chi(t)\rangle and χ(t)|χ(t)\langle\chi(t)|\chi(t)\rangle, and thus cc, dd and s=[χ(t)|ψ(t)]s=\Im[\langle\chi(t)|\psi(t)\rangle] are constants of motion. Note that cc is a real positive number which can be different from 1 (|χ(t)|\chi(t)\rangle is not necessarily normalized), while dd and ss can be any real number. Then, introducing, |ψ~(t)=U0(t)|ψ(t)|\tilde{\psi}(t)\rangle=U_{0}^{\dagger}(t)|\psi(t)\rangle and |χ~(t)=U0(t)|χ(t)|\tilde{\chi}(t)\rangle=U_{0}^{\dagger}(t)|\chi(t)\rangle, with U0(t)U_{0}(t) the time evolution generated by H0(t)H_{0}(t), we can show that

d|ψ~(t)dt\displaystyle\frac{d|\tilde{\psi}(t)\rangle}{dt} =\displaystyle= ωi(d|ψ~(t)|χ~(t)),\displaystyle-\omega_{i}\Big{(}d|\tilde{\psi}(t)\rangle-|\tilde{\chi}(t)\rangle\Big{)}, (14)
d|χ~(t)dt\displaystyle\frac{d|\tilde{\chi}(t)\rangle}{dt} =\displaystyle= ωi(c|ψ~(t)d|χ~(t)).\displaystyle-\omega_{i}\Big{(}c|\tilde{\psi}(t)\rangle-d|\tilde{\chi}(t)\rangle\Big{)}. (15)

We can parameterize |χ~(0)=|χ(0)|\tilde{\chi}(0)\rangle=|\chi(0)\rangle in the eigenbasis {|ei,|gi}\{|e_{i}\rangle,|g_{i}\rangle\} of HiH_{i} as

|χ(0)=(d+is)|ei+r|gi.|\chi(0)\rangle=(d+is)|e_{i}\rangle+r|g_{i}\rangle. (16)

This yields the relation c=d2+s2+|r|2c=d^{2}+s^{2}+|r|^{2}. In general, |ψ~(t)|\tilde{\psi}(t)\rangle and |χ~(t)|\tilde{\chi}(t)\rangle are not orthogonal, which can lead to some additional difficulties in solving the dynamics. Therefore, we introduce the normalized vector |ϕ~(t)|\tilde{\phi}(t)\rangle orthogonal to |ψ~(t)|\tilde{\psi}(t)\rangle as

|ϕ~(t)\displaystyle|\tilde{\phi}(t)\rangle :=\displaystyle:= (|χ~(t)ψ~(t)|χ~(t)|ψ~(t))|χ~(t)ψ~(t)|χ~(t)|ψ~(t)\displaystyle\frac{\Big{(}|\tilde{\chi}(t)\rangle-\langle\tilde{\psi}(t)|\tilde{\chi}(t)\rangle|\tilde{\psi}(t)\rangle\Big{)}}{|||\tilde{\chi}(t)\rangle-\langle\tilde{\psi}(t)|\tilde{\chi}(t)\rangle|\tilde{\psi}(t)\rangle||} (17)
=\displaystyle= 1|r|(|χ~(t)(d+is)|ψ~(t)).\displaystyle\frac{1}{|r|}\Big{(}|\tilde{\chi}(t)\rangle-(d+is)|\tilde{\psi}(t)\rangle\Big{)}.

Note that |ϕ~(0)=|ϕ(0)=r|r||gi|\tilde{\phi}(0)\rangle=|\phi(0)\rangle=\frac{r}{|r|}|g_{i}\rangle. Using this new vector |ϕ~(t)|\tilde{\phi}(t)\rangle, the dynamical system becomes

d|ψ~(t)dt\displaystyle\frac{d|\tilde{\psi}(t)\rangle}{dt} =\displaystyle= ωi(is|ψ~(t)|r||ϕ~(t)),\displaystyle-\omega_{i}\Big{(}-is|\tilde{\psi}(t)\rangle-|r||\tilde{\phi}(t)\rangle\Big{)}, (18)
d|ϕ~(t)dt\displaystyle\frac{d|\tilde{\phi}(t)\rangle}{dt} =\displaystyle= ωi(|r||ψ~(t)+is|ϕ~(t)).\displaystyle-\omega_{i}\Big{(}|r||\tilde{\psi}(t)\rangle+is|\tilde{\phi}(t)\rangle\Big{)}. (19)

Now, we introduce the kets

|X±(t):=x±|ψ~(t)+y±|ϕ~(t),|X_{\pm}(t)\rangle:=x_{\pm}|\tilde{\psi}(t)\rangle+y_{\pm}|\tilde{\phi}(t)\rangle, (20)

such that

ddt|X±(t)\displaystyle\frac{d}{dt}|X_{\pm}(t)\rangle =λ±|X±(t).\displaystyle=\lambda_{\pm}|X_{\pm}(t)\rangle. (21)

Using Eqs. (18) and (19), one can show that we have to choose

x±y±=i|r|(s±s2+|r|2),\frac{x_{\pm}}{y_{\pm}}=\frac{i}{|r|}(s\pm\sqrt{s^{2}+|r|^{2}}), (22)

with

λ±=±iωis2+|r|2.\lambda_{\pm}=\pm i\omega_{i}\sqrt{s^{2}+|r|^{2}}. (23)

Note that only the ratio x±/y±x_{\pm}/y_{\pm} matters, or, in other words, we can choose y±=1y_{\pm}=1. Then, we have

|X±(t)=eλ±t|X±(0),|X_{\pm}(t)\rangle=e^{\lambda_{\pm}t}|X_{\pm}(0)\rangle, (24)

from which we deduce the time evolution of |ψ~(t)|\tilde{\psi}(t)\rangle and |ϕ~(t)|\tilde{\phi}(t)\rangle by inverting the relation (20),

|ψ~(t)\displaystyle|\tilde{\psi}(t)\rangle =\displaystyle= 1x+yxy+\displaystyle\frac{1}{x_{+}y_{-}-x_{-}y_{+}} (25)
×(yeλ+t|X+(0)y+eλt|X(0))\displaystyle\times\left(y_{-}e^{\lambda_{+}t}|X_{+}(0)\rangle-y_{+}e^{\lambda_{-}t}|X_{-}(0)\rangle\right)
=\displaystyle= 1x+/y+x/y\displaystyle\frac{1}{x_{+}/y_{+}-x_{-}/y_{-}}
×[(eλ+tx+/y+eλtx/y)|ei\displaystyle\times\Big{[}\left(e^{\lambda_{+}t}x_{+}/y_{+}-e^{\lambda_{-}t}x_{-}/y_{-}\right)|e_{i}\rangle
+(eλ+teλt)r|r||gi],\displaystyle\hskip 42.67912pt+\left(e^{\lambda_{+}t}-e^{\lambda_{-}t}\right)\frac{r}{|r|}|g_{i}\rangle\Big{]},
|ϕ~(t)\displaystyle|\tilde{\phi}(t)\rangle =\displaystyle= 1x+yxy+\displaystyle\frac{1}{x_{+}y_{-}-x_{-}y_{+}}
×(xeλ+t|X+(0)+x+eλt|X(0))\displaystyle\times\left(-x_{-}e^{\lambda_{+}t}|X_{+}(0)\rangle+x_{+}e^{\lambda_{-}t}|X_{-}(0)\rangle\right)
=\displaystyle= 1x+/y+x/y\displaystyle\frac{1}{x_{+}/y_{+}-x_{-}/y_{-}}
×[x+xy+y(eλ+teλt)|ei\displaystyle\times\Big{[}-\frac{x_{+}x_{-}}{y_{+}y_{-}}\left(e^{\lambda_{+}t}-e^{\lambda_{-}t}\right)|e_{i}\rangle
+(eλ+tx/y+eλtx+/y+)r|r||gi].\displaystyle+\left(-e^{\lambda_{+}t}x_{-}/y_{-}+e^{\lambda_{-}t}x_{+}/y_{+}\right)\frac{r}{|r|}|g_{i}\rangle\Big{]}.

From this result, we find the expression of |χ~(t)=|r||ϕ~(t)+(d+is)|ψ~(t)|\tilde{\chi}(t)\rangle=|r||\tilde{\phi}(t)\rangle+(d+is)|\tilde{\psi}(t)\rangle, and we can deduce also H(t)H(t) using expression Eq. (13) as well as the optimal control,

vk(t)\displaystyle v_{k}^{*}(t) =\displaystyle= [χ(t)|σk|ψ(t)]\displaystyle\Im[\langle\chi(t)|\sigma_{k}|\psi(t)\rangle] (27)
=\displaystyle= [χ~(t)|σk0(t)|ψ~(t)],\displaystyle\Im[\langle\tilde{\chi}(t)|\sigma_{k}^{0}(t)|\tilde{\psi}(t)\rangle],

where σk0(t):=U0(t)σkU0(t)\sigma_{k}^{0}(t):=U_{0}^{\dagger}(t)\sigma_{k}U_{0}(t). Note that in order to have an explicit expression of the controls vk(t)v_{k}(t), we compute the evolution generated by the original protocol H0(t)H_{0}(t), which is easily done numerically.

The solutions we have just derived are candidates for optimality. The last step of the procedure consists in finding the solutions that also satisfy the boundary conditions. To this aim, we choose adequately the free parameters (rr, dd and ss, or equivalently |χ(0)|\chi(0)\rangle) such that |ψ(tf)=eiξf|ef|\psi(t_{f})\rangle=e^{i\xi_{f}}|e_{f}\rangle, as well as Tr[ρ(Hi+ωiv(0).σ)]=Tr[ρHi]{\rm Tr}[\rho(H_{i}+\omega_{i}\vec{v}(0).\vec{\sigma})]={\rm Tr}[\rho H_{i}] for all state ρ=pe|eiei|+pg|gigi|\rho=p_{e}|e_{i}\rangle\langle e_{i}|+p_{g}|g_{i}\rangle\langle g_{i}| that commutes with the Hamiltonian HiH_{i}, and a similar condition for the final Hamiltonian. This implies Tr[ρVopt(0)]=ωiTr[ρv(0).σ]=0{\rm Tr}[\rho V_{{\rm opt}}(0)]=\omega_{i}{\rm Tr}[\rho\vec{v}(0).\vec{\sigma}]=0 from which we obtain

0\displaystyle 0 =\displaystyle= Tr[ρ(|ψ(0)χ(0)||χ(0)ψ(0)|)]+is\displaystyle{\rm Tr}[\rho(|\psi(0)\rangle\langle\chi(0)|-|\chi(0)\rangle\langle\psi(0)|)]+is
=\displaystyle= pe[ei|ψ(0)χ(0)|eiei|χ(0)ψ(0)|ei]+is\displaystyle p_{e}[\langle e_{i}|\psi(0)\rangle\langle\chi(0)|e_{i}\rangle-\langle e_{i}|\chi(0)\rangle\langle\psi(0)|e_{i}\rangle]+is
+pg[gi|ψ(0)χ(0)|gigi|χ(0)ψ(0)|gi]\displaystyle+p_{g}[\langle g_{i}|\psi(0)\rangle\langle\chi(0)|g_{i}\rangle-\langle g_{i}|\chi(0)\rangle\langle\psi(0)|g_{i}\rangle]
=\displaystyle= pe[dis(d+is)]+is=(2pe1)is\displaystyle p_{e}[d-is-(d+is)]+is=-(2p_{e}-1)is

where in the last line we use |ψ(0)=|ei|\psi(0)\rangle=|e_{i}\rangle. Then, the initial condition gives s=0s=0. Similarly, for the condition at final time tft_{f},we arrive at

0\displaystyle 0 =\displaystyle= pe[ef|ψ(tf)χ(tf)|efef|χ(tf)ψ(tf)|ef]+is\displaystyle p_{e}[\langle e_{f}|\psi(t_{f})\rangle\langle\chi(t_{f})|e_{f}\rangle-\langle e_{f}|\chi(t_{f})\rangle\langle\psi(t_{f})|e_{f}\rangle]+is
+pg[gf|ψ(tf)χ(tf)|gfgf|χ(tf)ψ(tf)|gf]\displaystyle+p_{g}[\langle g_{f}|\psi(t_{f})\rangle\langle\chi(t_{f})|g_{f}\rangle-\langle g_{f}|\chi(t_{f})\rangle\langle\psi(t_{f})|g_{f}\rangle]
=\displaystyle= pe[dis(d+is)]+is=(2pe1)is,\displaystyle p_{e}[d-is-(d+is)]+is=-(2p_{e}-1)is,

where the last line assumes that the drive has successfully implemented the expected dynamics, namely |ψ(tf)=eiξf|ef|\psi(t_{f})\rangle=e^{i\xi_{f}}|e_{f}\rangle, which also implies that ef|χ(tf)=eiξf(d+is)\langle e_{f}|\chi(t_{f})\rangle=e^{-i\xi_{f}}(d+is). Then, we obtain that the condition s=0s=0 guarantees that the additional drive does not affect the initial and final energies of the system (as long as the initial and the final states commute respectively with the initial and final Hamiltonians).

The last condition to take into account is |ψ(tf)=eiξf|ef|\psi(t_{f})\rangle=e^{i\xi_{f}}|e_{f}\rangle. Before that, we analyze geometrically the consequence of the condition s=0s=0 that simplifies the equations of motion. Indeed, s=0s=0 leads to

x±y±=±i,λ=±iωi|r|,\frac{x_{\pm}}{y_{\pm}}=\pm i,\ \lambda=\pm i\omega_{i}|r|,

and

|ψ~(t)\displaystyle|\tilde{\psi}(t)\rangle =\displaystyle= cos(ωi|r|t)|ei+eiϕrsin(ωi|r|t)|gi,\displaystyle\cos{(\omega_{i}|r|t)}|e_{i}\rangle+e^{i\phi_{r}}\sin{(\omega_{i}|r|t)}|g_{i}\rangle, (28)
|ϕ~(t)\displaystyle|\tilde{\phi}(t)\rangle =\displaystyle= sin(ωi|r|t)|ei+eiϕrcos(ωi|r|t)|gi,\displaystyle-\sin{(\omega_{i}|r|t)}|e_{i}\rangle+e^{i\phi_{r}}\cos{(\omega_{i}|r|t)}|g_{i}\rangle, (29)

with ϕr:=arg(r)\phi_{r}:=\arg(r), which corresponds to a rotation on the Bloch sphere about the axis ur=(cosϕr,sinϕr,0)\vec{u}_{r}=(-\cos\phi_{r},\sin\phi_{r},0) at angular velocity 2ωi|r|2\omega_{i}|r|.

Equivalently, the resulting Hamiltonian in the rotating picture with respect to H0(t)H_{0}(t) is

H~(t)\displaystyle\tilde{H}(t) =\displaystyle= V~opt(t):=U0(t)Vopt(t)U0(t)\displaystyle\tilde{V}_{\text{opt}}(t):=U_{0}(t)^{\dagger}V_{\text{opt}}(t)U_{0}(t)
=\displaystyle= iωi(|ψ~(t)χ~(t)||χ~(t)ψ~(t)|)\displaystyle-i\omega_{i}\Big{(}|\tilde{\psi}(t)\rangle\langle\tilde{\chi}(t)|-|\tilde{\chi}(t)\rangle\langle\tilde{\psi}(t)|\Big{)}
=\displaystyle= iωi(|ψ~(t)ϕ~(t)||ϕ~(t)ψ~(t)|).\displaystyle-i\omega_{i}\Big{(}|\tilde{\psi}(t)\rangle\langle\tilde{\phi}(t)|-|\tilde{\phi}(t)\rangle\langle\tilde{\psi}(t)|\Big{)}.

This gives, using the above expressions, the following time-independent Hamiltonian,

V~opt=ωi|r|[sin(ϕr)σx(i)+cos(ϕr)σy(i)],\tilde{V}_{\text{opt}}=\omega_{i}|r|\big{[}-\sin(\phi_{r})\sigma_{x}^{(i)}+\cos(\phi_{r})\sigma_{y}^{(i)}\big{]}, (30)

which also corresponds to the aforementioned rotation in the Bloch sphere, where

σx(i)\displaystyle\sigma_{x}^{(i)} :=\displaystyle:= |eigi|+|giei|,\displaystyle|e_{i}\rangle\langle g_{i}|+|g_{i}\rangle\langle e_{i}|,
σy(i)\displaystyle\sigma_{y}^{(i)} :=\displaystyle:= i|eigi|+i|giei|,\displaystyle-i|e_{i}\rangle\langle g_{i}|+i|g_{i}\rangle\langle e_{i}|,
σz(i)\displaystyle\sigma_{z}^{(i)} :=\displaystyle:= |eiei||gigi|,\displaystyle|e_{i}\rangle\langle e_{i}|-|g_{i}\rangle\langle g_{i}|,

are the Pauli matrices in the eigenbasis of HiH_{i}, {|ei,|gi}\{|e_{i}\rangle,|g_{i}\rangle\}. Note that the parameter dd, which characterizes the real part of the overlap between |ψ~(t)|\tilde{\psi}(t)\rangle and |χ~(t)|\tilde{\chi}(t)\rangle, has disappeared from the dynamics, and is therefore an irrelevant parameter (which can actually already be seen by combining Eqs. (14), (15), and (16)).

Then, we deduce that for a fixed final time tft_{f}, |ψ~(t)|\tilde{\psi}(t)\rangle can reach any state on the Bloch sphere by choosing adequately |r||r| and ϕr\phi_{r}, and in particular we can fulfill our final condition |ψ(tf)=eiξf|ef|\psi(t_{f})\rangle=e^{i\xi_{f}}|e_{f}\rangle (which is equivalent to |ψ~(tf)=eiξfU0(tf)|ef|\tilde{\psi}(t_{f})\rangle=e^{i\xi_{f}}U_{0}^{\dagger}(t_{f})|e_{f}\rangle). Denoting by kf:=ef|U0(tf)σk(i)U0(tf)|efk_{f}:=\langle e_{f}|U_{0}(t_{f})\sigma_{k}^{(i)}U_{0}^{\dagger}(t_{f})|e_{f}\rangle, for k=x,y,zk=x,y,z, the Bloch coordinates of eiξfU0(tf)|efe^{i\xi_{f}}U_{0}^{\dagger}(t_{f})|e_{f}\rangle in the eigenbasis of HiH_{i}, a direct geometrical analysis shows that the final condition |ψ(tf)=eiξf|ef|\psi(t_{f})\rangle=e^{i\xi_{f}}|e_{f}\rangle is satisfied if

cos(ϕr)sin(2ωi|r|tf)=xf,\displaystyle\cos(\phi_{r})\sin(2\omega_{i}|r|t_{f})=x_{f},
sin(ϕr)sin(2ωi|r|tf)=yf,\displaystyle\sin(\phi_{r})\sin(2\omega_{i}|r|t_{f})=y_{f},
cos(2ωi|r|tf)=zf,\displaystyle\cos{(2\omega_{i}|r|t_{f})}=z_{f},

which is equivalent to

|r|=arccos(zf)2ωitf,\displaystyle|r|=\frac{\arccos(z_{f})}{2\omega_{i}t_{f}},
(r)|r|=xf1zf2,\displaystyle\frac{\Re(r)}{|r|}=\frac{x_{f}}{\sqrt{1-z_{f}^{2}}},
(r)|r|=yf1zf2,\displaystyle\frac{\Im(r)}{|r|}=\frac{y_{f}}{\sqrt{1-z_{f}^{2}}},

and finally to

(r)\displaystyle\Re(r) =\displaystyle= xfarccos(zf)2ωitf1zf2,\displaystyle\frac{x_{f}\arccos(z_{f})}{2\omega_{i}t_{f}\sqrt{1-z_{f}^{2}}}, (31)
(r)\displaystyle\Im(r) =\displaystyle= yfarccos(zf)2ωitf1zf2.\displaystyle\frac{y_{f}\arccos(z_{f})}{2\omega_{i}t_{f}\sqrt{1-z_{f}^{2}}}. (32)

We have shown that, for any initial and final Hamiltonian and any protocol H0(t)H_{0}(t), we can solve our problem completely and exactly, as long as we can compute, at least numerically, the transformation U0(t)U_{0}(t) generated by H0(t)H_{0}(t). The explicit expression of the optimal controls is

vk(t)=|r|2Tr([sin(ϕr)σx(i)+cos(ϕr)σy(i)]U0(t)σkU0(t)),v_{k}(t)=\frac{|r|}{2}{\rm Tr}\Big{(}\big{[}-\sin(\phi_{r})\sigma_{x}^{(i)}+\cos(\phi_{r})\sigma_{y}^{(i)}\big{]}U_{0}^{\dagger}(t)\sigma_{k}U_{0}(t)\Big{)}, (33)

for k=x,y,zk=x,y,z.

Finally, we conclude this derivation by a comment on the global phase. When using the Bloch representation, this phase is automatically discarded. However, it can be seen that Eqs. (28) and (29) do not allow any control of the global phase. This is actually due to the choice s=0s=0. It can be shown that the global phase can indeed be controlled by adjusting adequately ss. However, this would imply the loss of the conservation of initial and final energy with respect to H0(0)H_{0}(0) and H0(tf)H_{0}(t_{f}).

IV The case of a two-level quantum system with a constant energy gap

IV.1 The energy cost

We consider a qubit for which Hi=ωiσzH_{i}=\omega_{i}\sigma_{z}, with a control protocol of constant energy gap given by H0(t)=ωi[cos(νt)σz+sin(νt)σx]H_{0}(t)=\omega_{i}[\cos(\nu t)\sigma_{z}+\sin(\nu t)\sigma_{x}], and a final Hamiltonian defined by the final time tft_{f}. For instance, when tf=π2νt_{f}=\frac{\pi}{2\nu}, the final Hamiltonian is Hf=ωiσxH_{f}=\omega_{i}\sigma_{x}. In the following, we leave tft_{f} unspecified, and define HfH_{f} as H0(tf)H_{0}(t_{f}).

We deduce that our target state, expressed in the eigenbasis of Hi=ωiσzH_{i}=\omega_{i}\sigma_{z}, is

|ψtarget=cos(νtf2)|1+sin(νtf2)|0,|\psi_{\text{target}}\rangle=\cos\left(\frac{\nu t_{f}}{2}\right)|1\rangle+\sin\left(\frac{\nu t_{f}}{2}\right)|0\rangle, (34)

denoting respectively by |1|1\rangle and |0|0\rangle the excited and ground states of σz\sigma_{z}. The values of (r)\Re(r) and (r)\Im(r) are given by Eqs. (31) and (32), with

kf=ψtarget|U0(tf)σkU0(tf)|ψtarget.k_{f}=\langle\psi_{\text{target}}|U_{0}(t_{f})\sigma_{k}U_{0}^{\dagger}(t_{f})|\psi_{\text{target}}\rangle. (35)

Note that since Hi=ωiσzH_{i}=\omega_{i}\sigma_{z}, the above Pauli matrices are expressed in the {|1,|0}\{|1\rangle,|0\rangle\} basis. The associated cost is given by Eq. (8) as

𝒞=ωi20tf𝑑uv2(u),{\cal C}=\frac{\omega_{i}}{2}\int_{0}^{t_{f}}du\vec{v}^{2}(u), (36)

which, after some manipulation, can be shown to be equal to

𝒞=12ωi|r|2tf.{\cal C}=\frac{1}{2}\omega_{i}|r|^{2}t_{f}. (37)

For comparison, we consider the counter-diabatic drive given by Eq. (1), which here leads to

VCD=ν2σy,V_{CD}=\frac{\nu}{2}\sigma_{y}, (38)

with the associated cost

𝒞CD\displaystyle{\cal C}_{CD} =\displaystyle= 14ωi0tf𝑑uHCD(u)2\displaystyle\frac{1}{4\omega_{i}}\int_{0}^{t_{f}}du||H_{CD}(u)||^{2} (39)
=\displaystyle= ν2tf8ωi.\displaystyle\frac{\nu^{2}t_{f}}{8\omega_{i}}.

For ν/ωi=0.5\nu/\omega_{i}=0.5 and tf=3π/4νt_{f}=3\pi/4\nu, we obtain

𝒞0.0051Vs𝒞CD0.147.{\cal C}\simeq 0.0051~{}~{}~{}\text{Vs}~{}~{}~{}{\cal C}_{CD}\simeq 0.147. (40)

The ratio is of the order of 30. This represents a significant reduction in energy consumption.

In Fig. 1(a), we represent on the Bloch sphere the trajectories of the excited eigenstates of the original Hamiltonian H0(t)H_{0}(t), of the counter-diabatic drive HCD(t)=H0(t)+VCDH_{CD}(t)=H_{0}(t)+V_{CD}, and of the energetically optimal protocol derived here H(t)=H0(t)+Vopt(t)H(t)=H_{0}(t)+V_{\text{opt}}(t). Figure 1(b) displays the dynamics of the qubit state when driven respectively by HCD(t)=H0(t)+VCDH_{CD}(t)=H_{0}(t)+V_{CD} and by H(t)=H0(t)+Vopt(t)H(t)=H_{0}(t)+V_{\text{opt}}(t). We can verify that both trajectories realize the expected transformation, from |ei|e_{i}\rangle to |ef|e_{f}\rangle. We can also see that the trajectory when driven by the counter-diabatic protocol does indeed follow the adiabatic trajectory.

(a) Refer to caption

(b) Refer to caption

Figure 1: Trajectories (a) of the excited eigenstates of the Hamiltonians H0(t)H_{0}(t) (blue), HCD(t)=H0(t)+VCDH_{CD}(t)=H_{0}(t)+V_{CD} (in red) and by H(t)=H0(t)+Vopt(t)H(t)=H_{0}(t)+V_{\text{opt}}(t) (in green); (b) of the trajectories of the qubit state when driven respectively by HCD(t)=H0(t)+VCDH_{CD}(t)=H_{0}(t)+V_{CD} (in red) and by H(t)=H0(t)+Vopt(t)H(t)=H_{0}(t)+V_{\text{opt}}(t) (in green). All plots are for ν/ωi=0.5\nu/\omega_{i}=0.5 and tf=3/4νt_{f}=3/4\nu.

We emphasize that the optimal protocol tends to the counter-diabatic process when νωi\nu\ll\omega_{i}, i.e. as expected in the adiabatic limit. In the highly non-adiabatic limit, when νωi\nu\gg\omega_{i}, we observe that while the optimal protocol and the counter-diabatic process differs significantly, their associated costs tend to be equal, as well as the induced dynamics.

Finally, we briefly mention another type of control procedure, the time-rescaled protocols [52, 53, 54], which consists in accelerating an adiabatic process. Indeed, if one takes the ”movie” of the adiabatic dynamics accelerated by a factor aa, the same final state is reached, but in a time tf/atf/a. Therefore, by taking aa large enough, one obtains an adiabatic dynamics on a short timescale, which can be viewed as an STA protocol. Mathematically, this corresponds to a rescaling of the time variable. However, as explained in [53], simply multiplying the time variable by a factor aa would change the initial and final Hamiltonians (they would also be multiplied by a factor aa). In order to avoid this issue, one can rescale the time with the function [53] f(t)=ata12πatfsin(2πatft)f(t)=at-\frac{a-1}{2\pi a}t_{f}\sin\left(\frac{2\pi a}{t_{f}}t\right). The resulting Hamiltonian is f˙(t)H(f(t))\dot{f}(t)H(f(t)). The time-rescaled process has two interesting properties: (i) it is straightforward to obtain the associated Hamiltonian, (ii) experimentally, there is no need to use any additional control. However, there is an important energetic drawback. Roughly speaking, the Hamiltonian is multiplied by aa, which means that the energy gaps are also multiplied by aa, so that the adiabatic theorem remains valid in the accelerated dynamics. However, this also implies that the intensity of all controls are also multiplied by aa. We illustrate this issue on the above example. If one simply takes the dynamics generated by H0(t)H_{0}(t), the fidelity between the final state U0(tf)|1U_{0}(t_{f})|1\rangle and the target state given in Eq. (34) is, |ψtarget|U0(tf)|1|2=0.95|\langle\psi_{\text{target}}|U_{0}(t_{f})|1\rangle|^{2}=0.95, which is actually not too bad since ν/ωi=0.5\nu/\omega_{i}=0.5 does not correspond to a strong non-adiabatic dynamics. To improve this final fidelity using a time-rescaled adiabatic process, we can choose a ratio ν/ωi=0.1\nu/\omega_{i}=0.1, and accelerate the protocol by a factor a=5a=5 to obtain the same duration for the energetically optimized protocol and the counter-diabatic drive. In this case, the final fidelity is 0.9960.996, but the cost, computed with 𝒞TR=14ωi0tf𝑑u(f˙(u)1)2H0(u)2{\cal C}_{TR}=\frac{1}{4\omega_{i}}\int_{0}^{t_{f}}du(\dot{f}(u)-1)^{2}||H_{0}(u)||^{2}, is 𝒞TR=56.5{\cal C}_{TR}=56.5, which is 4 orders of magnitude larger than the energetically optimized protocol. In order to reach the same level of final fidelity as the energetically optimized protocol and the counter-diabatic drive, one needs to consider a ratio ν/ωi=0.001\nu/\omega_{i}=0.001, and accelerates the adiabatic process by a factor a=500a=500. Then, the energy cost explodes, with 𝒞TR=8.8×105{\cal C}_{TR}=8.8\times 10^{5}.

IV.2 Robustness

In this section, we compare the robustness of the counter-diabatic drive and the optimal control with respect to several experimental uncertainties. We consider static errors, where the Hamiltonian parameter is not exactly known, but is in a given interval fixed by the experimental setup. In Fig. 2(a), we show the fidelity of the final state |ψ(tf)|\psi(t_{f})\rangle generated by the counter-diabatic drive and the optimal drive with respect to the target state Eq. (34) when there are some uncertainties on the frequency ν\nu. The experimental frequency is equal to ν(1+ϵ)\nu(1+\epsilon), while the control is designed for ϵ=0\epsilon=0. In Fig. 2(b), we plot the same fidelity for uncertainties both in ωi\omega_{i} and ν\nu.

(a) Refer to caption

(b) Refer to caption

Figure 2: Robustness against uncertainty in (a) ν\nu and (b) both ωi\omega_{i} and ν\nu for central values given by ν/ωi=0.5\nu/\omega_{i}=0.5 and tf=3π/4νt_{f}=3\pi/4\nu. The optimal protocl and the counter-diabatic drive are respectively plotted in green and in blue.

(a) Refer to caption

(b) Refer to caption

(c) Refer to caption

Figure 3: Robustness against uncertainty in (a) tft_{f}, (b) amplitude of the control functions vk(t)v_{k}(t), (c) amplitude of the control function having 3% of uncertainty in the other parameters. All plots are for central values given by ν/ωi=0.5\nu/\omega_{i}=0.5 and tf=3π/4νt_{f}=3\pi/4\nu.

In Fig. 3, we test the robustness with respect to other experimental parameters. In Fig. 3(a), we plot the fidelity with respect to uncertainties in the final time tft_{f}, meaning that the protocol is not stopped at tft_{f} but at tf(1+ϵ)t_{f}(1+\epsilon). In Fig. 3(b), we plot the fidelity for some uncertainties with respect to the amplitude to the control functions, (1+ϵ)vk(t)(1+\epsilon)v_{k}(t) instead of vk(t)v_{k}(t), and (1+ϵ)ν(1+\epsilon)\nu instead of ν\nu for the counter-diabatic drive. Finally, in Fig. 3(c), we also plot the final fidelity with respect to the target state when there are some uncertainties on the driving amplitudes, but assuming an uncertainty of 3%3\% in the other parameters.

It can be seen that the energetically optimal protocol is significantly more robust than the counter-diabatic protocol. However, with respect to uncertainty in ωi\omega_{i} only, the counterdiabatic drive is more robust than the energetically optimized protocol, which is simply because the counterdiabatic drive does not depend on ωi\omega_{i}. Note that when ωi\omega_{i} depends on time, then the counterdiabatic drive does depend on ωi\omega_{i}, but it seems it is still more robust with respect to ωi\omega_{i} than the energetically optimal protocol.

IV.3 Analytical expressions

One can derive some analytical expressions for uncertainties with respect to the amplitude of the control. For completeness, we present such expression in the following.

We compare the target state |ψtarget|\psi_{\text{target}}\rangle, which corresponds to the final state with the ideal amplitude of the control functions of the optimal protocol, with the final state when the amplitude is perturbed by a factor (1+ϵ)(1+\epsilon), denoted by |ψpert(tf)|\psi_{\text{pert}}(t_{f})\rangle. Then, we can show that, for the optimal protocol,

|ψtarget|ψpert(tf)|2=opt protocol\displaystyle|\langle\psi_{\text{target}}|\psi_{\text{pert}}(t_{f})\rangle|^{2}\underset{\text{opt protocol}}{=} (41)
1ϵ2tf2|ψ(0)|Vopt|ψ(0)|2\displaystyle 1-\epsilon^{2}t_{f}^{2}|\langle\psi(0)|V_{\text{opt}}|\psi_{\perp}(0)\rangle|^{2}
=\displaystyle= 1(ϵ2arccos(ef|U0(tf)σzU0(tf)|ef))2\displaystyle 1-\left(\frac{\epsilon}{2}\arccos(\langle e_{f}|U_{0}(t_{f})\sigma_{z}U_{0}^{\dagger}(t_{f})|e_{f}\rangle)\right)^{2}
=\displaystyle= 1(ϵ2arccos(zf))2,\displaystyle 1-\left(\frac{\epsilon}{2}\arccos(z_{f})\right)^{2},

where |ψ(0)=|ei|\psi(0)\rangle=|e_{i}\rangle is the initial state of the protocol (here |1|1\rangle), and |ψ(0)|\psi_{\perp}(0)\rangle an orthogonal state to |ψ(0)|\psi(0)\rangle (for instance |0|0\rangle).

We can do a similar computation for the counter-diabatic protocol. We obtain

|ψtarget|ψpert(tf)|2=CD protocol\displaystyle|\langle\psi_{\text{target}}|\psi_{\text{pert}}(t_{f})\rangle|^{2}\underset{\text{CD protocol}}{=}
1ϵ2|0tf𝑑tψ(0)|UCD(t)VCDUCD(t)|ψ(0)|2,\displaystyle 1-\epsilon^{2}\Big{|}\int_{0}^{t_{f}}dt\langle\psi(0)|U_{CD}^{\dagger}(t)V_{CD}U_{CD}(t)|\psi_{\perp}(0)\rangle\Big{|}^{2},

where UCD(t):=𝒯ei0t𝑑uHCD(u)U_{CD}(t):={\cal T}e^{-\frac{i}{\hbar}\int_{0}^{t}duH_{CD}(u)} is the dynamics generated by the counter-diabatic drive HCD(t)=H0(t)+VCDH_{CD}(t)=H_{0}(t)+V_{CD}. Having |ψ(0)=|1|\psi(0)\rangle=|1\rangle and |ψ(0)=|0|\psi_{\perp}(0)\rangle=|0\rangle, we have UCD(t)|ψ(0)=eiφe(t)|e(t)U_{CD}(t)|\psi(0)\rangle=e^{i\varphi_{e}(t)}|e(t)\rangle and UCD(t)|ψ(0)=eiφg(t)|g(t)U_{CD}(t)|\psi_{\perp}(0)\rangle=e^{i\varphi_{g}(t)}|g(t)\rangle, since HCD(t)H_{CD}(t) generates a dynamics following the adiabatic trajectory. However, there are the dynamical phases [7, 10] that we denoted by φe(t)\varphi_{e}(t) and φg(t)\varphi_{g}(t), respectively. Then, using the above expressions and VCD=ν2σyV_{CD}=\frac{\nu}{2}\sigma_{y}, we obtain

|ψtarget|ψpert(tf)|2=CD protocol\displaystyle|\langle\psi_{\text{target}}|\psi_{\text{pert}}(t_{f})\rangle|^{2}\underset{\text{CD protocol}}{=}
1ϵ2|0tf𝑑tei[φg(t)φe(t)](i)ν2|2.\displaystyle 1-\epsilon^{2}\Big{|}\int_{0}^{t_{f}}dte^{i[\varphi_{g}(t)-\varphi_{e}(t)]}(-i)\frac{\nu}{2}\Big{|}^{2}. (43)

We find an excellent agreement with the numerical simulations Fig. 3 (b).

V The case of the Landau-Zener model

V.1 Energy cost

As a second example, we consider the Landau-Zener model, which corresponds to a protocol realizing an adiabatic population transfer from the ground state to the excited state, characterized by a time-dependent energy gap [55, 56, 57, 58, 10],

H0(t)=ω(t)σz+Δσx.H_{0}(t)=\omega(t)\sigma_{z}+\Delta\sigma_{x}. (44)

As in the previous example, the optimal protocol is determined by

kf=ψtarget|U0(tf)σk(i)U0(tf)|ψtarget,k_{f}=\langle\psi_{\text{target}}|U_{0}(t_{f})\sigma_{k}^{(i)}U_{0}^{\dagger}(t_{f})|\psi_{\text{target}}\rangle, (45)

with σk(i)\sigma_{k}^{(i)} being the Pauli matrices in the initial energy eigenbasis, given here by

|ei\displaystyle|e_{i}\rangle =\displaystyle= cosθ2|1+sinθ2|0\displaystyle\cos\frac{\theta}{2}|1\rangle+\sin\frac{\theta}{2}|0\rangle
|gi\displaystyle|g_{i}\rangle =\displaystyle= sinθ2|1+cosθ2|0,\displaystyle-\sin\frac{\theta}{2}|1\rangle+\cos\frac{\theta}{2}|0\rangle,

with θ=arctanΔω(0)\theta=\arctan\frac{\Delta}{\omega(0)}. Then

Vopt(t)=ωiU0(t)((r)σx(i)+(r)σy(i))U0(t),V_{\text{opt}}(t)=\omega_{i}U_{0}(t)\left(-\Im(r)\sigma_{x}^{(i)}+\Re(r)\sigma_{y}^{(i)}\right)U_{0}^{\dagger}(t), (46)

with (r)\Im(r) and (r)\Re(r) given by Eqs. (31) and (32), and ωi=ω2(0)+Δ2.\omega_{i}=\sqrt{\omega^{2}(0)+\Delta^{2}}. By comparison, using Eq. (1), the counter-diabatic drive is given by  [10]

VCD(t)=ω˙(t)Δ2(ω2(t)+Δ2)σy.V_{CD}(t)=-\frac{\dot{\omega}(t)\Delta}{2(\omega^{2}(t)+\Delta^{2})}\sigma_{y}. (47)

As in the previous example, in Fig. 4(a), we show on the Bloch sphere the trajectories of the excited eigenstates of the original Hamiltonian H0(t)H_{0}(t), of the counter-diabatic drive HCD(t)=H0(t)+VCDH_{CD}(t)=H_{0}(t)+V_{CD}, and of the energetically optimal protocol derived here, H(t)=H0(t)+Vopt(t)H(t)=H_{0}(t)+V_{\text{opt}}(t), for the choice of a driving function of the form ω(t)=ω0+ωdtT\omega(t)=\omega_{0}+\omega_{d}\frac{t}{T}. In Fig. 4(b), we show the trajectories of the state of the qubit when driven respectively by HCD(t)=H0(t)+VCDH_{CD}(t)=H_{0}(t)+V_{CD} and by H(t)=H0(t)+Vopt(t)H(t)=H_{0}(t)+V_{\text{opt}}(t). We can verify that both trajectories realize the expected transformation, from |ei|e_{i}\rangle to |ef|e_{f}\rangle. We can also see that the trajectory when driven by the counter-diabatic protocol indeed follows the adiabatic trajectory.

(a) Refer to caption

(b) Refer to caption

Figure 4: Trajectories (a) of the excited eigenstates of the Hamiltonians H0(t)H_{0}(t) (in blue), HCD(t)=H0(t)+VCDH_{CD}(t)=H_{0}(t)+V_{CD} (in red) and by H(t)=H0(t)+Vopt(t)H(t)=H_{0}(t)+V_{\text{opt}}(t) (in green); (b) of the trajectories of the qubit state when driven respectively by HCD(t)=H0(t)+VCDH_{CD}(t)=H_{0}(t)+V_{CD} (in red) and by H(t)=H0(t)+Vopt(t)H(t)=H_{0}(t)+V_{\text{opt}}(t) (in green). We use ω(t)=ω0+ωdt/T\omega(t)=\omega_{0}+\omega_{d}t/T with ω0/Δ=10\omega_{0}/\Delta=-10, ωd/Δ=20\omega_{d}/\Delta=20 and tf/Δ=T/Δ=1t_{f}/\Delta=T/\Delta=1.

For the energy cost of the optimal protocol, we obtain the same expression as in the previous example, namely

𝒞=12ωi|r|2tf,{\cal C}=\frac{1}{2}\omega_{i}|r|^{2}t_{f}, (48)

and for the counter-diabatic drive we arrive at

𝒞CD=12ωi0tf𝑑tΔ2[v˙(t)]24(ω2(t)+Δ2)2.{\cal C}_{CD}=\frac{1}{2\omega_{i}}\int_{0}^{t_{f}}dt\frac{\Delta^{2}[\dot{v}(t)]^{2}}{4(\omega^{2}(t)+\Delta^{2})^{2}}. (49)

Taking the settings of Fig. 4, ω(t)=ω0+ωdt/T\omega(t)=\omega_{0}+\omega_{d}t/T with ω0/Δ=10\omega_{0}/\Delta=-10, ωd/Δ=20\omega_{d}/\Delta=20 and tf/Δ=T/Δ=1t_{f}/\Delta=T/\Delta=1, we get

𝒞opt\displaystyle{\cal C}_{\text{opt}} \displaystyle\simeq 0.07,\displaystyle 0.07,
𝒞CD\displaystyle{\cal C}_{CD} \displaystyle\simeq 0.39.\displaystyle 0.39.

Note that the optimal protocol has an additional advantage, i.e. the control amplitudes are much smaller than the counter-diabatic drive (see Fig. 5).

Refer to caption
Figure 5: Amplitudes of the control functions for the optimal protocol and counter-diabatic drive for ω(t)=ω0+ωdt/T\omega(t)=\omega_{0}+\omega_{d}t/T with ω0/Δ=10\omega_{0}/\Delta=-10, ωd/Δ=20\omega_{d}/\Delta=20 and tf/Δ=T/Δ=1t_{f}/\Delta=T/\Delta=1.

Additionally, for other choices of the parameters ω0/Δ\omega_{0}/\Delta, ωd/Δ\omega_{d}/\Delta and t/Tt/T, we can have a very significant reduction of energy cost, with 𝒞CD/𝒞opt{\cal C}_{CD}/{\cal C}_{\text{opt}} of the order of 200 or more. However, in such situations, although the energetically optimized protocol is order of magnitude more energetically efficient, it is also significantly less robust.

V.2 Robustness

As for the previous example, it is interesting to compare the robustness of the optimal protocol with the one of the counter-diabatic drive. For the robustness with respect to uncertainty in the amplitude of the controls, we can obtain the same analytical expressions as previously, namely Eqs. (41) and (IV.3), which leads to a very good agreement with the numerical plot in Fig. 6(a). In Fig. 6(b), we represent the robustness of the protocols with respect to uncertainty in the amplitude of the parameter vdv_{d} (which implies that the counter-diabatic drive and the optimal protocol are determined with a value of vdv_{d} which is not the exact one). We observe that for these uncertainties, the energetically optimal protocol is approximately as robust as the counterdiabatic drive. In Fig. 7, we show the uncertainty with respect to (a) Δ\Delta and (b) the duration of the operation tft_{f}. For these two kinds of uncertainty, one can see that the counter-diabatic drive is more robust than the energetically optimized one.

(a) Refer to caption

(b) Refer to caption

Figure 6: Robustness of the counter-diabatic drive (in blue) compared with the robustness of the optimal protocol (in green), for uncertainty (a) in the amplitude of the control; (b) in the amplitude of H0(t)H_{0}(t). We used ω(t)=ω0+ωdt/T\omega(t)=\omega_{0}+\omega_{d}t/T with ω0/Δ=0\omega_{0}/\Delta=0, ωd/Δ=10\omega_{d}/\Delta=10 and tf/Δ=T/Δ=1t_{f}/\Delta=T/\Delta=1.

(a) Refer to caption

(b) Refer to caption

Figure 7: Robustness of the counter-diabatic drive (in blue) compared with the robustness of the optimal protocol (in green), for uncertainty (a) in the Hamiltonian parameter Δ\Delta; (b) in the duration of the operation tft_{f}. We use ω(t)=ω0+ωdt/T\omega(t)=\omega_{0}+\omega_{d}t/T with ω0/Δ=0\omega_{0}/\Delta=0, ωd/Δ=10\omega_{d}/\Delta=10 and tf/Δ=T/Δ=1t_{f}/\Delta=T/\Delta=1.

VI Conclusion

For an arbitrary time-dependent qubit Hamiltonian H0(t)H_{0}(t), we introduce an energetically-optimized protocol that reproduces the final state of the adiabatic trajectory associated with H0(t)H_{0}(t). We use the figure of merit suggested in [48] to quantify the energy cost of the quantum control. Analytical solutions to the optimal control problem can be derived by applying the PMP. We compare the energy consumed by the optimal protocol with that of the counter-diabatic drive in the Landau-Zener model and a model with a constant energy gap. The energy difference can be very significant, reaching several orders of magnitude. We also briefly consider another STA technique, namely time-rescaling of the adiabatic process. Although it has practical advantages, its energetic bill is several orders of magnitude higher than the energetically optimised one. Finally, we compare the robustness of the optimal protocol with the counter-diabatic drive. We find that the optimised protocol is indeed more robust than the counter-diabatic drive for many, but not all, experimental uncertainties. This raises the question of whether the optimal procedure can be made even more robust by design, as suggested in [59, 60, 61]. This will be the focus of future work. Other perspectives are to extend the present framework to systems of arbitrary dimensions, and to consider a reduced number of control functions.

Acknowledgment
C.L.L. acknowledges funding from the French National Research Agency (ANR) under grant ANR-23-CPJ1- 0030-01. The research work of D. Sugny has been supported by the ANR project “QuCoBEC” ANR-22-CE47-0008-02 and by the ANR-DFG ”CoRoMo” Projects No. 505622963/KO 2301/15-1 and No. ANR-22-CE92-0077-01.

Appendix A Derivation of the optimal protocol

Using the Pontryagin Maximal Principle, the optimal control functions can be expressed as

vk(t)=[χ(t)|σk|ψ(t)],v_{k}^{*}(t)=\Im[\langle\chi(t)|\sigma_{k}|\psi(t)\rangle], (50)

which leads to an optimal protocol given by

Vopt(t)\displaystyle V_{\text{opt}}(t) =\displaystyle= ωik=x,y,z[χ(t)|σk|ψ(t)]σk\displaystyle\omega_{i}\sum_{k=x,y,z}\Im[\langle\chi(t)|\sigma_{k}|\psi(t)\rangle]\sigma_{k}
=\displaystyle= ωi2ik=x,y,z(Tr[|ψ(t)χ(t)|σk]\displaystyle\frac{\omega_{i}}{2i}\sum_{k=x,y,z}\Big{(}{\rm Tr}[|\psi(t)\rangle\langle\chi(t)|\sigma_{k}]
Tr[|χ(t)ψ(t)|σk])σk\displaystyle-{\rm Tr}[|\chi(t)\rangle\langle\psi(t)|\sigma_{k}]\Big{)}\sigma_{k}

The next step consists in using the property that {𝕀,σx,σy,σz}\{\mathbb{I},\sigma_{x},\sigma_{y},\sigma_{z}\} is a basis of the operator vectorial space acting on \mathcal{H}, and that any operator MM can be decompsed as M=12Tr[M]𝕀+12Tr[Mσx]σx+12Tr[Mσy]σy+12Tr[Mσz]σzM=\frac{1}{2}{\rm Tr}[M]\mathbb{I}+\frac{1}{2}{\rm Tr}[M\sigma_{x}]\sigma_{x}+\frac{1}{2}{\rm Tr}[M\sigma_{y}]\sigma_{y}+\frac{1}{2}{\rm Tr}[M\sigma_{z}]\sigma_{z}. Using this relation for M=|ψχ|M=|\psi\rangle\langle\chi|, we obtain

Vopt(t)\displaystyle V_{\text{opt}}(t) =\displaystyle= ωii(|ψ(t)χ(t)|12χ(t)|ψ(t)\displaystyle\frac{\omega_{i}}{i}\Big{(}|\psi(t)\rangle\langle\chi(t)|-\frac{1}{2}\langle\chi(t)|\psi(t)\rangle
|χ(t)ψ(t)|+12ψ(t)|χ(t))\displaystyle-|\chi(t)\rangle\langle\psi(t)|+\frac{1}{2}\langle\psi(t)|\chi(t)\rangle\Big{)}
=\displaystyle= iωi(|ψ(t)χ(t)|\displaystyle-i\omega_{i}\Big{(}|\psi(t)\rangle\langle\chi(t)|
|χ(t)ψ(t)|)ωi[χ(t)|ψ(t)],\displaystyle-|\chi(t)\rangle\langle\psi(t)|\Big{)}-\omega_{i}\Im[\langle\chi(t)|\psi(t)\rangle],

which is the expression given in Eq. (13).

References

  • Glaser et al. [2015] S. J. Glaser, U. Boscain, T. Calarco, C. P. Koch, W. Köckenberger, R. Kosloff, I. Kuprov, B. Luy, S. Schirmer, T. Schulte-Herbrüggen, D. Sugny, and F. K. Wilhelm, Training Schrödinger’s cat: quantum optimal control, Eur. Phys. J. D 69, 1 (2015).
  • Koch et al. [2022] C. P. Koch, U. Boscain, T. Calarco, G. Dirr, S. Filipp, S. J. Glaser, R. Kosloff, S. Montangero, T. Schulte-Herbrüggen, D. Sugny, and F. K. Wilhelm, Quantum optimal control in quantum technologies. strategic report on current status, visions and goals for research in europe, EPJ Quantum Technology 9, 19 (2022).
  • Brif et al. [2010] C. Brif, R. Chakrabarti, and H. Rabitz, Control of quantum phenomena: past, present and future, New Journal of Physics 12, 075008 (2010).
  • Altafini and Ticozzi [2012] C. Altafini and F. Ticozzi, Modeling and control of quantum systems: An introduction, IEEE Trans. Automat. Control 57, 1898 (2012).
  • Dong and Petersen [2010] D. Dong and I. A. Petersen, Quantum control theory and applications: A survey, IET Control Theory A 4, 2651 (2010).
  • Koch et al. [2019] C. P. Koch, M. Lemeshko, and D. Sugny, Quantum control of molecular rotation, Rev. Mod. Phys. 91, 035005 (2019).
  • Guéry-Odelin et al. [2019] D. Guéry-Odelin, A. Ruschhaupt, A. Kiely, E. Torrontegui, S. Martínez-Garaot, and J. G. Muga, Shortcuts to adiabaticity: Concepts, methods, and applications, Rev. Mod. Phys. 91, 045001 (2019).
  • Stefanatos and Paspalakis [2021] D. Stefanatos and E. Paspalakis, A shortcut tour of quantum control methods for modern quantum technologies, Europhysics Letters 132, 60001 (2021).
  • Torrontegui et al. [2013] E. Torrontegui, S. Ibáñez, S. Martínez-Garaot, M. Modugno, A. del Campo, D. Guéry-Odelin, A. Ruschhaupt, X. Chen, and J. G. Muga, Chapter 2 - shortcuts to adiabaticity, in Advances in Atomic, Molecular, and Optical Physics, Advances In Atomic, Molecular, and Optical Physics, Vol. 62, edited by E. Arimondo, P. R. Berman, and C. C. Lin (Academic Press, 2013) pp. 117–169.
  • Duncan et al. [2025] C. W. Duncan, P. M. Poggi, M. Bukov, N. T. Zinner, and S. Campbell, Taming quantum systems: A tutorial for using shortcuts-to-adiabaticity, quantum optimal control, and reinforcement learning, arXiv 10.48550/arXiv.2501.16436 (2025), 2501.16436 .
  • Unanyan et al. [1997] R. G. Unanyan, L. P. Yatsenko, K. Bergmann, and B. W. Shore, Laser-induced adiabatic atomic reorientation with control of diabatic losses, Opt. Commun. 139, 48 (1997).
  • Demirplak and Rice [2003] M. Demirplak and S. A. Rice, Adiabatic Population Transfer with Control Fields, ACS Publications  (2003).
  • Demirplak and Rice [2005] M. Demirplak and S. A. Rice, Assisted Adiabatic Passage Revisited†, ACS Publications  (2005).
  • Demirplak and Rice [2008] M. Demirplak and S. A. Rice, On the consistency, extremal, and global properties of counterdiabatic fields, J. Chem. Phys. 129, 154111 (2008).
  • Berry [2009] M. V. Berry, Transitionless quantum driving, J. Phys. A: Math. Theor. 42, 365303 (2009).
  • Campo et al. [2014] A. d. Campo, J. Goold, and M. Paternostro, More bang for your buck: Super-adiabatic quantum engines, Sci. Rep. 4, 1 (2014).
  • Deng et al. [2013] J. Deng, Q.-h. Wang, Z. Liu, P. Hänggi, and J. Gong, Boosting work characteristics and overall heat-engine performance via shortcuts to adiabaticity: Quantum and classical systems, Phys. Rev. E 88, 062122 (2013).
  • Beau et al. [2016] M. Beau, J. Jaramillo, and A. Del Campo, Scaling-Up Quantum Heat Engines Efficiently via Shortcuts to Adiabaticity, Entropy 18, 168 (2016).
  • Abah and Paternostro [2019a] O. Abah and M. Paternostro, Shortcut-to-adiabaticity Otto engine: A twist to finite-time thermodynamics, Phys. Rev. E 99, 022110 (2019a).
  • Hartmann et al. [2020] A. Hartmann, V. Mukherjee, W. Niedenzu, and W. Lechner, Many-body quantum heat engines with shortcuts to adiabaticity, Phys. Rev. Res. 2, 023145 (2020).
  • Dann and Kosloff [2020] R. Dann and R. Kosloff, Quantum signatures in the quantum Carnot cycle, New J. Phys. 22, 013055 (2020).
  • Deng et al. [2018] S. Deng, A. Chenu, P. Diao, F. Li, S. Yu, I. Coulamy, A. del Campo, and H. Wu, Superadiabatic quantum friction suppression in finite-time thermodynamics, Sci. Adv. 410.1126/sciadv.aar5909 (2018).
  • Kosloff and Feldmann [2002] R. Kosloff and T. Feldmann, Discrete four-stroke quantum heat engine exploring the origin of friction, Phys. Rev. E 65, 055102 (2002).
  • Feldmann and Kosloff [2003] T. Feldmann and R. Kosloff, Quantum four-stroke heat engine: Thermodynamic observables in a model with intrinsic friction, Phys. Rev. E 68, 016101 (2003).
  • Feldmann and Kosloff [2004] T. Feldmann and R. Kosloff, Characteristics of the limit cycle of a reciprocating quantum heat engine, Phys. Rev. E 70, 046110 (2004).
  • Hegade et al. [2021] N. N. Hegade, K. Paul, Y. Ding, M. Sanz, F. Albarrán-Arriagada, E. Solano, and X. Chen, Shortcuts to Adiabaticity in Digitized Adiabatic Quantum Computing, Phys. Rev. Appl. 15, 024038 (2021).
  • Chen et al. [2021] Y.-H. Chen, W. Qin, X. Wang, A. Miranowicz, and F. Nori, Shortcuts to Adiabaticity for the Quantum Rabi Model: Efficient Generation of Giant Entangled Cat States via Parametric Amplification, Phys. Rev. Lett. 126, 023602 (2021).
  • Santos et al. [2020] A. C. Santos, A. Nicotina, A. M. Souza, R. S. Sarthour, I. S. Oliveira, and M. S. Sarandy, Optimizing NMR quantum information processing via generalized transitionless quantum driving, Europhys. Lett. 129, 30008 (2020).
  • Liberzon [2012] D. Liberzon, Calculus of variations and optimal control theory (Princeton University Press, Princeton, NJ, 2012) pp. xviii+235.
  • D’Alessandro [2008] D. D’Alessandro, Introduction to quantum control and dynamics. (Applied Mathematics and Nonlinear Science Series. Boca Raton, FL: Chapman, Hall/CRC., 2008).
  • Kirk [2004] D. E. Kirk, Optimal control theory: an introduction (Courier Corporation, New York, 2004).
  • Pontryagin et al. [1962] L. S. Pontryagin, V. Boltianski, R. Gamkrelidze, and E. Mitchtchenko, The Mathematical Theory of Optimal Processes (John Wiley and Sons, New York, 1962).
  • Lee and Markus [1967] M. M. Lee and L. Markus, Foundations of Optimal Control Theory (John Wiley and Sons, New York, 1967).
  • Ansel et al. [2024] Q. Ansel, E. Dionis, F. Arrouas, B. Peaudecerf, S. Guérin, D. Guéry-Odelin, and D. Sugny, Introduction to theoretical and experimental aspects of quantum optimal control, Journal of Physics B: Atomic, Molecular and Optical Physics 57, 133001 (2024).
  • Boscain et al. [2021] U. Boscain, M. Sigalotti, and D. Sugny, Introduction to the Pontryagin Maximum Principle for Quantum Optimal Control, PRX Quantum 2, 030203 (2021).
  • Bonnard and Sugny [2012] B. Bonnard and D. Sugny, Optimal Control with Applications in Space and Quantum Dynamics, AIMS on applied mathematics, Vol. 5 (American Institute of Mathematical Sciences, Springfield, 2012).
  • Dupont et al. [2021] N. Dupont, G. Chatelain, L. Gabardos, M. Arnal, J. Billy, B. Peaudecerf, D. Sugny, and D. Guéry-Odelin, Quantum state control of a bose-einstein condensate in an optical lattice, PRX Quantum 2, 040303 (2021).
  • Abah and Paternostro [2019b] O. Abah and M. Paternostro, Shortcut-to-adiabaticity Otto engine: A twist to finite-time thermodynamics, Phys. Rev. E 99, 022110 (2019b).
  • Daems et al. [2008] D. Daems, S. Guérin, and N. J. Cerf, Quantum search by parallel eigenvalue adiabatic passage, Phys. Rev. A 78, 042322 (2008).
  • Born and Fock [1928] M. Born and V. Fock, Beweis des Adiabatensatzes, Z. Phys. 51, 165 (1928).
  • Allahverdyan and Nieuwenhuizen [2005] A. E. Allahverdyan and Th. M. Nieuwenhuizen, Minimal work principle: Proof and counterexamples, Phys. Rev. E 71, 046107 (2005).
  • Albash et al. [2012] T. Albash, S. Boixo, D. A. Lidar, and P. Zanardi, Quantum adiabatic Markovian master equations, New J. Phys. 14, 123016 (2012).
  • Latune [2021] C. L. Latune, Energetic advantages of nonadiabatic drives combined with nonthermal quantum states, Phys. Rev. A 103, 062221 (2021).
  • Moutinho et al. [2023] J. P. Moutinho, M. Pezzutto, S. S. Pratapsi, F. F. da Silva, S. De Franceschi, S. Bose, A. T. Costa, and Y. Omar, Quantum Dynamics for Energetic Advantage in a Charge-Based Classical Full Adder, PRX Energy 2, 033002 (2023).
  • Góis et al. [2024] F. Góis, M. Pezzutto, and Y. Omar, Towards Energetic Quantum Advantage in Trapped-Ion Quantum Computation, arXiv 10.48550/arXiv.2404.11572 (2024), 2404.11572 .
  • Campbell and Deffner [2017] S. Campbell and S. Deffner, Trade-Off Between Speed and Cost in Shortcuts to Adiabaticity, Phys. Rev. Lett. 118, 100601 (2017).
  • Zheng et al. [2016] Y. Zheng, S. Campbell, G. De Chiara, and D. Poletti, Cost of counterdiabatic driving and work output, Phys. Rev. A 94, 042132 (2016).
  • Kiely et al. [2022] A. Kiely, S. Campbell, and G. T. Landi, Classical dissipative cost of quantum control, Phys. Rev. A 106, 012202 (2022).
  • Fellous-Asiani et al. [2023a] M. Fellous-Asiani, J. H. Chai, Y. Thonnart, H. K. Ng, R. S. Whitney, and A. Auffèves, Optimizing Resource Efficiencies for Scalable Full-Stack Quantum Computers, PRX Quantum 4, 040319 (2023a).
  • Auffèves [2022] A. Auffèves, Quantum Technologies Need a Quantum Energy Initiative, PRX Quantum 3, 020101 (2022).
  • Fellous-Asiani et al. [2023b] M. Fellous-Asiani, J. H. Chai, Y. Thonnart, H. K. Ng, R. S. Whitney, and A. Auffèves, Optimizing Resource Efficiencies for Scalable Full-Stack Quantum Computers, PRX Quantum 4, 040319 (2023b).
  • Zhu et al. [2022] J.-j. Zhu, X. Laforgue, X. Chen, and S. Guérin, Robust quantum control by smooth quasi-square pulses, J. Phys. B: At. Mol. Opt. Phys. 55, 194001 (2022).
  • Bernardo [2020] B. d. L. Bernardo, Time-rescaled quantum dynamics as a shortcut to adiabaticity, Phys. Rev. Res. 2, 013133 (2020).
  • Ferreira et al. [2024] J. L. M. Ferreira, Â. F. d. S. França, A. Rosas, and B. d. L. Bernardo, Shortcuts to adiabaticity designed via time-rescaling follow the same transitionless route, arXiv 10.48550/arXiv.2406.07433 (2024), 2406.07433 .
  • Bason et al. [2012] M. G. Bason, M. Viteau, N. Malossi, P. Huillery, E. Arimondo, D. Ciampini, R. Fazio, V. Giovannetti, R. Mannella, , and O. Morsch, High-fidelity quantum driving, Nature Physics 8, 147 (2012).
  • Hegerfeldt [2013] G. C. Hegerfeldt, Driving at the quantum speed limit: Optimal control of a two-level system, Phys. Rev. Lett. 111, 260501 (2013).
  • Zenesini et al. [2009] A. Zenesini, H. Lignier, G. Tayebirad, J. Radogostowicz, D. Ciampini, R. Mannella, S. Wimberger, O. Morsch, and E. Arimondo, Time-resolved measurement of landau-zener tunneling in periodic potentials, Phys. Rev. Lett. 103, 090403 (2009).
  • Tayebirad et al. [2010] G. Tayebirad, A. Zenesini, D. Ciampini, R. Mannella, O. Morsch, E. Arimondo, N. Lörch, and S. Wimberger, Time-resolved measurement of landau-zener tunneling in different bases, Phys. Rev. A 82, 013633 (2010).
  • Dridi et al. [2020] G. Dridi, K. Liu, and S. Guérin, Optimal Robust Quantum Control by Inverse Geometric Optimization, Phys. Rev. Lett. 125, 250403 (2020).
  • Harutyunyan et al. [2023] M. Harutyunyan, F. Holweck, D. Sugny, and S. Guérin, Digital optimal robust control, Phys. Rev. Lett. 131, 200801 (2023).
  • Van Damme et al. [2017] L. Van Damme, Q. Ansel, S. J. Glaser, and D. Sugny, Robust optimal control of two-level quantum systems, Phys. Rev. A 95, 063403 (2017).