Brought to you by:
Paper

How fast and robust is the quantum adiabatic passage?

Published 17 July 2013 © 2013 IOP Publishing Ltd
, , Citation Kazutaka Takahashi 2013 J. Phys. A: Math. Theor. 46 315304 DOI 10.1088/1751-8113/46/31/315304

1751-8121/46/31/315304

Abstract

We study the assisted adiabatic passage, and equivalently the transitionless quantum driving, as a quantum brachistochrone trajectory. The optimal Hamiltonian for given constraints is constructed from the quantum brachistochrone equation. We discuss how the adiabatic passage is realized as the solution of the equation. The formulation of the quantum brachistochrone is based on the principle of least action. We utilize it to discuss the stability of the adiabatic passage.

Export citation and abstract BibTeX RIS

1. Introduction

Coherent manipulation of quantum systems is of fundamental and practical importance and has been discussed intensively. Theoretically, the problem can be stated in a simple question: what is the optimal Hamiltonian under a given condition? For example, we seek the Hamiltonian minimizing the traveling time between two states, or maximizing the fidelity between the state under evolution and a reference state. Depending on experimental situations, we can consider optimizations in many possible ways. Several optimization methods have been proposed, each based on a different philosophy. The assisted adiabatic passage (AP) [1, 2] and the transitionless quantum driving [3] are known as a nonadiabatic driving following an adiabatic state. In this method, the counter-diabatic Hamiltonian is introduced to make the evolution transitionless. Several closely related methods have been proposed such as shortcuts to adiabaticity or the Lewis–Riesenfeld invariant-based engineering [4] and the fast-forward scaling [57]. These methods have been intensively studied in recent years [824].

It has been discussed in, e.g., [11, 14] that some of those methods are related to each other. This implies that there exists a common principle behind. It can be useful to reformulate the problem from a different general perspective. In this paper, we study the AP by using the method of the quantum brachistochrone (QB) [2527]. For a quantum trajectory, we define an action and the optimal Hamiltonian and state are determined from the variational principle. The advantage of this method is in its simplicity and generality. The QB equation is derived in a general form and the system-dependent conditions are implemented in the constraint part. By using the QB equation, we show that the solution of the equation is interpreted as the AP. It is not trivial what kind of adiabatic state appears as the solution and we examine the equation in several examples.

We also study the stability of the AP. In several examples, the AP was shown to be robust against variations of control parameters [13, 19]. This problem was closely studied in [18, 23, 24]. Since the QB is formulated by the variational principle, we can consider the stability of the solution in a rather general way. We derive a general stability condition and examine several examples explicitly.

The paper is organized as follows. In section 2, we formulate the QB method and discuss the relation with the method of the AP. Then, several examples are studied in section 3. The stability of the AP is discussed in section 4. Section 5 is devoted to a summary.

2. Quantum brachistochrone

2.1. Quantum brachistochrone equation

In order to derive the optimal solution of the time-dependent Hamiltonian H(t) and state |ψ(t)〉, we define the action

Equation (1)

as an integral between the initial and final time [25]. Each term is expressed as

Equation (2)

Equation (3)

Equation (4)

We use the notations P = |ψ〉〈ψ|, ΔE2 = 〈ψ|H2|ψ〉 − 〈ψ|H|ψ〉2 and a dotted symbol for the time derivative. LT dt represents the Fubini–Study distance 〈dψ|(1 − P)|dψ〉 divided by ΔE. The energy variance ΔE plays the role of the velocity in quantum systems [28], which means that ∫LT dt represents the time to be taken for the evolution of the quantum state. The other two terms are introduced to impose constraints which must hold throughout the evolution. The Schrödinger equation ${\rm i}|\dot{\psi }\rangle =H|\psi \rangle$ is implemented by LS and other constraints fa(H) = 0 by LC. |ϕ〉 and λa represent the Lagrange multipliers. They are determined in course of the calculation.

For the fixed initial state |ψ(0)〉 and the final one |ψ(T)〉, the action is extremized by variations HH + δH and |ψ〉 → |ψ〉 + |δψ〉. Setting the linear term to be zero, we obtain

Equation (5)

Equation (6)

where

Equation (7)

The prime denotes the derivative with respect to H. After some calculations, we obtain

Equation (8)

where U(t) is the time-evolution operator for the Hamiltonian H(t) and

Equation (9)

Thus, F satisfies the QB equation [25]

Equation (10)

with the initial condition

Equation (11)

We note that condition (11) is unchanged throughout the evolution. These equations are solved under fixed T. Then, we minimize the passage time T to accomplish our purpose to obtain the optimal solution.

For convenience of the calculation, we rewrite the QB equation by using the basis operators $\lbrace X_\mu \rbrace _{\mu =0,1,\ldots ,N^2-1}$ where N is the dimension of the Hilbert space. These satisfy the orthonormal relation

Equation (12)

and the completeness relation

Equation (13)

The structure constant fμνρ is defined by the commutation relation

Equation (14)

We set X0 to be the identity operator and use the vector representation X for other components $\lbrace X_a\rbrace _{a=1,\ldots ,N^2-1}$. Then, the Hamiltonian is represented as

Equation (15)

We note that the first term does not bring any quantum effect and can be eliminated by a proper gauge transformation |ψ(t)〉 → eiφ(t)|ψ(t)〉.

Using this representation, we can parametrize P and F as

Equation (16)

Equation (17)

where e and l are normalized vectors with the angle between them fixed: l(t) · e(t) = l(0) · e(0). These representations are derived from Tr P = Tr P2 = 1 and Tr F = Tr F P. In the following analysis, the overall constant λ is not important and is left undetermined. We also note that the first term in (17) is derived from the condition that the first term in (15) is fixed and is not important as well. From equation (10), we can write the equations of motion as

Equation (18)

where the vector product is defined by $(\boldsymbol {h}\times \boldsymbol {l})_a=\sum _{b,c}f_{abc}h_bl_c$. Since the state |ψ(t)〉 satisfies the Schrödinger equation, P follows the same equation as (10) and, as a result, $\dot{\boldsymbol {e}}(t)=\boldsymbol {h}(t)\times \boldsymbol {e}(t)$.

The solution of the QB equation strongly depends on the constraints imposed. We consider the case where components a = 1, 2, ..., k are fixed as

Equation (19)

Correspondingly, fa and F are given by

Equation (20)

Equation (21)

The QB equation is solved to give other components:

Equation (22)

The total Hamiltonian is given by H(t) = HC(t) + HQB(t). For a given $h_a^{(0)}(t)$, time-dependences of la(t) and $h^{(1)}_p(t)$ are obtained by solving equation (18) and the initial condition is determined by equation (11).

2.2. Lewis–Riesenfeld invariant and transitionless quantum driving

The QB equation is written in terms of the operator $F=\sum _a\lambda _af_a^{\prime }(H)$ coming from the constraints fa(H) = 0. It is crucial in the following analysis to notice that this quantity F is nothing but the Lewis–Riesenfeld invariant [29]. It satisfies equation (10) and, as a result, all eigenvalues of F become independent of time. We can write

Equation (23)

where the time-independent eigenvalue is denoted by λn and |n(t)〉 is the eigenstate defined at each time. By using the eigenstates, we can write the state as

Equation (24)

with a phase factor

Equation (25)

The real constant cn is shown to be independent of time, which means that if we start the evolution from |n(0)〉, the state remains the eigenstate |n(t)〉 at arbitrary t.

In order to represent the optimal Hamiltonian in the basis of {|n(t)〉}, we multiply 〈m(t)| from the left and |n(t)〉 from the right to equation (10). We assume that there is no degeneracy in the spectrum for simplicity. Then, the off-diagonal element 〈m(t)|H(t)|n(t)〉 with mn is calculated as

Equation (26)

The Hamiltonian is thus written as H(t) = H0(t) + H1(t) where

Equation (27)

Equation (28)

The diagonal part H0(t) commutes with the invariant F(t). Its eigenvalue En(t) cannot be specified from equation (10). We see that this Hamiltonian represents the formula of the transitionless quantum driving [3]. In that method, the adiabatic state (24), with H(t) replaced by H0(t), is constructed for a given Hamiltonian H0(t). This is not the solution of the Schrödinger equation and we apply the counter-diabatic Hamiltonian H1(t) to make the system transitionless. Then, the adiabatic state becomes the exact solution of the equation.

The result here shows that the assisted AP, transitionless quantum driving and shortcuts to adiabaticity are essentially equivalent and they can be derived from the QB equation.

2.3. Adiabatic passage as a quantum brachistochrone trajectory

Although we have shown that the solution of the QB equation can be interpreted as the AP, it is not clear from the beginning what kind of adiabatic state is obtained. It will be practically useful if the adiabatic state is determined by the constrained part HC(t) so that we can control the system at will. That is, we want to know the cases where the relations H0(t) = HC and H1(t) = HQB hold. In the following, we set the initial state |ψ(0)〉 to be an eigenstate of F(0). In this case, we have |ϕ(t)〉∝|ψ(t)〉 and F(t)∝P(t).

The condition that HC(t) determines the adiabatic state is given by the commutation relation

Equation (29)

This can be written as

Equation (30)

We easily see that a possible solution is given by l(t) ∝ h(0)(t). It is hard to imagine that the equation has other nontrivial solutions since the commutation relation (29) must hold for arbitrary t. Although it is an interesting problem to find such a solution in a specific example, we discuss the solution l(t) ∝ h(0)(t) in the following general analysis. Thus, F is equivalent to HC(t) as

Equation (31)

where $h^{(0)}(t)=|\boldsymbol {h}^{(0)}(t)|=\sqrt{\sum _ah_a^{(0)}(t)h_a^{(0)}(t)}$. We neglected the unimportant constant term and overall factor. The QB equation is given by

Equation (32)

We note that this equation does not necessarily have the solution. If no solution exists, H0(t) ≠ HC(t) is implied. When the solution exists, each part of the Hamiltonian is given by

Equation (33)

Equation (34)

where

Equation (35)

Generally, there exists the Hamiltonian δH1(t) in HQB(t) such that the adiabatic state is not disturbed. Since the QB equation is the most general equation, the result depends strongly on the constraints imposed. If the constraints are too loose, the solution has many ambiguities and is not determined uniquely. On the other hand, tight constraints will give trivial and ineffective solutions. It is important to choose proper constraints to obtain nontrivial results. Since it is hard to study general conditions, we examine the QB equation explicitly in the next section.

3. Examples

3.1. N = 2

First, we consider the simplest example of N = 2. The Hamiltonian is written by the Pauli matrices σ as

Equation (36)

The basis operator is given by X = (1, σ1, σ2, σ3) and the structure constant by fabc = 2epsilonabc. Equation (32) is rewritten as

Equation (37)

where the definition of the vector product is changed as $(\boldsymbol {h}^{(1)}\times \boldsymbol {h}^{(0)})_a=\sum _{bc}\epsilon _{abc}h_b^{(1)}h_c^{(0)}$. Considering the vector product with h(0)(t), we obtain

Equation (38)

Here, we used the property of the antisymmetric tensor ∑aepsilonabcepsilonade = δbdδce − δbeδcd. This result gives the counter-diabatic part H1(t) derived in the method of the transitionless quantum driving [3]. That is, we conclude that HC(t) = H0(t) and HQB(t) = H1(t) in the present case.

We note that the nontrivial result is obtained only in the case where two components of h(0) are constrained. In that case, the third component is determined by equation (38). When one component of h(0) is constrained, no quantum effect appears and we do not have any interesting results. For three components fixed, the Hamiltonian is completely specified and there is no room for finding the optimal Hamiltonian.

3.2. N = 3

Next, we consider the case of N = 3. In this case, we use the Gell–Mann matrices

Equation (39)

as the basis operators. Normalizing the operators, we have $X_{a}=\frac{\sqrt{6}}{2}\lambda _a$ and the structure constant is given in table 1. Since we do not have useful relations on the structure constant, it is hard to consider the general properties as we did in the case of N = 2. In the following, we study the properties of the solution explicitly.

Table 1. The structure constant fabc for the basis operators $X_a=\frac{\sqrt{6}}{2}\lambda _a$ (a = 1, ...8) at N = 3. λa represent the Gell–Mann matrices in equation (39). fabc is antisymmetric with respect to all pairs of indices. The other components not shown in the table are zero.

a b c fabc a b c fabc
1 2 3 $\sqrt{6}$ 3 4 5 $\frac{1}{2}\sqrt{6}$
1 4 7 $\frac{1}{2}\sqrt{6}$ 3 6 7 $\frac{1}{2}\sqrt{6}$
1 5 6 $\frac{1}{2}\sqrt{6}$ 4 5 8 $\frac{\sqrt{3}}{2}\sqrt{6}$
2 4 6 $\frac{1}{2}\sqrt{6}$ 6 7 8 $\frac{\sqrt{3}}{2}\sqrt{6}$
2 5 7 $\frac{1}{2}\sqrt{6}$        

3.2.1. k = 2

We first consider the case where two components of h(t), a and b, are fixed. We assume that there exist components p such that fabp ≠ 0. Otherwise, no quantum effect appears. The explicit analysis depends on how many p exist. Although it is always possible to choose basis operators such that the unique p exists, we consider the case when the Gell–Mann matrices are used as the basis operators. In that case, we have the following two patterns.

  • a = 1, b = 2.In this case, p = 3 is the only component with fabp ≠ 0. From equation (32), we obtain
    Equation (40)
    We also see that $h_8^{(1)}(t)$ is left undetermined from the QB equation. In fact, X8 commutes with $H_{\rm C}(t)=h_1^{(0)}X_1+h_2^{(0)}X_2$ and does not disturb the adiabatic state. On the other hand, by using the formula of the transitionless quantum driving, equations (27) and (28), we obtain the counter-diabatic Hamiltonian $H_1(t)=h_3^{(1)}(t)X_3$. We conclude that the AP is realized as
    Equation (41)
    Equation (42)
  • a = 4, b = 5.There exist two components f345 and f458 with nonzero value. Then, we obtain
    Equation (43)
    The solution is given by
    Equation (44)
    Equation (45)
    where h(t) is an arbitrary function. $h(t)(\sqrt{3}X_3-X_8)$ commutes with the constraint part and does not affect the adiabatic part. Each first term in the above equations is derived from equations (27) and (28). Therefore, we conclude also in this case as
    Equation (46)
    Equation (47)

3.2.2. k = 3

We next consider the case where three components a, b and c are fixed. In this case, possible numbers of components p such that fabp ≠ 0 are four at the maximum. We can consider each pattern to find the formula of the transitionless driving. In the present case, another problem arises such that the structure constant among the constraint components is nonzero: fabc ≠ 0.

  • a = 1, b = 2, c = 4.In this case, components 3, 6, 7 and 8 participate in equations to determine h(1). The equations to be solved are given by
    Equation (48)
    Equation (49)
    Equation (50)
    Equation (51)
    The third equation is derived from the first and second equations. Since three independent equations are imposed on four variables $h_3^{(1)}$, $h_6^{(1)}$, $h_7^{(1)}$ and $h_8^{(1)}$, we have an arbitrariness in choosing the solution as in the previous example.On the other hand, if we apply the formula of the transitionless quantum driving by setting $H_0(t)=H_{\rm C}(t)=h_1^{(0)}(t)X_1+h_2^{(0)}(t)X_2+h_4^{(0)}(t)X_4$, we obtain the counter-diabatic Hamiltonian
    Equation (52)
    Equation (53)
    Equation (54)
    Equation (55)
    Equation (56)
    This result is a solution of the QB equation. The arbitrariness comes from the choice of the diagonal part of H1, δH1(t) in equation (34).
  • a = 3, b = 4, c = 5.This case is different from the previous one due to the property f345 ≠ 0. Then, the ansatz l(t) ∝ h(0)(t) does not solve the QB equation. Equation (18) is written explicitly as
    Equation (57)
    Equation (58)
    Equation (59)
    Equation (60)
    Equation (61)
    Equation (62)
    Equation (63)
    Equation (64)
    From the first and last equations, we obtain
    Equation (65)
    where θ is a constant. Using the second or third equations, we can calculate $h_8^{(1)}$ as
    Equation (66)
    The other components $h_1^{(1)}$, $h_2^{(1)}$, $h_6^{(1)}$ and $h_7^{(1)}$ are shown to be zero. Thus, we obtain $H_{\rm QB}(t) = h_8^{(1)}(t)X_8$.Correspondingly, for $H_0(t)=h^{(0)}_3X_3+h^{(0)}_4X_4+h^{(0)}_5X_5$, the AP is given by the counter-diabatic Hamiltonian of the form $H_1(t)=h_3^{(1)}(t)X_3+h_4^{(1)}(t)X_4+h_5^{(1)}(t)X_5+h_8^{(1)}(t)X_8$. Due to the property f345 ≠ 0, H1 inevitably has the same components as H0. Therefore, in this case, H0(t) ≠ HC(t).

In the same way, we can show that the formula of the transitionless driving is derived when fabc = 0. We have examined four possible patterns depending on the number of components p such that fabp ≠ 0. In some cases, the QB equation cannot determine the solution uniquely and there exists some ambiguity. It does not disturb the adiabatic state and we can conclude that the constraint part gives the adiabatic state. This is generalized to the case where higher numbers of components are constrained.

4. Stability of the adiabatic passage

4.1. General consideration

The advantage of describing the AP as the QB is that one can study the stability of the driving. It is shown in several examples that the AP is robust against parameter variations [13, 19]. Here, we show that it is generally correct. We also find that the instability arises when we consider perturbation by different kinds of operators.

The QB equation is derived by expanding the action up to first order in δH and |δψ〉. The stability of the extremized solution is found by examining the second order. As a possible situation, we consider the case when the Hamiltonian is changed as HH + δH. H and |ψ〉 are determined from the QB equation and we see what happens if the counter-diabatic Hamiltonian deviates from the ideal form.

The second order of the action in δH is calculated as

Equation (67)

where 〈(⋅⋅⋅)〉 = 〈ψ(t)|(⋅⋅⋅)|ψ(t)〉. The solution of the QB equation is stable when I(t) > 0. This is the general result applied to any QB trajectories. We are interested in the case where H(t) = H0(t) + H1(t) with H0(t) is the Hamiltonian giving the adiabatic state and H1(t) is the counter-diabatic Hamiltonian. In this case, the state |ψ〉 is the eigenstate of H0(t) and the diagonal elements of H1(t) are zero in this representation. Then, we can write

Equation (68)

This is the main result in this section. The stability of the AP is given by the condition I(t) > 0.

When the variation is proportional to the counter-diabatic Hamiltonian as δH(t) = c(t)H1(t) with an arbitrary scalar function c(t), we can show that I(t) = c2(t), which means that the AP is stable against any parameter variations in the counter-diabatic Hamiltonian. This result is consistent with previous examples showing stable driving [13, 19]. It is also possible to consider the unstable perturbation in principle. Such a perturbation is realized when the second term of equation (68) becomes smaller compared with the first term. It is accomplished when H1(t) and δH(t) anticommute with each other: H1(tH(t) + δH(t)H1(t) = 0. We can generally say that the AP is unstable against perturbations by operators anticommuting with the counter-diabatic Hamiltonian.

4.2. Example: N = 3

Equation I(t) > 0 represents a necessary condition of the stability. We examine this condition by using an example to see whether the expected behavior is obtained. Although the simplest example of the quantum system is the two-level system N = 2, unstable perturbations cannot be considered in this case. If we fix two components of the magnetic field, the third component is determined by the QB equation and no other components exist. For this reason, we treat the three-level system N = 3 in this subsection.

As an adiabatic Hamiltonian, we consider the magnetic field in the xy plane

Equation (69)

Equation (70)

Equation (71)

where S1 and S2 are spin operators,

Equation (72)

Equation (73)

and can be expressed by linear combinations of the Gell–Mann matrices. The adiabatic state is given by

Equation (74)

By using the formulas (27) and (28) of the transitionless quantum driving, we obtain the counter-diabatic Hamiltonian

Equation (75)

From the general consideration, the driving should be stable against the variation ω → ω + δω(t) where δω(t) is an arbitrary function. It is also possible to consider perturbations inducing instabilities. For example, as operators which anticommute with S3, we can use the Gell–Mann matrices λ4 and λ5.

We numerically solve the Schrödinger equation to calculate the fidelity

Equation (76)

where |ψad(t)〉 denotes the adiabatic state of H0(t) in equation (74) and |ψ(t)〉 the solution of the Schrödinger equation with the Hamiltonian H0(t) + H1(t) + δH(t). We set the initial condition at t = 0 to be the eigenstate of S1 with the eigenvalue +1 and calculate the probability of observing the eigenstate of S2 with the eigenvalue +1.

We set parameters h0 = 1 and ω = π/20. As possible perturbations, we consider the following four cases:

Equation (77)

We set each coefficient so that the instability I(t), calculated respectively as

Equation (78)

gives the same magnitude in average. We set δh(t) = 0.5.

The results are plotted in figures 14. We see in the stable case that the trajectory follows that of the ideal driving case while oscillating uniformly. In unstable cases, we see more nonuniform behavior and large deviations for λ4 and λ5. The deviation is relatively small for λ8 but the oscillations are nonuniform compared with the stable case of S3. We note that λ8 commutes with the driving term S3 and does not disturb the system significantly. Such a difference cannot be seen in the quantity I(t). We consider only the local instability of the solution of the QB equation. Although it is interesting to see nonlinear effects, such an analysis is beyond the scope of this study. The result here shows that the condition I(t) > 0 can be one of the criteria for the stable driving.

Figure 1.

Figure 1. Transitionless quantum driving with S3-perturbation. The red dotted line represents the fidelity and the blue solid line is the probability of observing the eigenstate of S2 with the eigenvalue +1. The black dashed line is the probability for the case of the ideal driving without perturbation.

Standard image High-resolution image
Figure 2.

Figure 2. Driving with λ4-perturbation.

Standard image High-resolution image
Figure 3.

Figure 3. Driving with λ5-perturbation.

Standard image High-resolution image
Figure 4.

Figure 4. Driving with λ8-perturbation.

Standard image High-resolution image

5. Summary

We have studied the AP as a QB trajectory. The QB equation shows that the Lewis–Riesenfeld invariant exists for the optimized trajectory and the solution is interpreted as an AP. As a practical situation, we closely discussed the case where the constraint part of the Hamiltonian gives the adiabatic state so that one can control the system in a favorable way. By considering several cases explicitly, we find that such a case is realized when the commutator of the basis operators between constraint parts [Xa, Xb] does not belong to the constraint part. This result will be a guiding principle in considering the ideal manipulation.

We note that we only considered the case where the constraint is given by the form (19). We can consider other types of constraints in the QB equation such as fixing |h(t)|. It is not clear what kind of the AP is realized in that case. Such an analysis will be an interesting future problem. In order to discuss the property of the AP more generally, it may be useful to consider a Lie-algebraic classification of the Lewis–Riesenfeld invariants as discussed in [30].

We also studied the stability of the solution and derived a general necessary condition for the stability. The result is confirmed in a three-level system. Generally, the AP is stable against the variation of the operators which commutes with the counter-diabatic Hamiltonian H1(t) and unstable with operators anticommutes with H1(t).

In this paper, we examined only the variation with respect to the Hamiltonian. It is possible to consider the variation of the state. Then, it can be possible to consider the stability under a change of the initial state for example.

Another possible study to be done is to interpret the fast-forward scaling [57] by using the QB. In the fast-forward scaling, we utilize a reference state in fast-forwarding the evolution. In such a formulation, the fidelity-optimized QB is expected to be suitable [27]. By clarifying the relations between various methods, we hope that we can understand the nature of quantum fluctuations in a deeper way, which will be useful for optimal coherent manipulations.

Acknowledgments

The author is grateful to S Masuda, Y Shikano and J Tsuda for useful discussions and comments.

Please wait… references are loading.
10.1088/1751-8113/46/31/315304