Next Article in Journal
A Regularization Homotopy Strategy for the Constrained Parameter Inversion of Partial Differential Equations
Next Article in Special Issue
Thermality versus Objectivity: Can They Peacefully Coexist?
Previous Article in Journal
The Composition of Saturated Vapor over 1-Butyl-3-methylimidazolium Tetrafluoroborate Ionic Liquid: A Multi-Technique Study of the Vaporization Process
Previous Article in Special Issue
Many-Body Localization and the Emergence of Quantum Darwinism
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Environment-Assisted Shortcuts to Adiabaticity

1
Department of Physics, University of Maryland, Baltimore County, Baltimore, MD 21250, USA
2
Instituto de Física ‘Gleb Wataghin’, Universidade Estadual de Campinas, Campinas, São Paulo 13083-859, Brazil
*
Author to whom correspondence should be addressed.
Entropy 2021, 23(11), 1479; https://doi.org/10.3390/e23111479
Submission received: 16 October 2021 / Revised: 4 November 2021 / Accepted: 5 November 2021 / Published: 9 November 2021
(This article belongs to the Special Issue Quantum Darwinism and Friends)

Abstract

:
Envariance is a symmetry exhibited by correlated quantum systems. Inspired by this “quantum fact of life,” we propose a novel method for shortcuts to adiabaticity, which enables the system to evolve through the adiabatic manifold at all times, solely by controlling the environment. As the main results, we construct the unique form of the driving on the environment that enables such dynamics, for a family of composite states of arbitrary dimension. We compare the cost of this environment-assisted technique with that of counterdiabatic driving, and we illustrate our results for a two-qubit model.

1. Introduction

An essential step in the development of viable quantum technologies is to achieve precise control over quantum dynamics [1,2]. In many situations, optimal performance relies on the ability to create particular target states. However, in dynamically reaching such states, the quantum adiabatic theorem [3] poses a formidable challenge since finite-time driving inevitably causes parasitic excitations [4,5,6,7]. Acknowledging and addressing this issue, the field of “shortcuts to adiabaticity” (STA) [8,9,10,11] has developed a variety of techniques that permit to facilitate effectively adiabatic dynamics in finite time.
Recent years have seen an explosion of work on, for instance, counterdiabatic driving [12,13,14,15,16,17,18,19], the fast-forward method [20,21,22,23], time-rescaling [24,25], methods based on identifying the adiabatic invariant [26,27,28,29], and even generalizations to classical dynamics [30,31,32]. For comprehensive reviews of the various techniques, we refer to the recent literature [9,10,11].
Among these different paradigms, counterdiabatic driving (CD) stands out, as it is the only method that forces evolution through the adiabatic manifold at all times. However, experimentally realizing the CD method requires applying a complicated control field, which often involves non-local terms that are hard to implement in many-body systems [15,17]. This may be particularly challenging if the system is not readily accessible, due to, for instance, geometric restrictions of the experimental set-up.
In the present paper, we propose an alternative method to achieve transitionless quantum driving by leveraging the system’s (realistically) inevitable interaction with the environment. Our novel paradigm is inspired by “envariance,” which is short for entanglement-assisted invariance. Envariance is a symmetry of composite quantum systems, first described by Wojciech H. Zurek [33]. Consider a quantum state | ψ SE that lives on a composite quantum universe comprising the system, S , and its environment, E . Then, | ψ SE is called envariant under a unitary map u S I E if and only if there exists another unitary I S u E acting on E such that the composite state remains unaltered after applying both maps, i.e., u S I E | ψ SE = | ϕ SE and I S u E | ϕ SE = | ψ SE . In other words, the state is envariant if the action of a unitary on S can be inverted by applying a unitary on E .
Envariance was essential to derive Born’s rule [33,34], and in formulating a novel approach to the foundations of statistical mechanics [35]. Moreover, experiments [36,37] showed that this inherent symmetry of composite quantum states is indeed a physical reality, or rather a “quantum fact of life” with no classical analog [34]. Drawing inspiration from envariance, we develop a novel method for transitionless quantum driving. In the following, we will see that instead of inverting the action of a unitary on S , we can suppress undesirable transitions in the energy eigenbasis of S by applying a control field on the environment E . In particular, we consider the unitary evolution of an ensemble of composite states { | ψ SE } on a Hilbert space H S H E of arbitrary dimension, and we determine the general analytic form of the time-dependent driving on H E , which suppresses undesirable transitions in the system of interest S . This general driving on the environment E guarantees that the system S evolves through the adiabatic manifold at all times. We dub this technique environment-assisted shortcuts to adiabaticity, or “EASTA” for short. In addition, we prove that the cost associated with the EASTA technique is exactly equal to that of counterdiabatic driving. We illustrate our results in a simple two-qubit model, where the system and the environment are each described by a single qubit. Finally, we conclude with discussing a few implications of our results in the general context of decoherence theory and quantum Darwinism.

2. Counterdiabatic Driving

We start by briefly reviewing counterdiabatic driving to establish notions and notations. Consider a quantum system S , in a Hilbert space H S of dimension d S , driven by the Hamiltonian H 0 ( t ) with instantaneous eigenvalues { E n ( t ) } n 0 , d S 1 and eigenstates { | n ( t ) } n 0 , d S 1 . For slowly varying H 0 ( t ) , according to the quantum adiabatic theorem [3], the driving of S is transitionless. In other words, if the system starts in the eigenstate | n ( 0 ) , at t = 0 , it evolves into the eigenstate | n ( t ) at time t (with a phase factor) as follows:
| ψ n ( t ) U ( t ) | n ( 0 ) = e i 0 t E n ( s ) d s 0 t n s n d s | n ( t ) e i f n ( t ) | n ( t ) .
For arbitrary driving H 0 ( t ) , namely for driving rates larger than the typical energy gaps, the system undergoes transitions. However, it was shown [12,13,14] that the addition of a counterdiabatic field H CD ( t ) forces the system to evolve through the adiabatic manifold. Using the following total Hamiltonian,
H = H 0 ( t ) + H CD ( t ) = H 0 ( t ) + i n t n n n t n n n | ,
the system evolves with the corresponding unitary U CD ( t ) = n | ψ n ( t ) n ( 0 ) | such that the following holds:
U CD ( t ) | n ( 0 ) = e i f n ( t ) | n ( t ) .
This evolution is exact no matter how fast the system is driven by the total Hamiltonian. However, the counterdiabatic driving (CD) method requires adding a complicated counterdiabatic field H CD ( t ) involving highly non-local terms that are hard to implement in a many-body set-up [15,17]. Constructing this counterdiabatic field requires determining the instantaneous eigenstates { | n ( t ) } n 0 , d S 1 of the time-dependent Hamiltonian H 0 ( t ) . Moreover, changing the dynamics of the system of interest (i.e., adding the counterdiabatic field) requires direct access and control on S .
In the following, we will see how (at least) the second issue can be circumvented by relying on the environment E that inevitably couples to the system of interest. In particular, we make use of the entanglement between system and environment to avoid any transitions in the system. To this end, we construct the unique driving of the environment E that counteracts the transitions in S .

3. Open System Dynamics and STA for Mixed States

We start by stating three crucial assumptions: (i) the joint state of the system S and the environment E is described by an initial wave function | ψ SE ( 0 ) evolving unitarily, according to the Schrödinger equation; (ii) the environment’s degrees of freedom do not interact with each other; (iii) the S - E joint state belongs to the ensemble of singly branching states [38]. These branching states have the following general form:
| ψ SE = n = 0 N 1 p n | n l = 1 N E | E n l ,
where p n [ 0 , 1 ] is the probability associated with the nth branch of the wave function, with orthonormal states | n H S and l = 1 N E | E n l H E .
Without loss of generality, we can further assume p n = 1 / N for all n 0 , N 1 since if p n 1 / N we can always find an extended Hilbert space [33,34] such that the state | ψ SE becomes even. Thus, we can consider branching states | ψ SE of the simpler form as follows:
| ψ SE = 1 N n = 0 N 1 | n l = 1 N E | E n l .
In the following, we will see that EASTA can actually only be facilitated for even states (5). In Appendix B, we show that EASTA cannot be implemented for arbitrary probabilities (i.e., ( n ) ; p n 1 / N ).

3.1. Two-Level Environment E

We start with the instructive case of a two-level environment, cf. Figure 1. To this end, consider the following branching state:
| ψ SE ( 0 ) = 1 2 | g ( 0 ) | E g ( 0 ) + 1 2 | e ( 0 ) | E e ( 0 ) ,
where the states | E g ( 0 ) and | E e ( 0 ) form a basis on the environment E , and the states | g ( 0 ) and | e ( 0 ) represent the ground and excited states of S at t = 0 , respectively.
It is then easy to see that there exists a unique unitary U such that the system evolves through the adiabatic manifold in each branch of the wave function as follows:
( ! U ) ; ( U U ) | ψ SE ( 0 ) = ( U CD I E ) | ψ SE ( 0 ) .
Starting from the above equality, we obtain the following:
U | g ( 0 ) U | E g ( 0 ) + U | e ( 0 ) U | E e ( 0 ) = e i f g ( t ) | g ( t ) | E g ( 0 ) + e i f e ( t ) | e ( t ) | E e ( 0 ) .
Projecting the environment E into the state “ | E g ( 0 ) ”, we have
U | g ( 0 ) E g ( 0 ) | U | E g ( 0 ) + U | e ( 0 ) E g ( 0 ) | U | E e ( 0 ) = e i f g ( t ) | g ( t ) ,
equivalently written as
( U g , g ) U | g ( 0 ) + ( U g , e ) U | e ( 0 ) = e i f g ( t ) | g ( t ) ,
which implies the following:
( U g , g ) | g ( 0 ) + ( U g , e ) | e ( 0 ) = e i f g ( t ) U | g ( t ) .
Therefore,
U g , g = e i f g ( t ) g ( 0 ) | U | g ( t ) , and U g , e = e i f g ( t ) e ( 0 ) | U | g ( t ) .
Additionally, by projecting E into the state “ | E e ( 0 ) ” we obtain the following:
U e , g = e i f e ( t ) g ( 0 ) | U | e ( t ) , and U e , e = e i f e ( t ) e ( 0 ) | U | e ( t ) .
It is straightforward to check that the operator U , which reads as follows:
U = U g , g U g , e U e , g U e , e ,
is indeed a unitary on E .
In conclusion, we have constructed a unique unitary map that acts only on E , but counteracts transitions in S . Note that coupling the system and environment implies that the state of the system is no longer described by a wave function. Hence the usual counterdiabatic scheme evolves the density matrix ρ S ( 0 ) to another density ρ S ( t ) such that both matrices have the same populations and coherence in the instantaneous eigenbasis of H 0 ( t ) (which is what EASTA accomplishes, as well).

3.2. N-Level Environment E

We can easily generalize the two-level analysis to an N-level environment. Similar to the above description, coupling the system to the environment leads to a branching state of the following form:
| ψ SE ( 0 ) = 1 N n = 0 N 1 | n ( 0 ) | E n ( 0 ) ,
where the states { | E n ( 0 ) } n form a basis on the environment E . We can then construct a unique unitary U such that the system evolves through the adiabatic manifold in each branch of the wave function as follows:
( ! U ) ; ( U U ) | ψ SE ( 0 ) = ( U CD I E ) | ψ SE ( 0 ) .
The proof follows the exact same strategy as the two-level case, and we find the following:
( ( m , n ) 0 , N 1 2 ) ; U m , n = e i f m ( t ) n ( 0 ) | U | m ( t ) .
The above expression of the elements of the unitary U is our main result, which holds for any driving H 0 ( t ) and any N-dimensional system.

3.3. Process Cost

Having established the general analytic form of the unitary applied on the environment, the next logical step is to compute and compare the cost of both schemes: (a) the usual counterdiabatic scheme and (b) the environment-assisted shortcut scheme presented above (cf. Figure 1). More specifically, we now compare the time integral of the instantaneous cost [39] for both driving schemes [39,40,41,42,43], (a) C CD ( t ) = ( 1 / τ ) 0 t H CD ( s ) d s and (b) C env ( t ) = ( 1 / τ ) 0 t H env ( s ) d s ( . is the operator norm), where the driving Hamiltonian on the environment can be determined from the expression of U ( t ) , H env ( t ) = i d U ( t ) d t U ( t ) .
In fact, from Equation (17) it is not too hard to see that the field applied on the environment H env ( t ) has the same eigenvalues as the counterdiabatic field H CD ( t ) , since there exists a similarity transformation between H env ( t ) and H CD * ( t ) . Therefore, the cost of both processes is exactly the same, C CD = C env , for any arbitrary driving H 0 ( t ) . Details of the derivation can be found in Appendix A. Note that for t = τ , the above definition of the cost becomes the total cost for the duration “ τ ” of the process.

3.4. Illustration

We illustrate our results in a simple two-qubit model, where the system and and the environment are each described by a single qubit. Note that the environment can live in a larger Hilbert space while still being characterized as a virtual qubit [44]. The aforementioned virtual qubit notion simply means that the state of the environment is of rank equal to two.
We choose a driving Hamiltonian H 0 ( t ) , such that
H 0 ( t ) = B 2 σ x + J ( t ) 2 σ z ,
where J ( t ) is the driving/control field, B is a constant, and σ z and σ x are Pauli matrices. Depending on the physical context, B and J ( t ) can be interpreted in various ways. In particular, as noted in ref. [45], in some contexts, the constant B can be regarded as the energy splitting between the two levels [46,47,48], and in others, the driving J ( t ) can be interpreted as a time-varying energy splitting between the states [49,50,51,52]. To illustrate our results we choose the following:
( t [ 0 , τ ] ) ; J ( t ) = B cos 2 π t 2 τ .
The above driving evolves the system beyond the adiabatic manifold, and we quantify this by plotting, in Figure 2, the overlap between the evolved state | ϕ n ( t ) U ( t ) | n ( 0 ) and the instantaneous eigenstate | n ( t ) of the Hamiltonian H 0 ( t ) , for n { g , e } . To illustrate our main result, we also plot the overlap between the states resulting from the two shortcut schemes (illustrated in Figure 1): the first scheme is the usual counterdiabatic (CD) driving, where we add a counterdiabatic field H CD to the system of interest, and we note the resulting composite state as “ | ψ SE CD ”. The second scheme is the environment-assisted shortcut to adiabaticity (EASTA), and we note the resulting composite state as “ | ψ SE EASTA ”. Confirming our analytic results, the local driving on the environment ensures that the system evolves through the adiabatic manifold at all times since the state overlap is equal to one for all t [ 0 , τ ] .
Finally, we compute and plot the cost of both shortcut schemes and verify that they are both equal to each other for all times “t” (cf. Figure 2b), and for all “ τ ” (cf. Figure 2c).

4. Concluding Remarks

4.1. Summary

In the present manuscript, we considered branching states { | ψ SE } , on a Hilbert space H S H E of arbitrary dimension, and we derived the general analytic form of the time-dependent driving on H E , which guarantees that the system S evolves through the adiabatic manifold at all times. Through this environment-assisted shortcuts to adiabaticity scheme, we explicitly showed that the environment can act as a proxy to control the dynamics of the system of interest. Moreover, for branching states | ψ SE with equal branch probabilities, we further proved that the cost associated with the EASTA technique is exactly equal to that of counterdiabatic driving. We illustrated our results in a simple two-qubit model, where the system and the environment are each described by a single qubit.
It is interesting to note that while we focused in the present manuscript on counterdiabatic driving, the technique can readily be generalized to any type of control unitary map “ U control ”, resulting in a desired evolved state | κ n ( t ) U control | n ( 0 ) . The corresponding unitary U on H E has then the following form:
( ( m , n ) 0 , N 1 2 ) ; U m , n = n ( 0 ) | U | κ m ( t ) .
In the special case, for which the evolved state is equal to the nth instantaneous eigenstate of H 0 ( t ) (with a phase factor),
| κ n ( t ) = e i f n ( t ) | n ( t ) ,
we recover the main result of the manuscript. The above generalization illustrates the broad scope of our results. Any control unitary on the system S can be realized solely by acting on the environment E , without altering the dynamics of the system of interest S (i.e., for any arbitrary driving H 0 ( t ) and thus, any driving rate).

4.2. Envariance and Pointer States

In the present work, we leveraged the presence of an environment to induce the desired dynamics in a quantum system. Interestingly, our novel method for shortcuts to adiabaticity relies on branching states, which play an essential role in decoherence theory and in the framework of quantum Darwinism.
In open system dynamics [53,54,55], the interaction between system and environment superselects states that survive the decoherence process, also known as the pointer states [56,57]. It is exactly these pointer states that are the starting point of our analysis, and for which EASTA is designed. While previous studies [58,59,60] have explored STA methods for open quantum systems, to the best of our understanding, the environment was only considered a passive source of additional noise described by quantum master equations. In our paradigm, we recognize the active role that an environment plays in quantum dynamics, which is inspired by envariance and reminiscent of the mindset of quantum Darwinism. In this framework [44,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77], the environment is understood as a communication channel through which we learn about the world around us, i.e., we learn about the state of systems of interest by eavesdropping on environmental degrees of freedom [44].
Thus, in true spirit of the teachings by Wojciech H. Zurek, we have understood the agency of quantum environments and the useful role they can assume. To this end, we have applied a small part of the many lessons we learned from working with Wojciech, to connect and merge tools from seemingly different areas of physics to gain a deeper and more fundamental understanding of nature.

Author Contributions

Conceptualization, A.T. and S.D.; Formal analysis, A.T.; Funding acquisition, S.D.; Supervision, S.D.; Writing—original draft, A.T. and S.D.; Writing—review & editing, A.T. and S.D. All authors have read and agreed to the published version of the manuscript.

Funding

S.D. acknowledges support from the U.S. National Science Foundation under Grant No. DMR-2010127. This research was supported by grant number FQXi-RFP-1808 from the Foundational Questions Institute and Fetzer Franklin Fund, a donor advised fund of Silicon Valley Community Foundation (SD).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to thank Wojciech H. Zurek for many years of mentorship and his unwavering patience and willingness to teach us how to think about the mysteries of the quantum universe. Enlightening discussions with Agniva Roychowdhury are gratefully acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A. Cost of Environment-Assisted Shortcuts to Adiabaticity

In this appendix, we show that CD and EASTA have the same cost. Generally, we have the following:
H env ( t ) = i d U ( t ) d t U ( t ) , = i i , j k d U i , k d t ( U j , k ) * | E i ( 0 ) E j ( 0 ) | .
From the main result in Equation (17), we obtain the following:
H env ( t ) = i , j k k ( 0 ) | i t U | ψ i ( t ) ( U j , k ) * + i k ( 0 ) | U | t ψ i ( t ) ( U j , k ) * × | E i ( 0 ) E j ( 0 ) | .
Given that H 0 ( t ) = i d U ( t ) d t U ( t ) , we also have
H env ( t ) = i , j k k ( 0 ) | ( U H 0 ) | ψ i ( t ) ( U j , k ) * + i k ( 0 ) | U | t ψ i ( t ) ( U j , k ) * × | E i ( 0 ) E j ( 0 ) | ,
which implies
H env ( t ) = i , j k k ( 0 ) | ( U H 0 ) | ψ i ( t ) ( U j , k ) * + k ( 0 ) | U H | ψ i ( t ) ( U j , k ) * × | E i ( 0 ) E j ( 0 ) | ,
and hence,
H env ( t ) = i , j k k ( 0 ) | U H CD | ψ i ( t ) ( U j , k ) * | E i ( 0 ) E j ( 0 ) | .
Using | ϕ k ( t ) U ( t ) | k ( 0 ) we can write the following:
H env ( t ) = i , j k ψ j ( t ) | ϕ k ( t ) ϕ k ( t ) | H CD | ψ i ( t ) | E i ( 0 ) E j ( 0 ) | .
Therefore, the following holds:
H env ( t ) = i , j ψ j ( t ) | H CD | ψ i ( t ) | E i ( 0 ) E j ( 0 ) | .
By definition, we also have the following:
H CD = i , j ψ i ( t ) | H CD | ψ j ( t ) | ψ i ( t ) ψ j ( t ) | ,
and hence,
H CD T = H CD * = i , j ψ j ( t ) | H CD | ψ i ( t ) | ψ i ( t ) ψ j ( t ) | .
Thus, there exists a similarity transformation between H CD * and H env , and C CD = C env for any arbitrary driving H 0 ( t ) . The similarity transformation is given by the matrix S = j | E j ( 0 ) ψ j ( t ) | , such that S H CD * S 1 = H env . Since we proved that the Hamiltonians H CD and H env have the same eigenvalues, our result can be valid for other definitions of the cost function C , which might involve other norms (e.g., the Frobenius norm).
It is noteworthy that in the above analysis, we do not consider the effect of quantum fluctuations [78] in the control fields, since their energetic contribution to the cost function is negligible in our context.

Appendix B. Generalization to Arbitrary Branching Probabilities

Finally, we briefly inspect the case of non-even branching states. We begin by noting the consequences of our assumptions. In particular, we have assumed that the state of system+environment evolves unitarly. Thus, consider a joint map of the form U M , where U is a unitary on S . Then, it is a simple exercise to show that the map M, on E , is also unitary, M M = M M = I . In what follows, we prove by contradiction that there exists no unitary map M that suppresses transitions in S , for branching states with arbitrary probabilities.
Consider the following:
| ψ SE ( 0 ) = n = 0 N 1 p n | n ( 0 ) l = 1 N E | E n l ( 0 ) ,
and assume that there exists a unitary map M on E that suppresses transitions in S , i.e.,
n = 0 N 1 p n U | n ( 0 ) M l = 1 N E | E n l ( 0 ) = n = 0 N 1 p n e i f n ( t ) | n ( t ) l = 1 N E | E n l ( 0 ) .
Following the same steps of Section 3, we obtain the following:
( ( m , n ) 0 , N 1 2 ) ; M m , n = p m p n e i f m ( t ) n ( 0 ) | U | m ( t ) .
Comparing the above map with our main result in Equation (17), we conclude that the additional factor p m p n violates unitarity, and hence we conclude that EASTA cannot work for non-even branching states (A10).
This can be seen more explicitly from the form of the matrices M M and M M . Generally, and by dropping the superscript in environmental states l = 1 N E | E n l ( 0 ) | E n ( 0 ) , we have the following:
M M = i , j , k M i , k M j , k * | E i ( 0 ) E j ( 0 ) | ,
from the expression of the elements of M (cf. Equation (A12)), and by adopting the notation | ϕ n U ( t ) | n ( 0 ) , we obtain the following:
M M = i , j , k p i p j p k ϕ k | ψ i ψ j | ϕ k | E i ( 0 ) E j ( 0 ) | ,
which implies
M M = I + i , j p j p i ψ j | D ( i ) | ψ i | E i ( 0 ) E j ( 0 ) | ,
such that the matrix
D ( i ) = k p i p k | ϕ k ϕ k | I
is diagonal in the basis spanned by the orthonormal vectors { | ϕ k } k . This matrix is generally (for any choice of H 0 ( t ) and initial state of S ) different from the null matrix for non-equal branch probabilities. A similar decomposition can be made for the matrix M M , such that
M M = I + i , j p i p j ϕ j | D ( i ) | ϕ i | E i ( 0 ) E j ( 0 ) | ,
where
D ( i ) = k p k p i | ψ k ψ k | I .
In conclusion, for branching states with non-equal probabilities, there is no unitary map that guarantees that the system evolves through the adiabatic manifold at all times and for any arbitrary driving H 0 ( t ) . Hence, we can realize the EASTA technique only for a system maximally entangled with its environment (cf. Equation (A10) with p n = 1 / N for all n 0 , N 1 ), or in the general case (non-equal branch probabilities) when we can access an extended Hilbert space.

References

  1. Deutsch, I.H. Harnessing the Power of the Second Quantum Revolution. PRX Quantum 2020, 1, 020101. [Google Scholar] [CrossRef]
  2. Dowling, J.P.; Milburn, G.J. Quantum technology: The second quantum revolution. Philos. Trans. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. 2003, 361, 1655–1674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Born, M.; Fock, V. Beweis des adiabatensatzes. Z. Phys. 1928, 51, 165–180. [Google Scholar] [CrossRef]
  4. Messiah, A. Quantum Mechanics; John Wiley & Sons: Amsterdam, The Netherlands, 1966; Volume II. [Google Scholar]
  5. Nenciu, G. On the adiabatic theorem of quantum mechanics. J. Phys. A Math. Gen. 1980, 13, L15–L18. [Google Scholar] [CrossRef]
  6. Nenciu, G. On the adiabatic limit for Dirac particles in external fields. Commun. Math. Phys. 1980, 76, 117. [Google Scholar] [CrossRef]
  7. Nenciu, G. Adiabatic theorem and spectral concentration. Commun. Math. Phys. 1981, 82, 121. [Google Scholar] [CrossRef]
  8. Torrontegui, E.; Ibáñez, S.; Martínez-Garaot, S.; Modugno, M.; del Campo, A.; Guéry-Odelin, D.; Ruschhaupt, A.; Chen, X.; Muga, J.G. Shortcuts to Adiabaticity. Adv. At. Mol. Opt. Phys. 2013, 62, 117. [Google Scholar] [CrossRef] [Green Version]
  9. Guéry-Odelin, D.; Ruschhaupt, A.; Kiely, A.; Torrontegui, E.; Martínez-Garaot, S.; Muga, J.G. Shortcuts to adiabaticity: Concepts, methods, and applications. Rev. Mod. Phys. 2019, 91, 045001. [Google Scholar] [CrossRef]
  10. del Campo, A.; Kim, K. Focus on Shortcuts to Adiabaticity. New J. Phys. 2019, 21, 050201. [Google Scholar] [CrossRef]
  11. Deffner, S.; Bonança, M.V.S. Thermodynamic control —An old paradigm with new applications. EPL (Europhys. Lett.) 2020, 131, 20001. [Google Scholar] [CrossRef]
  12. Demirplak, M.; Rice, S.A. Assisted adiabatic passage revisited. J. Phys. Chem. B 2005, 109, 6838. [Google Scholar] [CrossRef]
  13. Demirplak, M.; Rice, S.A. Adiabatic Population Transfer with Control Fields. J. Chem. Phys. A 2003, 107, 9937. [Google Scholar] [CrossRef]
  14. Berry, M.V. Transitionless quantum driving. J. Phys. A Math. Theor. 2009, 42, 365303. [Google Scholar] [CrossRef] [Green Version]
  15. Campbell, S.; De Chiara, G.; Paternostro, M.; Palma, G.M.; Fazio, R. Shortcut to Adiabaticity in the Lipkin-Meshkov-Glick Model. Phys. Rev. Lett. 2015, 114, 177206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. del Campo, A. Shortcuts to Adiabaticity by Counterdiabatic Driving. Phys. Rev. Lett. 2013, 111, 100502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. del Campo, A.; Rams, M.M.; Zurek, W.H. Assisted Finite-Rate Adiabatic Passage Across a Quantum Critical Point: Exact Solution for the Quantum Ising Model. Phys. Rev. Lett. 2012, 109, 115703. [Google Scholar] [CrossRef] [PubMed]
  18. Deffner, S.; Jarzynski, C.; del Campo, A. Classical and Quantum Shortcuts to Adiabaticity for Scale-Invariant Driving. Phys. Rev. X 2014, 4, 021013. [Google Scholar] [CrossRef] [Green Version]
  19. An, S.; Lv, D.; Del Campo, A.; Kim, K. Shortcuts to adiabaticity by counterdiabatic driving for trapped-ion displacement in phase space. Nat. Commun. 2016, 7, 1–5. [Google Scholar] [CrossRef]
  20. Masuda, S.; Nakamura, K. Acceleration of adiabatic quantum dynamics in electromagnetic fields. Phys. Rev. A 2011, 84, 043434. [Google Scholar] [CrossRef]
  21. Masuda, S.; Nakamura, K. Fast-forward of adiabatic dynamics in quantum mechanics. Proc. R. Soc. A 2010, 466, 1135. [Google Scholar] [CrossRef]
  22. Masuda, S.; Rice, S.A. Fast-Forward Assisted STIRAP. J. Phys. Chem. A 2015, 119, 3479. [Google Scholar] [CrossRef] [Green Version]
  23. Masuda, S.; Nakamura, K.; del Campo, A. High-Fidelity Rapid Ground-State Loading of an Ultracold Gas into an Optical Lattice. Phys. Rev. Lett. 2014, 113, 063003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Bernardo, B.d.L. Time-rescaled quantum dynamics as a shortcut to adiabaticity. Phys. Rev. Res. 2020, 2, 013133. [Google Scholar] [CrossRef] [Green Version]
  25. Roychowdhury, A.; Deffner, S. Time-Rescaling of Dirac Dynamics: Shortcuts to Adiabaticity in Ion Traps and Weyl Semimetals. Entropy 2021, 23, 81. [Google Scholar] [CrossRef] [PubMed]
  26. Chen, X.; Ruschhaupt, A.; Schmidt, S.; del Campo, A.; Guéry-Odelin, D.; Muga, J.G. Fast Optimal Frictionless Atom Cooling in Harmonic Traps: Shortcut to Adiabaticity. Phys. Rev. Lett. 2010, 104, 063002. [Google Scholar] [CrossRef] [PubMed]
  27. Torrontegui, E.; Martínez-Garaot, S.; Muga, J.G. Hamiltonian engineering via invariants and dynamical algebra. Phys. Rev. A 2014, 89, 043408. [Google Scholar] [CrossRef] [Green Version]
  28. Kiely, A.; McGuinness, J.P.L.; Muga, J.G.; Ruschhaupt, A. Fast and stable manipulation of a charged particle in a Penning trap. J. Phys. B At. Mol. Opt. Phys. 2015, 48, 075503. [Google Scholar] [CrossRef] [Green Version]
  29. Jarzynski, C.; Deffner, S.; Patra, A.; Subaş ı, Y.b.u. Fast forward to the classical adiabatic invariant. Phys. Rev. E 2017, 95, 032122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Patra, A.; Jarzynski, C. Classical and Quantum Shortcuts to Adiabaticity in a Tilted Piston. J. Phys. Chem. B 2017, 121, 3403–3411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Patra, A.; Jarzynski, C. Shortcuts to adiabaticity using flow fields. New J. Phys. 2017, 19, 125009. [Google Scholar] [CrossRef]
  32. Iram, S.; Dolson, E.; Chiel, J.; Pelesko, J.; Krishnan, N.; Güngör, Ö.; Kuznets-Speck, B.; Deffner, S.; Ilker, E.; Scott, J.G.; et al. Controlling the speed and trajectory of evolution with counterdiabatic driving. Nat. Phys. 2021, 17, 135–142. [Google Scholar] [CrossRef]
  33. Zurek, W.H. Environment-Assisted Invariance, Entanglement, and Probabilities in Quantum Physics. Phys. Rev. Lett. 2003, 90, 120404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Zurek, W.H. Probabilities from entanglement, Born’s rule pk = ∣ψk2 from envariance. Phys. Rev. A 2005, 71, 052105. [Google Scholar] [CrossRef] [Green Version]
  35. Deffner, S.; Zurek, W.H. Foundations of statistical mechanics from symmetries of entanglement. New J. Phys. 2016, 18, 063013. [Google Scholar] [CrossRef] [Green Version]
  36. Vermeyden, L.; Ma, X.; Lavoie, J.; Bonsma, M.; Sinha, U.; Laflamme, R.; Resch, K.J. Experimental test of environment-assisted invariance. Phys. Rev. A 2015, 91, 012120. [Google Scholar] [CrossRef] [Green Version]
  37. Harris, J.; Bouchard, F.; Santamato, E.; Zurek, W.H.; Boyd, R.W.; Karimi, E. Quantum probabilities from quantum entanglement: Experimentally unpacking the Born rule. New J. Phys. 2016, 18, 053013. [Google Scholar] [CrossRef] [Green Version]
  38. Blume-Kohout, R.; Zurek, W.H. Quantum Darwinism: Entanglement, branches, and the emergent classicality of redundantly stored quantum information. Phys. Rev. A 2006, 73, 062310. [Google Scholar] [CrossRef] [Green Version]
  39. Abah, O.; Puebla, R.; Kiely, A.; De Chiara, G.; Paternostro, M.; Campbell, S. Energetic cost of quantum control protocols. New J. Phys. 2019, 21, 103048. [Google Scholar] [CrossRef]
  40. Zheng, Y.; Campbell, S.; De Chiara, G.; Poletti, D. Cost of counterdiabatic driving and work output. Phys. Rev. A 2016, 94, 042132. [Google Scholar] [CrossRef] [Green Version]
  41. Campbell, S.; Deffner, S. Trade-Off Between Speed and Cost in Shortcuts to Adiabaticity. Phys. Rev. Lett. 2017, 118, 100601. [Google Scholar] [CrossRef] [Green Version]
  42. Abah, O.; Puebla, R.; Paternostro, M. Quantum State Engineering by Shortcuts to Adiabaticity in Interacting Spin-Boson Systems. Phys. Rev. Lett. 2020, 124, 180401. [Google Scholar] [CrossRef]
  43. Ilker, E.; Güngör, Ö.; Kuznets-Speck, B.; Chiel, J.; Deffner, S.; Hinczewski, M. Counterdiabatic control of biophysical processes. arXiv 2021, arXiv:2106.07130. [Google Scholar]
  44. Touil, A.; Yan, B.; Girolami, D.; Deffner, S.; Zurek, W.H. Eavesdropping on the Decohering Environment: Quantum Darwinism, Amplification, and the Origin of Objective Classical Reality. arXiv 2021, arXiv:2107.00035. [Google Scholar]
  45. Barnes, E.; Das Sarma, S. Analytically Solvable Driven Time-Dependent Two-Level Quantum Systems. Phys. Rev. Lett. 2012, 109, 060401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Greilich, A.; Economou, S.E.; Spatzek, S.; Yakovlev, D.; Reuter, D.; Wieck, A.; Reinecke, T.; Bayer, M. Ultrafast optical rotations of electron spins in quantum dots. Nat. Phys. 2009, 5, 262–266. [Google Scholar] [CrossRef] [Green Version]
  47. Poem, E.; Kenneth, O.; Kodriano, Y.; Benny, Y.; Khatsevich, S.; Avron, J.E.; Gershoni, D. Optically Induced Rotation of an Exciton Spin in a Semiconductor Quantum Dot. Phys. Rev. Lett. 2011, 107, 087401. [Google Scholar] [CrossRef] [Green Version]
  48. Martinis, J.M.; Cooper, K.B.; McDermott, R.; Steffen, M.; Ansmann, M.; Osborn, K.D.; Cicak, K.; Oh, S.; Pappas, D.P.; Simmonds, R.W.; et al. Decoherence in Josephson Qubits from Dielectric Loss. Phys. Rev. Lett. 2005, 95, 210503. [Google Scholar] [CrossRef] [Green Version]
  49. Petta, J.R.; Johnson, A.C.; Taylor, J.M.; Laird, E.A.; Yacoby, A.; Lukin, M.D.; Marcus, C.M.; Hanson, M.P.; Gossard, A.C. Coherent manipulation of coupled electron spins in semiconductor quantum dots. Science 2005, 309, 2180–2184. [Google Scholar] [CrossRef] [Green Version]
  50. Foletti, S.; Bluhm, H.; Mahalu, D.; Umansky, V.; Yacoby, A. Universal quantum control of two-electron spin quantum bits using dynamic nuclear polarization. Nat. Phys. 2009, 5, 903–908. [Google Scholar] [CrossRef] [Green Version]
  51. Maune, B.M.; Borselli, M.G.; Huang, B.; Ladd, T.D.; Deelman, P.W.; Holabird, K.S.; Kiselev, A.A.; Alvarado-Rodriguez, I.; Ross, R.S.; Schmitz, A.E.; et al. Coherent singlet-triplet oscillations in a silicon-based double quantum dot. Nature 2012, 481, 344–347. [Google Scholar] [CrossRef]
  52. Wang, X.; Bishop, L.S.; Kestner, J.; Barnes, E.; Sun, K.; Sarma, S.D. Composite pulses for robust universal control of singlet–triplet qubits. Nat. Commun. 2012, 3, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Zeh, H.D. On the interpretation of measurement in quantum theory. Found. Phys. 1970, 1, 69–76. [Google Scholar] [CrossRef]
  54. Zurek, W.H. Decoherence, einselection, and the quantum origins of the classical. Rev. Mod. Phys. 2003, 75, 715–775. [Google Scholar] [CrossRef] [Green Version]
  55. Zurek, W.H. Preferred States, Predictability, Classicality and the Environment-Induced Decoherence. Prog. Theo. Phys. 1993, 89, 281–312. [Google Scholar] [CrossRef]
  56. Zurek, W.H. Pointer basis of quantum apparatus: Into what mixture does the wave packet collapse? Phys. Rev. D 1981, 24, 1516–1525. [Google Scholar] [CrossRef]
  57. Zurek, W.H. Environment-induced superselection rules. Phys. Rev. D 1982, 26, 1862–1880. [Google Scholar] [CrossRef]
  58. Alipour, S.; Chenu, A.; Rezakhani, A.T.; del Campo, A. Shortcuts to adiabaticity in driven open quantum systems: Balanced gain and loss and non-Markovian evolution. Quantum 2020, 4, 336. [Google Scholar] [CrossRef]
  59. Santos, A.C.; Sarandy, M.S. Generalized transitionless quantum driving for open quantum systems. arXiv 2021, arXiv:2109.11695. [Google Scholar]
  60. Yin, Z.; Li, C.; Zhang, Z.; Zheng, Y.; Gu, X.; Dai, M.; Allcock, J.; Zhang, S.; An, S. Shortcuts to Adiabaticity for Open Systems in Circuit Quantum Electrodynamics. arXiv 2021, arXiv:2107.08417. [Google Scholar]
  61. Riedel, C.J.; Zurek, W.H. Quantum Darwinism in an Everyday Environment: Huge Redundancy in Scattered Photons. Phys. Rev. Lett. 2010, 105, 020404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Riedel, C.J.; Zurek, W.H. Redundant information from thermal illumination: Quantum Darwinism in scattered photons. New J. Phys. 2011, 13, 073038. [Google Scholar] [CrossRef]
  63. Zurek, W.H. Quantum Darwinism. Nat. Phys. 2009, 5, 181. [Google Scholar] [CrossRef]
  64. Zurek, W.H. Quantum Darwinism, classical reality, and the randomness of quantum jumps. arXiv 2014, arXiv:1412.5206. [Google Scholar] [CrossRef] [Green Version]
  65. Zurek, W.H. Quantum origin of quantum jumps: Breaking of unitary symmetry induced by information transfer in the transition from quantum to classical. Phys. Rev. A 2007, 76, 052110. [Google Scholar] [CrossRef] [Green Version]
  66. Brandao, F.G.; Piani, M.; Horodecki, P. Generic emergence of classical features in quantum Darwinism. Nat. Commun. 2015, 6, 1–8. [Google Scholar] [CrossRef] [Green Version]
  67. Riedel, C.J.; Zurek, W.H.; Zwolak, M. The rise and fall of redundancy in decoherence and quantum Darwinism. New J. Phys. 2012, 14, 083010. [Google Scholar] [CrossRef]
  68. Milazzo, N.; Lorenzo, S.; Paternostro, M.; Palma, G.M. Role of information backflow in the emergence of quantum Darwinism. Phys. Rev. A 2019, 100, 012101. [Google Scholar] [CrossRef] [Green Version]
  69. Kiciński, M.; Korbicz, J.K. Decoherence and objectivity in higher spin environments. arXiv 2021, arXiv:2105.09093. [Google Scholar]
  70. Korbicz, J. Roads to objectivity: Quantum Darwinism, Spectrum Broadcast Structures, and Strong quantum Darwinism. arXiv 2020, arXiv:2007.04276. [Google Scholar]
  71. Blume-Kohout, R.; Zurek, W.H. Quantum Darwinism in Quantum Brownian Motion. Phys. Rev. Lett. 2008, 101, 240405. [Google Scholar] [CrossRef] [Green Version]
  72. Zwolak, M.; Riedel, C.J.; Zurek, W.H. Amplification, Redundancy, and Quantum Chernoff Information. Phys. Rev. Lett. 2014, 112, 140406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Zwolak, M.; Quan, H.T.; Zurek, W.H. Quantum Darwinism in a Mixed Environment. Phys. Rev. Lett. 2009, 103, 110402. [Google Scholar] [CrossRef] [Green Version]
  74. Paz, J.P.; Roncaglia, A.J. Redundancy of classical and quantum correlations during decoherence. Phys. Rev. A 2009, 80, 042111. [Google Scholar] [CrossRef] [Green Version]
  75. Riedel, C.J.; Zurek, W.H.; Zwolak, M. Objective past of a quantum universe: Redundant records of consistent histories. Phys. Rev. A 2016, 93, 032126. [Google Scholar] [CrossRef] [Green Version]
  76. Riedel, C.J. Classical Branch Structure from Spatial Redundancy in a Many-Body Wave Function. Phys. Rev. Lett. 2017, 118, 120402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Fu, H.F. Uniqueness of the observable leaving redundant imprints in the environment in the context of quantum Darwinism. Phys. Rev. A 2021, 103, 042210. [Google Scholar] [CrossRef]
  78. Calzetta, E. Not-quite-free shortcuts to adiabaticity. Phys. Rev. A 2018, 98, 032107. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Sketch of the two different schemes of applying STA in a branching state of the form presented in  Equation (6). In both panels, the state preparation involves a Hadamard gate (H) applied on S , and coupling with the environment through a c-not operation. In panel (a), we describe the “usual” counterdiabatic scheme. As shown in Section 3, local driving on E suppresses any transitions of S in the instantaneous eigenbasis of H 0 ( t ) . The latter scheme is illustrated in panel (b).
Figure 1. Sketch of the two different schemes of applying STA in a branching state of the form presented in  Equation (6). In both panels, the state preparation involves a Hadamard gate (H) applied on S , and coupling with the environment through a c-not operation. In panel (a), we describe the “usual” counterdiabatic scheme. As shown in Section 3, local driving on E suppresses any transitions of S in the instantaneous eigenbasis of H 0 ( t ) . The latter scheme is illustrated in panel (b).
Entropy 23 01479 g001
Figure 2. In panel (a), the blue curve illustrates the overlap between the nth evolved state | ϕ n ( t ) and the nth instantaneous eigenstate | n ( t ) of the Hamiltonian H 0 ( t ) . This curve shows that the driving H 0 ( t ) evolves the system beyond the adiabatic manifold. The red curve illustrates that EASTA guarantees an exact evolution through the adiabatic manifold. In panel (b), we illustrate the cost of both the EASTA and the CD schemes, and numerically verify that they are equal C = C CD = C env for all t / τ [ 0 , 1 ] , and τ = 1 . Note that in the illustrations we pick the driving field J ( t ) = B cos 2 π t 2 τ and B = 1 . In panel (c), we illustrate the costs for different values of τ . For infinitely fast processes ( τ 0 ) the cost diverges and it tends to zero for infinitely slow processes ( τ ).
Figure 2. In panel (a), the blue curve illustrates the overlap between the nth evolved state | ϕ n ( t ) and the nth instantaneous eigenstate | n ( t ) of the Hamiltonian H 0 ( t ) . This curve shows that the driving H 0 ( t ) evolves the system beyond the adiabatic manifold. The red curve illustrates that EASTA guarantees an exact evolution through the adiabatic manifold. In panel (b), we illustrate the cost of both the EASTA and the CD schemes, and numerically verify that they are equal C = C CD = C env for all t / τ [ 0 , 1 ] , and τ = 1 . Note that in the illustrations we pick the driving field J ( t ) = B cos 2 π t 2 τ and B = 1 . In panel (c), we illustrate the costs for different values of τ . For infinitely fast processes ( τ 0 ) the cost diverges and it tends to zero for infinitely slow processes ( τ ).
Entropy 23 01479 g002aEntropy 23 01479 g002b
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Touil, A.; Deffner, S. Environment-Assisted Shortcuts to Adiabaticity. Entropy 2021, 23, 1479. https://doi.org/10.3390/e23111479

AMA Style

Touil A, Deffner S. Environment-Assisted Shortcuts to Adiabaticity. Entropy. 2021; 23(11):1479. https://doi.org/10.3390/e23111479

Chicago/Turabian Style

Touil, Akram, and Sebastian Deffner. 2021. "Environment-Assisted Shortcuts to Adiabaticity" Entropy 23, no. 11: 1479. https://doi.org/10.3390/e23111479

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop