3,718
Views
2
CrossRef citations to date
0
Altmetric
Reviews

Quantum control in open and periodically driven systems

, , & ORCID Icon
Article: 1870559 | Received 15 Oct 2020, Accepted 24 Dec 2020, Published online: 11 Jan 2021

ABSTRACT

Quantum technology resorts to efficient utilization of quantum resources to realize technique innovation. The systems are controlled such that their states follow the desired manners to realize different quantum protocols. However, the decoherence caused by the system-environment interactions causes the states deviating from the desired manners. How to protect quantum resources under the coexistence of active control and passive decoherence is of significance. Recent studies have revealed that the decoherence is determined by the feature of the system-environment energy spectrum: Accompanying the formation of bound states in the energy spectrum, the decoherence can be suppressed. It supplies a guideline to control decoherence. Such idea can be generalized to systems under periodic driving. By virtue of manipulating Floquet bound states in the quasienergy spectrum, coherent control via periodic driving dubbed as Floquet engineering has become a versatile tool not only in controlling decoherence, but also in artificially synthesizing exotic topological phases. We will review the progress on quantum control in open and periodically driven systems. Special attention will be paid to the distinguished role played by the bound states and their controllability via periodic driving in suppressing decoherence and generating novel topological phases.

1. Introduction

Quantum technology is a class of technology that works by using quantum effects, such as quantum coherence, squeezing, and entanglement, to develop revolutionary techniques. Including quantum computation and algorithm [Citation1–4], quantum metrology and sensor [Citation5–8], quantum simulation [Citation9], and quantum communication [Citation10,Citation11], it opens an avenue to break through the technique limits for their classical counterparts from the principles of quantum mechanics. The main idea is to actively control the quantum states of relevant systems in the desired manners to realize different protocols. However, its realization is challenged by decoherence [Citation12]. As a ubiquitous phenomenon in the microscopic world, decoherence is caused by the inevitable interaction of quantum systems with their environments. It makes the quantum resources degraded and the states deviating from the desired manners. Therefore, a thorough understanding of the detrimental influences of decoherence on quantum protocols and their efficient control are of importance in quantum technology. Many methods, such as entanglement concentration [Citation13,Citation14], dynamical decoupling [Citation15,Citation16], and feedback control [Citation17,Citation18], have been proposed to suppress the effects of decoherence. Recent studies show that the structured environment, e.g. photonic crystal [Citation19–22] and localized surface plasmons [Citation23,Citation24], dramatically changes or even inhibits the decoherence of quantum emitters [Citation20,Citation25,Citation26]. Further studies reveal that these phenomena are due to the formation of a bound state in the energy spectrum of the total system [Citation27]. It has been reported that such a bound state has profound impacts on many quantum protocols and nonequilibrium physics, e.g. entanglement preservation [Citation28–30], noncanonical thermalization [Citation31], quantum speed limit [Citation32], and quantum metrology [Citation33–35]. The results give us an insightful guideline to control decoherence via forming the bound state by quantum reservoir engineering [Citation36,Citation37].

On the other hand, coherent control via periodic driving dubbed as Floquet engineering has become a versatile tool in quantum control [Citation38]. The main idea resides in that the concepts of the energy spectrum and bound states of static systems can be completely inherited in the quasienergy spectrum of periodically driven systems according to Floquet theory. The bound-state mechanism in static systems was generalized to the periodically driven systems as the Floquet bound-state mechanism [Citation39–42]. Periodic driving largely extends the controllability of systems because time as a new control dimension is introduced. It overcomes the practical difficulty of reservoir engineering in static systems that the parameter is hard to change once the material is fabricated [Citation43–45]. Engineering the quasienergy spectrum to form the Floquet bound states via periodic driving can also be applied to artificially synthesizing topological phases. The periodic driving induces an effective long-range hopping in lattice systems which is hard to experimentally realize in the static system [Citation46–49]. In this manner, widely tunable numbers of edge or surface states have been realized by periodic driving [Citation47–53]. It is found that not only the topological phases hard to reach in the same setting of static systems but also many exotic topological phases totally absent in static systems can be realized by Floquet engineering [Citation50,Citation54,Citation55].

We here review the progress on quantum control in open and periodically driven systems. The paper is organized as follows. In Sec. 2, the bound-state mechanism is given and the decoherence control based on the reservoir engineering is reviewed. The profound impacts of the bound state on entanglement preservation, quantum speedup, noncanonical thermalization, quantum metrology, and quantum plasmonics are explicitly analyzed. The Floquet engineering is reviewed in Sec. 3. The applications of Floquet engineering in decoherence control and the generation of exotic topological phases are explicitly discussed. Conclusion and outlook are made in Sec. 4.

2. Decoherence control in open quantum systems

Quantum technology is based on the efficient utilization of quantum resources, e.g. quantum coherence and entanglement, of the relevant systems to realize the technique innovation. However, any microscopic system is inevitably influenced by its surrounding environment and becomes an open system. It causes unwanted loss of quantum resources. Therefore, one of the important purposes of quantum control in open systems is to suppress decoherence. Depending on whether the open system has energy exchange with the environment, its decoherence can be classified into dissipation and dephasing. We focus on dissipation.

Open system influenced by its environment has a wide physical relevance in different experimental platforms of quantum technology. There are two types of open systems: Discrete- and continuous-variable systems. The discrete-variable system is generally a two-level system, which may be a two-level atom interacting with the radiation electromagnetic field [Citation56], an electron spin in a quantum dot interacting with the nuclear spin [Citation57,Citation58], or a superconducting charge qubit exposed to the charge noise [Citation59]. The continuous-variable one may be a harmonic oscillator or a single-mode quantized optical field. The microscopic description of dissipation is based on the ‘system’ + ‘reservoir’ approach pioneered by Senitzky [Citation60,Citation61]. The Hamiltonian is (=1)

(1) Hˆ=ω0oˆoˆ+kgk(oˆaˆk+aˆkoˆ)+kωkaˆkaˆk,(1)

where oˆ and aˆk are the annihilation operator of the system with frequency ω0 and the k-th environmental mode with frequency ωk, and gk is the coupling strength. The coupling can be further characterized by the spectral density J(ω)=k|gk|2δ(ωωk) [Citation62]. The operator oˆ=σˆ with σˆ=|ge| is the transition operator between the two levels for the discrete-variable system, oˆ=bˆ is the annihilation operator for the bosonic continuous-variable system. One can check that the total excitation number Nˆ=σˆ+σˆ+kaˆkaˆk for both of the systems is conserved. Thus, the Hilbert space is split into the subspaces with a definite excitation number N.

The dynamics of the open system is obtainable by tracing out the environmental degrees of freedom from ρT(t) satisfying the Liouville equation ρ˙T(t)=i[Hˆ,ρT(t)]. However, it is difficult to be exactly done. A widely used reduction is the Markovian approximation [Citation63–65], which works under the conditions that the system-environment coupling is weak and the environmental correlation time is much smaller than the one of the systems. Many systems, such as quantum dot [Citation57,Citation58] and superconducting qubit [Citation66], have shown the conditions where this approximation is inapplicable. It inspires particular attention to the non-Markovian dynamics [Citation62,Citation67].

Supposing that the coupling is turned on at the initial time, we have ρT(0)=ρ(0)ρE, where ρ(0) and ρE are the initial state of system and environment, respectively. We consider that the environment is initially in a vacuum state. The dynamics governed by EquationEq. (1) in this case for the discrete-variable system is exactly solvable because only the Hilbert subspaces with N=0 and 1 are involved [Citation28]. The exact dynamics for the continuous-variable system is obtainable by the Feynman-Vernon influence functional theory [Citation68] in the coherent-state representation [Citation69]. After tracing out the environmental degrees of freedom, we obtain the non-Markovian master equation for both of the systems as

(2) ρ˙(t)=iω(t)[oˆoˆ,ρ(t)]+γ(t)2[2oˆρ(t)oˆ{oˆoˆ,ρ(t)}],(2)

where γ(t)iω(t)=u˙(t)/u(t) with u(t) satisfying

(3) u˙(t)+iω0u(t)+0tf(tτ)u(τ)dτ=0,(3)

under u(0)=1. Here the kernel function is f(tτ)=0J(ω)eiω(tτ)dω.

A Laplace transform can linearize EquationEq. (3) into u˜(z)=[z+iω0+0J(ω)z+iωdω]1. The solution of u(t) is obtained by the inverse Laplace transform of u˜(z), which can be done by finding its pole from

(4) y(E)ω00J(ω)ωEdω=E,(E=iz).(4)

Note that the roots E of EquationEq. (4) is just the eigenenergies of the total system (1) in the single-excitation space. Specifically, expanding the eigenstate as |Ψ=(xoˆ+kykbˆk)|0,{0k} and substituting it into Hˆ|Ψ=E|Ψ with E being the eigenenergy, we have (Eω0)x=kgkyk and yk=gkx/(Eωk). They readily lead to EquationEq. (4). Since y(E) is a decreasing function in the regime E<0, EquationEq. (4) has one isolated root Eb in this regime provided y(0)<0. While y(E) is ill defined when E>0, EquationEq. (4) has infinite roots in this regime forming a continuous energy band. We call the eigenstate of the isolated eigenenergy Eb bound state [Citation28,Citation34]. After the inverse Laplace transform, we obtain

(5) u(t)=ZeiEbt+0J(E)eiEtdE[Eω0Δ(E)]2+[πJ(E)]2,(5)

where Δ(E)=P0J(ω)Eωdω with P being the Cauchy principal value. The first term with Z=[1+0J(ω)dω(Ebω)2]1 is from the bound state, and the second one is from the band energies. Oscillating with time in continuously changing frequencies, the integral tends to zero in the long-time condition due to out-of-phase interference. Thus, if the bound state is absent, then limtu(t)=0 characterizes a complete dissipation, while if the bound state is formed, then limtu(t)=ZeiEbt implies a dissipation suppression. A widely used type of spectral density reads J(ω)=ηωsωc1seω/ωc, where η is a dimensionless coupling constant, ωc is a cutoff frequency, and s is the Ohmicity parameter. Depending on the value of s, the environments are classified into a sub-Ohmic one (0<s<1), an Ohmic one (s=1), and a super-Ohmic one (s>1) [Citation70]. We can check that the bound state is formed for the Ohmic-family spectral density when ω0<ηωcΓ_(s), where Γ_(s) is the Euler’s gamma function.

From the above analysis, we see that the non-Markovian dissipation dynamics for both of the discrete- and continuous-variable systems intrinsically depends on the character of the energy spectrum of the total system. In the following, we review the profound impact of the bound state on the different nonequilibrium physics of the systems.

2.1. Entanglement preservation

Quantum entanglement describes the nonclassical correlations between quantum subsystems, which was first recognized by Einstein, Podolsky, and Rosen [Citation71]. It is an important resource to implement quantum protocols [Citation72], e.g. quantum communication [Citation10,Citation11,Citation73–75], quantum computation [Citation1–3], and quantum metrology [Citation5–8]. However, it is generally degraded by the ubiquitous decoherence in the microscopic world. It was found that, quite different from the asymptotically decaying to zero of quantum coherence of a single two-level system, the entanglement of a bipartite two-level system influenced by two independent Markovian dissipative environments abruptly disappears in a finite time duration [Citation76–78]. It is called entanglement sudden death. Further studies revealed that the dead entanglement can partially revive due to the non-Markovian effect [Citation79–81]. Such finite extension of the entangled time is obviously not enough for quantum protocols and one generally desires to preserve the entanglement in the long-time limit. Indeed, it was shown that some entanglement can be preserved in the steady state by engineering a photonic crystal-structured environment [Citation82]. Actually, as demonstrated in the following, the entanglement preservation is a universal consequence of the existence of the system-environment bound state and the non-Markovian effect. The bound state provides the ability and the non-Markovian effect provides the dynamical way to preserve the entanglement [Citation28,Citation29,Citation83].

Consider two independent two-level systems interacting with their respective dissipative environment. The dynamics of each two-level system is governed by EquationEq. (2) with oˆ=σˆ [Citation83]. Solving the master equation under the initial condition |ψ(0)=12(|gg+|ee), we obtain

(6) ρ(t)=12{[|u(t)|2|ee|+(1|u(t)|2)|gg|]2+|gg|2+[u(t)2|eg|2+H.c.]}.(6)

The entanglement of a bipartite two-level system is quantified by concurrence [Citation84]. It is defined as C(ρ)=max{0,λ1λ2λ3λ4}, where λi is the i-th decreasing ordered eigenvalues of Rρρ˜ρ and ρ˜ is defined as ρ˜(σˆyσˆy)ρ(σˆyσˆy) [Citation72,Citation85]. Then we can obtain that the concurrence of EquationEq. (6) is C(t)=max{0,|u(t)|4}. Combining with the knowledge that u(t) is preserved when the bound state exists, we find that the formation of the bound state can also preserve the entanglement, which paves the way to stably protect entanglement from dissipation.

Figure 1. (a) Diagrammatic solution of EquationEq. (4) with different parameters of a super-Ohmic spectral density s=3. Entanglement evolution with parameters in the absence (b, c) and presence (d) of the bound state. Reproduced figures from [Citation28]

Figure 1. (a) Diagrammatic solution of EquationEq. (4)(4) y(E)≡ω0−∫0∞J(ω)ω−Edω=E,(E=iz).(4) with different parameters of a super-Ohmic spectral density s=3. Entanglement evolution with parameters in the absence (b, c) and presence (d) of the bound state. Reproduced figures from [Citation28]

Taking the super-Ohmic spectral density J(ω)=ηωsω01seω/ωc with s=3, where the bound state is formed when ω0ηω0(ωc/ω0)sΓ_(s)<0, as an example [Citation28], ) shows the diagrammatic solution of EquationEq. (4) for different parameter values. It reveals that as long as ω0<2ηω0(ωc/ω0)3, the bound state is formed. The corresponding entanglement dynamics are shown in ). Under the Markovian approximation, the concurrence always decays to zero irrespective of the existence of the bound state, whereas the exact non-Markovian dynamics reveals that the entanglement is preserved when the bound state is formed. The one-to-one correspondence between the formation of bound state and entanglement preservation clearly signifies the ability of the bound state in protecting the entanglement from decoherence.

Note that the entanglement preservation mechanism given above is also applicable to other spectral density [Citation28] and other systems, e.g. nitrogen-vacancy centers embedded into the planar photonic-crystal cavities [Citation30], continuous-variable system [Citation29,Citation86], and quantum emitters coupled to the localized surface plasmons [Citation26].

2.2. Quantum speed limit

The Heisenberg’s uncertainty principle sets a fundamental limit to measurement precision. One well-known example is that the precision of position is limited by the uncertainty of momentum as Δx/(2Δp). Such limitation also exists for the detection of time, which is known as the quantum speed limit (QSL) [Citation87–90]. It sets a bound on the minimal time a system needs to evolve between two distinguishable states [Citation91,Citation92]. The QSL time for isolated systems is determined by the maximum of the Mandelstam-Tamn bound τML=π/(2ΔE) [Citation88,Citation89] and Margolus-Levitin bound τML=π/(2Eˉ) [Citation92], where ΔE and Eˉ are the variation and the average of the Hamiltonian over the initial state, respectively. The QSL time for open systems characterizes the most efficient response of the system to the environmental influences. The Mandelstam-Tamm-type bound is generalized to open systems by using positive non-unitary maps [Citation93,Citation94]. Based on a geometric approach, a unified bound including both Mandelstam-Tamm and Margolus-Levitin types has been formulated [Citation95]. The QSL has attracted considerable attention because the efficient speedup of quantum systems plays remarkable roles in various areas of quantum physics, including nonequilibrium thermodynamics [Citation96], quantum metrology [Citation97], quantum optimal control [Citation98], quantum computation [Citation99], and quantum communication [Citation100].

The previous studies revealed that the non-Markovian effect characterized by non-Markovianity can speed up the evolution of open systems [Citation101,Citation102]. As a dynamical quantity, the non-Markovianity could not be seen as an essential reflection on the QSL of open systems. Furthermore, such attribution of the quantum speedup to the non-Markovianity is experimentally meaningless because the non-Markovianity is not observable. Actually, it is the formation of the system-environment bound state that is the essential reason for the quantum speedup.

Figure 2. (a) Non-Markovianity N (blue dashed line) and QSL time τQSL (read solid line) for the Ohmic spectral density as a function of coupling constant η. The inset is the energy spectrum of the total system, where green dot-dashed line denotes the bound state. The obtained |u(t)|2 (b) and the corresponding decay rate γ(t) (c) with and without bound state. Parameters in (a), (b), and (c) are ωcτ=800, ω0/ωc=0.1, and s=1, which determines that the bound state is formed when η>0.1. QSL time τQSL (d) and non-Markovianity N (e) as a function of the spectral power index s and the system frequency ω0 for the spectral density J(ω)=ηωsωc1seω/ωc. The red dashed line shows the critical values for forming the bound state. Parameters in (d) and (e) are ωcτ=800 and η=0.2. Reproduced figures from [Citation32]

Figure 2. (a) Non-Markovianity N (blue dashed line) and QSL time τQSL (read solid line) for the Ohmic spectral density as a function of coupling constant η. The inset is the energy spectrum of the total system, where green dot-dashed line denotes the bound state. The obtained |u(t)|2 (b) and the corresponding decay rate γ(t) (c) with and without bound state. Parameters in (a), (b), and (c) are ωcτ=800, ω0/ωc=0.1, and s=1, which determines that the bound state is formed when η>0.1. QSL time τQSL (d) and non-Markovianity N (e) as a function of the spectral power index s and the system frequency ω0 for the spectral density J(ω)=ηωsωc1−se−ω/ωc. The red dashed line shows the critical values for forming the bound state. Parameters in (d) and (e) are ωcτ=800 and η=0.2. Reproduced figures from [Citation32]

To demonstrate this result, we consider again a dissipative two-level system [Citation32]. Its dynamics follows the master Equationequation (2) with oˆ=σˆ. The QSL time between an initial state ρ(0)=|ψ0ψ0| and its target state ρ(τ) is defined as τQSL=sin2B[ρ(0),ρ(τ)]/Λτ() [Citation101], where B[ρ(0),ρ(τ)]arccosψ0|ρ(τ)|ψ0 is the Bures angle between ρ(0) and ρ(τ), and Λτ()=(1/τ)0τdt||ρ˙(t)|| with Aˆ equaling to the largest eigenvalue of AˆAˆ. From EquationEq. (2), the QSL time for this system

(7) τQSL=1|u(τ)|2(1/τ)0τ|t|u(t)|2|dt,(7)

is achieved when |ψ0=|e. In order to make a comparison with the previous works, we also calculate the non-Markovianity defined as N=maxρ1,2(0)σ>0dtσ(t,ρ1,2(0)), where σ(t,ρ1,2(0))=D˙(ρ1(t),ρ2(t)) is the change rate of the trace distance D(ρ1(t),ρ2(t))=12Tr|ρ1(t)ρ2(t)| between states ρ1,2(t) evolving from their respective initial forms ρ1,2(0) [Citation103]. For the two-level system, it has been proven that the optimal pair of initial states to maximize N are ρ1(0)=|gg| and ρ2(0)=|ee| [Citation101], which means D(ρ1(t),ρ2(t))=|u(t)|2. Then one can readily verify N=12[|u(τ)|21+0τ|t|u(t)|2|dt], which connects to τQSL as τQSL=τ1|u(τ)|21|u(τ)|2+2N.

The numerical results for the Ohmic-family spectral density are shown in . From ) one can find that both the non-Markovianity and quantum speedup show a perfect correspondence with the formation of bound state. In the absence of the bound state, there is no non-Makrovianity and quantum speedup. Whereas, an abrupt increase of the non-Markovianity and an abrupt decrease of the QSL time are observed when the bound state is formed. The corresponding excited-state population Tr[σˆ+σˆρ(t)]=|u(t)|2 and the decay rate γ(t) are plotted in , c), respectively. We can see that γ(t) in the absence of the bound state tends to a positive constant after a short-time jolt. Its complete positivity causes |u(t)|2 to decay to zero monotonically. Since the system equilibrates in an asymptotic manner to its ground state, there is no more efficient evolution than the evolution characterized by τ and thus τQSL=τ. When the bound state is formed, the competition between the environmental backaction and the dissipation causes γ(t) to transiently take a negative value and asymptotically approaches zero. Thus, after some short-time oscillations, |u(t)|2 tends to a finite value. The transient increase of |u(t)|2 causes the increase of the distinguishability and the decrease of τQSL. The system here has a great capacity to speedup. Hence the bound state supplies a latent capability, while the non-Markovianity only supplies the dynamical way to the system for quantum speedup. The result is confirmed in the general cases with changing ω0 and s in , e).

The relationship between the bound state and quantum speedup reveals a mechanism for quantum speedup. It may be of both theoretical and experimental interests in exploring the ultimate QSL in realistic environments, and may open new perspectives for devising active quantum speedup devices.

2.3. Noncanonical thermalization

In statistical mechanics, systems in equilibrium with its weakly coupled environment are described by a canonical state at the same temperature as the environment. Being called canonical thermalization, this is based on the assumption of ergodicity. The assumption relies on the subjective lack of knowledge of systems. Therefore, how and when systems microscopically equilibrate from any initial state to a unique canonical state has attracted much attention [Citation104–115]. It was revealed that the assumption of ergodicity could be abandoned by examining the entanglement induced by the system–environment interaction, by which the canonical state can be achieved without referring to ensembles or time averages [Citation104]. It inspires the studies of canonical thermalization from the dynamics of open quantum systems [Citation116–120]. It is expected that equilibrium statistical physics should be an emergent steady-state limit of the nonequilibrium dynamics. It was shown that the canonical thermalization is valid only in the limit of the vanishing coupling and the memoryless correlation function of the environment [Citation116,Citation118]. However, in real physical situations, e.g., photonic crystal [Citation121] and semiconductor phonon [Citation122] environments, low-dimensional solid-state system [Citation123], and optical systems [Citation124,Citation125], these conditions are generally invalid and the strong memory effect of the environment may even cause the anomalous decoherence behavior [Citation25,Citation26,Citation126]. Therefore, to microscopically bridge between an arbitrary initial state of the open system and its equilibrium state via studying its equilibration dynamics is desired to check the validity regime of the assumption of ergodicity [Citation31].

Figure 3. Γ(t) (a) and Γβ(t) (b) in different ωc. (c) Teff evaluated when t=100/ω0. η=0.1, β=0.1/ω0, and s=1 have been used. Mean photon number N(t) in different initial condition α0 when η=0.05 (d) and 0.2 (e), where the red dashed line is obtained from EquationEq. (10). Reproduced figure from [Citation31]

Figure 3. Γ(t) (a) and Γβ(t) (b) in different ωc. (c) Teff evaluated when t=100/ω0. η=0.1, β=0.1/ω0, and s=1 have been used. Mean photon number N(t) in different initial condition α0 when η=0.05 (d) and 0.2 (e), where the red dashed line is obtained from EquationEq. (10)(10) ρcon≡ρ(∞)=e−βeffHˆs/Trse−βeffHˆs,(βeff=1/kBTeff),(10) . Reproduced figure from [Citation31]

To address this issue, we consider a continuous-variable system described by a quantum harmonic oscillator interacting with a finite-temperature environment [Citation31]. The Hamiltonian follows EquationEq. (1) with oˆ=bˆ. After generalizing EquationEq. (2) to the finite-temperature case, we obtain the exact non-Markovian master equation as

(8) ρ˙(t)=iΩ(t)[bˆbˆ,ρ(t)]+[Γ(t)+Γβ(t)2][2bˆρ(t)bˆ{bˆbˆ,ρ(t)}]+Γβ(t)2[2bˆρ(t)bˆ{bˆbˆ,ρ(t)}],(8)

where Γ(t)+iΩ(t)=u˙(t)/u(t) and Γβ(t)=v˙(t)+2v(t)Γ(t). The function u(t) satisfies EquationEq. (3) and v(t)=0tdt10tdt2u(t1)μ(t1t2)u(t2), where μ(x)=0nˉ(ω)J(ω)eiωxdω with β=(kBT)1 and nˉ(ω)=(eβω1)1. According to the detailed balance condition, as long as Γ(t) and Γβ(t) are positive, the steady state of EquationEq. (8) must be

(9) ρ()=n=0[Γβ()/(2Γ())]n[1+Γβ()/(2Γ())]n+1|nn|,(9)

which defines a unique canonical state irrespective of the initial state

(10) ρconρ()=eβeffHˆs/TrseβeffHˆs,(βeff=1/kBTeff),(10)

with Teff=ω0kB{ln[1+2Γ()Γβ()]}1. The effective temperature Teff reduces to the environmental temperature T in the Markovian limit. This can be verified from the Markovian solution of EquationEq. (3), i.e. uMA(t)=e(Γ0+iΩ0)t with Γ0=πJ(ω0) and Ω0=ω0+PJ(ω)ω0ωdω, which leads to ΓMA(t)=κ, ΩMA(t)=Ω0, and ΓMAβ(t)=2κnˉ(ω0). We readily have Teff,MA=T. Thus, the canonical ensemble assumption in statistical mechanics can be dynamically confirmed under the Markovian approximation.

In the non-Markovian situation, the obtained Γ(t), Γβ(t), and Teff for the Ohmic spectral density are plotted in . In the absence of the bound state when ωc<10ω0, both Γ(t) and Γβ(t) tend to positive constant values asymptotically. Their positivity guarantees the validity of EquationEq. (9). Consequently, the system approaches a canonical state governed by EquationEq. (10), as shown in ), irrespective of in what state the system initially resides. Here it is qualitatively consistent with the one under the Markovian approximation. However, whenever the bound state is formed in the regime ωc10ω0 of , Γ(t) and Γβ(t) possess transient negative values and approach zero asymptotically. Their vanishing values cause the ill definition of the canonical state (10). In ), we really see that Teff changes to be divergent when the parameters tend to the critical point forming the bound state. Another interesting observation is that even when the equilibration to a canonical state is valid in the absence of the bound state, Teff still shows a dramatically quantitative difference from T. ) indicates that the equilibrium state matching with EquationEq. (9) is independent of the initial condition when the bound state is absent. However, this is not true as long as the bound state is formed [see )]. This result confirms again the breakdown of canonical thermalization due to the formation of the bound state.

Building a unified framework to treat the issues in equilibrium and nonequilibrium statistical mechanics, the result gives a microscopic description of the canonical thermalization and reveals the working condition of the assumption of ergodicity as the basis of the equilibrium statistical mechanics.

2.4. Quantum metrology

Quantum metrology employs quantum resources, such as entanglement and squeezing, to attain a measurement precision surpassing the limit achievable in classical physics [Citation5,Citation127–130]. It has been found that the metrology precision with a N-body entangled state reaches the Heisenberg limit (HL) scaling N1, which is far better than the shot-noise limit (SNL) N1/2 revealed by N-body product state [Citation5]. Such an enhancement inspires the application in gravitation wave detection [Citation131,Citation132] and atomic clocks [Citation133,Citation134]. The main obstacle of quantum metrology in practical applications is decoherence. It has been shown that the Markovian decoherence degrades the scaling from the HL back to the SNL [Citation8,Citation135]. Some studies revealed that the non-Markovian effect can retrieve quantum superiority to beat the SNL [Citation136–138]. However, it works only in the short encoding time condition. The precision gets worse in the long-time condition. Therefore, how to achieve metrology precision surpassing the SNL in the long encoding time condition in the presence of the dissipative environment is of significance in the practical application of quantum metrology.

2.4.1. Quantum metrology based on Mech-Zehnder interferometer

To estimate a frequency γ of a system, we prepare two modes of optical fields with frequency ω0 and the initial state |Ψin as the probe to input into the Mech-Zehnder interferometer (MZI). The input-output relation of the MZI reads |Ψout=VˆUˆ0(γ,t)Vˆ|Ψin, where Vˆ=exp[iπ4(bˆ1bˆ2+bˆ2bˆ1)] is the action of BSi and Uˆ0(γ,t)=exp(iHˆ0t/) with Hˆ0=ω0m=1,2bˆmbˆm+γbˆ2bˆ2 describes the free evolution of the probe and its coupling to the system for encoding. The photon difference Mˆ=bˆ1bˆ1bˆ2bˆ2 is measured by Di. If we use the squeezed state |Ψin=Dˆbˆ1Sˆbˆ2|0,0, where Dˆbˆ=eαbˆαbˆ with α=|α|eiφ, Sˆbˆ=e12(ξbˆ2ξbˆ2) with ξ=reiϕ, and |0,0 is the vacuum state, then we have the best precision of estimating γ via δγ=δM|Mˉ/γ| as [Citation139]

(11) minδγ=[(1β)e2r+β]12tN|12β|,(βsinh2r/N)(11)

when ϕ=2φ and γt=(2m+1)π/2 for mZ. If the squeezing is absent, then minδγ|β=0=(tN01/2)1 with N0=|α|2 is just the SNL. For β0, using e2r1/(4sinh2r) for N1 and optimizing β, we obtain minδγ|β=(2N)1=(tN3/4)1, which is called Zeno limit. It beats the SNL and manifests the superiority of the squeezing in metrology.

In practice, the above ideal encoding dynamics is obscured by the decoherence. The decoherence in MZI comes from the photon loss. Consider that the second optical field is affected by photon loss during the parameter encoding process, then its dynamics is governed by EquationEq. (2) with oˆ=bˆ2. We can calculate [Citation34]

Mˉ=Re[u(t)eiω0t](sinh2r|α|2),
δM2=Im[u(t)eiω0t]2[|αcoshrαsinhreiϕ|2+sinh2r]
(12) +Re[u(t)eiω0t]2[|α|2+12sinh2(2r)]+1|u(t)|22N,(12)

where u(t) satisfies EquationEq. (3) with the replacement of ω0 by ω0+γ. Applying the Markovian solution uMA(t)=e[κ+i(ω0+γ+Δ)]t with κ=πJ(ω0+γ) and Δ=P0J(ω)ω0+γωdω, we obtain minδγe2κt12Nt21/2 when β=(2N)1 and φ=2ϕ. Getting divergent in the long-time limit, its minimum at t=κ1 returns the SNL eκ(2N)1/2. Thus, the quantum superiority of the scheme in the Markovian noise disappears completely, which is consistent with the result in the Ramsey spectroscopy [Citation135].

Figure 4. (a) Scheme of Mach-Zehnder interference. Metrology precision as a function of time t (b) and total photon number (c) in different η of the Ohmic spectral density. The inset of (c) shows steady-state value Z=|c()| and the energy spectrum. Parameters are β=(2N)1, ωc=300ω0, γ=πω0, t=10ω01 for (c) and N=100 for (b). Reproduced figures from [Citation34]

Figure 4. (a) Scheme of Mach-Zehnder interference. Metrology precision as a function of time t (b) and total photon number (c) in different η of the Ohmic spectral density. The inset of (c) shows steady-state value Z=|c(∞)| and the energy spectrum. Parameters are β=(2N)−1, ωc=300ω0, γ=πω0, t=10ω0−1 for (c) and N=100 for (b). Reproduced figures from [Citation34]

In the non-Markovian case, focusing on the situation in the presence of the bound state, we can obtain the precision using the long-time solution ZeiEbt of EquationEq. (5)

(13) minδγ|β=(2N)1=(tN3/4)1Z[1+1Z22Z2N12]12,(13)

when t=(2m+1)π2|ω0ϖb| and φ=2ϕ. It is interesting to see that, different from the previous results [Citation8,Citation135–138], where δγ gets divergent with time, δγ decreases with the encoding time monotonically. This is numerically confirmed by the results in ). Thus, the bound state makes the superiority of the encoding time as a resource in the ideal metrology case retrieved. The relation of the precision with the total photon number in ) reveals that, with the formation of the bound state, not only the SNL can be surpassed, but also the ideal Zeno limit is asymptotically retrieved with Z approaching unit [see EquationEq. (13)].

2.4.2. Quantum metrology based on Ramsey spectroscopy

Another widely used metrology scheme is based on the atomic Ramsey spectroscopy [Citation135]. To estimate atomic frequency ω0, we use n two-level atoms prepared in |Ψin=(|gn+|en)/2 as the probe. It freely evolves to |Ψω0=(|gn+einω0t|en)/2. After performing a CNOT gate with the first atom as the controller and the others as the target to disentangle them and a Hadamard gate on the first atom, the measuring of the excited-state population Mˆ=σˆ+σˆ on the first atom results in Mˉ=sin2nω0t2 and ΔM=|sinnω0t|/2. Repeating the experiment in duration T, we have N=T/t results. According to the central limit theorem, we have δM=ΔMN. Then the propagation of error leads to the best error in estimating ω0 as

(14) δω0|ideal=(n2Tt)1/2,(14)

which is called HL. It has a n1/2 time enhancement to the SNL.

Same as the MZI, it is important to estimate the performance of Ramsey interferometry under decoherence. Consider that the atomic free evolution in the parameter encoding step is obscured by a dissipative environment. Thus, the dynamics of each atom is governed by EquationEq. (2) with oˆ=σˆ. The state after the encoding process reads

(15) ρ(t)=12{[|u(t)|2|ee|+(1|u(t)|2)|gg|]n+|gg|n+[u(t)n|eg|n+H.c.]}.(15)

Then repeating the same readout process as the ideal case, we can evaluate [Citation33]

(16) δω0=T[ω0Re(u(t)n)]2t[1Re2(u(t)n)]1/2.(16)

Using the Markovian solution uMA(t)=e[κ+i(ω0+Δ)]t, we obtain that δω0 tends to be divergent in the long-time limit. After optimizing t in short-time condition, we have the best precision min(δω0)=(nT/κe)1/2, which returns to the SNL. Thus, the quantum superiority of the scheme in the Markovian environment disappears completely [Citation135].

In the non-Markovian dynamics, by substituting the form ZeiEbt of large-time u(t) in the presence of the bound state into EquationEq. (16), we have

(17) min(δω0)=Z(n+1)(n2Tt)1/2,(17)

where the dependence of Eb on ω0 has been considered via ω0Eb=Z. Once again, the bound state causes the encoding time as a metrology resource recovered. The precision asymptotically approaches the ideal HL (14) for n<<1/lnZ when Z reaches unity.

In summary, we find that whether the superiority of the quantum metrology under dissipation exists or not is highly depends on the formation of the bound state and the non-Markovian effect. When the decoherence is Markovian or the bound state is absent, the quantum superiority is destroyed; whereas when the decoherence is non-Markovian and the bound states are formed, the quantum superiority is retrieved. The result suggests a guideline to experimentation to implement the ultrasensitive measurement in the practical noise situation by engineering the formation of the bound state.

2.5. Quantum plasmonics

The hybrid system composed of the metal-dielectric interface and quantum emitters (QEs) has attracted much attention due to its application in modern science, ranging from physics and chemistry to materials science [Citation24,Citation140–143]. The light is confined in the mode of surface plasmon polariton (SPP), which is a hybrid mode of the electromagnetic field and charge density wave called surface plasmon near the metal-dielectric interface [Citation23,Citation144–148]. The fascinating effect of SPP comes from its unique dispersion relation kSPP=ωcεm(ω)εdεm(ω)+εd, where εd is the relative permittivities of the dielectric. The metal has a permittivity denoted by a complex Drude model εm(ω)=εωp2/[ω(ω+iγp)], where ωp is the bulk plasma frequency, ε is the high-frequency limit of εm(ω), and γp is the Ohmic loss of light in the metal [Citation149,Citation150]. The dispersion relation shows that the SPP has a shorter wavelength than the light. It makes that the SPP can confine light within regions far below the diffraction limit. This effect endows SPP an idea platform to study sub-wavelength optics [Citation149,Citation151] and strong light–matter interactions [Citation152–158].

Figure 5. (a) Evolution and (b) long-time values of the excited-state population in different Δz. The purple line in (b) shows Z2 evaluated by the bound state. (c) Energy spectrum of the whole system in different Δz. (d) Decay rate of the QE in different Δz. The parameters are chosen as ω0=1.2 eV, γ0=104 eV, and εd=25. The inset of (a) shows the evolution when γ0=106 eV. Reproduced figures from [Citation25]

Figure 5. (a) Evolution and (b) long-time values of the excited-state population in different Δz. The purple line in (b) shows Z2 evaluated by the bound state. (c) Energy spectrum of the whole system in different Δz. (d) Decay rate of the QE in different Δz. The parameters are chosen as ℏω0=1.2 eV, ℏγ0=10−4 eV, and εd=25. The inset of (a) shows the evolution when ℏγ0=10−6 eV. Reproduced figures from [Citation25]

The quantized QE-SPP interaction is studied in Ref. [Citation159] based on the macroscopic quantum electrodynamics. In the vicinity of the nanoparticle-dielectric interface, the spectral density of the SPPs shows a Lorentzian lineshape [Citation160]. Thus, the SPPs is described by an artificially damping cavity. Based on this assumption, a lot of studies on QE-metal nanoparticle hybrid system have been given [Citation161–165]. However, the Lorentzian spectral density approximation only correctly captures the short-time dynamics of the QEs. On the other hand, due to absorption of the metal to the SPPs, QE coupled to the SPPs tends to its ground state in a long-time limit. It is the main obstacle in the practical application of a hybrid QE-SSP system. Therefore, how to stabilize the coherence of QE is an urgent problem. Recent studies have shown that the formation of the QE-SPP bound states could supply an efficient way to suppress the decoherence of the QE and the damping of the SPPs in the metal [Citation25,Citation26].

Consider a QE with frequency ω0 placed at a distance Δz above the silver interface. The Hamiltonian follows EquationEq. (1) with the k-th SPP mode–emitter interaction strength gk=ic2ωk2πε0/d3rIm[εm(ωk)]μG(r,r,ωk)ek. Here ε0 being the vacuum permittivity, c the speed of light, G(r,r,ω) is the Green’s function satisfying the Helmholtz equation [××ω2εm(ω)/c2]G(r,r,ω)=δ(rr), r is the position of the emitter with an electric dipole μ, and ek is the polarization direction of the SPP mode. Taking the dipole μ along the z direction, the spectral density of the SPPs reads [Citation25]

(18) J(ω)=3γ0εdω34πω03Re[0dss3(1rpe2ikzdΔz)1s2],(18)

where s=kρ/kd, rp=εdkzmεm(ω)kzdεdkzm+εm(ω)kzd with kzm and kzd being the axial components of the wave vector in the metal and the dielectric, respectively, and γ0=ω03μ/(3πε0c3) is the vacuum spontaneous emission rate. The evolution of the QE obeys EquationEq. (2). shows the evolution of the excited-state population Pe(t)=Tr[σˆ+σˆρ(t)] of the QE in different Δz. When the intrinsic decay rate γ0 is small, Pe(t) monotonically decays to zero, especially for large Δz [see the inset of )]. When γ0 is large, the dynamical oscillation becomes significant. It represents a rapid excitation exchange and thus manifests the strong QE-SPP interactions. A remarkable difference between the two results is that although Pe(t) approaches zero when Δz=2.0 nm, while it tends to a stable nonzero value when Δz=1.2 nm. It is in sharp contrast to the previous result and represents a dissipation suppression of the QE. ) shows the steady-state population Pe() in different Δz. It reveals that Pe() matches well with Z2 obtained by the bound-state analysis. ) shows clearly that the regions where the QE population is preserved correspond exactly to the one where a bound state is formed in the environmental bandgap. ) is the decay rate γ(t) of the QE. The fast oscillations manifest the rapid energy exchange between the QE and the SPPs. It is interesting to see that γ(t) asymptotically approaches zero with the formation of the bound state. Such behavior implies that the plasmonic nanostructure in the presence of the bound state can be an ideal platform for the coherent manipulation of QE dynamics, even it is inherently dissipative. This is helpful for utilizing plasmonic nanostructures in quantum devices, such as entanglement generators [Citation166].

The formation of the bound state in the QE-SPPs system also indicates that the SPPs can be used to realize entanglement preservation [Citation26].

3. Quantum control in periodically driven systems

Coherent control via periodic driving dubbed as Floquet engineering has become a versatile tool in artificially synthesizing exotic states of matter in systems of ultracold atoms [Citation167,Citation168], photonics [Citation169,Citation170], superconductor qubits [Citation171], and graphene [Citation172]. Different from the static systems, the systems under periodic driving has no well-defined energy spectrum because its energy is not conserved. Floquet theory [Citation173,Citation174] supplies us with a powerful approach to map a nonequilibrium system under periodic driving to a static one. According to Floquet theory, the periodic system with Hˆ(t+T)=Hˆ(t) has a complete set of basis |ϕα(t) determined by

(19) Hˆ(t)it]|ϕα(t)=α|ϕα(t)(19)

such that the evolution of any state can be expanded as |ψ(t)=αCαeiαt|ϕα(t) with Cα=uα(0)|Ψ(0). The time-independence of Cα implies that α and |ϕα(t) play the same roles in a periodic system as eigenenergies and stationary states do in a static system. They are thus called quasienergies and quasi-stationary states, respectively. Carrying all the quasi-stationary-state characters, the quasienergy spectrum formed by all α is a key to study the periodic system. Note that α is periodic with a period 2π/T because eilωt|uα(t) with ω=2π/T is also the eigenstate of EquationEq. (19) with eigenvalue α+lω. It can be proven that the Floquet eigen Equationequation (19) is equivalent to

(20) UˆT|ϕα(0)=eiαT|ϕα(0)(20)

with UˆT the one-period evolution operator. Thus, the Floquet equation defines an effective static system Hˆeff=iTlnUˆT whose eigenvalues are the quasienergies. Then one can engineer the quasienergy spectrum by changing the parameters of the external periodic driving and artificially synthesize many exotic states of matter absent in the static systems. Its merits reside in that the controllability of the systems is dramatically enhanced because time as a new control dimension is introduced to the systems by the periodic driving. This greatly lifts the experimental difficulty in realizing different types of quantum states of matter via changing the intrinsic parameters of the systems, which is hard to be manipulated once the material of the system is fabricated.

In the above section, we reviewed the decoherence control via engineering the bound state in the energy spectrum. A practical difficulty is that one generally cannot change the energy spectrum once the system is explicitly given. In this section, we will review that a parallel decoherence control can be realized by periodic driving as long as a Floquet bound state (FBS) is formed in the quasienergy spectrum. Actually, such engineering to the FBS in the quasienergy spectrum can be generalized to realize artificial topological phases. We will also review its recent progress.

3.1. Floquet control of quantum dissipation

Figure 6. (a) Evolution of Pt and (b) quasienergy spectrum of the whole system in different driving amplitude a2. The parameters T=0.25πJ1, a1=0, τ=0.1πJ1, g=1.0J, λ=20.0J, and L=800 are used. Evolution of Pt for the initial state |ϕ=|0 in (c) and Ft for the initial state |ϕ=(|0+|0)/2 in (d) when a2=36.0J with the FBS (cyan solid line) and a2=1.5J without the FBS (red dashed line). The blue dotdashed lines show the results obtained via analytically evaluating the contribution of the FBS to the asymptotic state. The parameters in (c, d) are T=0.05πJ1 and τ=0.02πJ1. Reproduced figures from [Citation39]

Figure 6. (a) Evolution of Pt and (b) quasienergy spectrum of the whole system in different driving amplitude a2. The parameters T=0.25πJ−1, a1=0, τ=0.1πJ−1, g=1.0J, λ=20.0J, and L=800 are used. Evolution of Pt for the initial state |ϕ⟩=|↑0⟩ in (c) and Ft for the initial state |ϕ⟩=(|↑0⟩+|↓0⟩)/2 in (d) when a2=36.0J with the FBS (cyan solid line) and a2=1.5J without the FBS (red dashed line). The blue dotdashed lines show the results obtained via analytically evaluating the contribution of the FBS to the asymptotic state. The parameters in (c, d) are T=0.05πJ−1 and τ=0.02πJ−1. Reproduced figures from [Citation39]

The idea of controlling quantum dissipation via periodic driving is based on the correspondence between the quasienergy spectrum in a temporally periodic system and the energy spectrum in the static system [Citation39]. Compared with the static case, the periodic driving provides a much flexible tool in manipulating the formation of the bound state.

Consider a periodically driven spin-1/2 system interacting with an XX-coupled spin chain environment. The Hamiltonian reads Hˆ(t)=HˆS(t)+HˆI+HˆE with

HˆS(t)=12[λ+A(t)]σˆ0z,HˆI=g2(σˆ0xσˆ1x+σˆ0yσˆ1y),
(21) HˆE=λ2j=1Lσˆjz+J2j=1L1(σˆjxσˆj+1x+σˆjyσˆj+1y).(21)

Here σˆjα (α=x,y,z) being the Pauli matrix of the j-th spin, λ is the Zeeman interaction induced by the longitudinal magnetic field, J and g are the coupling strengths between the nearest-neighbor spins in the spin chain and the coupling between system spin and the first spin of the chain, respectively, and L is the total number of spins in XX-chain. Here the periodic driving field reads

(22) A(t)=a1,nT<tnT+τa2,nT+τ<t(n+1)T,(22)

where T is the driving period, τ is the duration of the first piece, and a1,2 are the driving amplitudes. We consider the initial state as |Ψ(0)=|0|{j}. Because the total excitation number Nˆ=j=0Lσˆj+σˆj is conserved, the time-evolved state can be expanded as |Ψ(t)=j=0LeiLλt2cj(t)σˆj+|{j}. Here the excited-state population Pt|c0(t)|2 of the system spin is used to characterize the decoherence effect.

) shows the evolution of Pt in different driving amplitude a2. When the driving is switched off, i.e., a2=0, Pt decays monotonically to zero, which means a complete dissipation exerted by the spin chain to the system spin. When the driving is switched on, it is interesting to see that Pt is stabilized repeatedly with the increase of a2. It is remarkable to see that the regimes where the dissipation is inhibited match well with the ones where the FBS is formed in the quasienergy spectrum [see )]. To understand the behavior, we, according to Floquet theory, rewrite |Ψ(t) as

(23) |Ψ(t)=eiLλt2[xeiFBSt|ϕFBS(t)+αBandyαeiαt|ϕα(t)],(23)

where x=ϕFBS(0)|Ψ(0) and yα=ϕα(0)|Ψ(0). One can see that Pt evolves asymptotically to Px2|Ψ(0)|uFBS(t)|2 with all the components in the quasi-energy band vanishing due to the out-of-phase interference contributed by the continuous phases, as confirmed in ). In the absence of the FBS, Pt decays to zero finally. It means that the presence of the FBS would cause Pt to survive in the only component of the FBS and thus synchronize with the driving field [Citation175]. For a general initial state |ϕ(0)=12(|0+|0), we use the initial-state fidelity Ft=ϕ(t)|TrE[|Ψ(t)Ψ(t)|]|ϕ(0) to reflect the damping of the quantum coherence. The dynamics are shown in ). In the absence of the FBS, Ft decays to a trivial value 1/2, which indicates that the system ρs=TrE[|Ψ(t)Ψ(t)|]=I2 has no coherence, while Ft shows persistent oscillation once the FBS is formed. The long-time dynamics in ) verifies again the coincidence of the FBS analysis (blue dashed line) with the exact numerical results (cyan line).

As a stationary state, the bound state in both the static system and the periodically driven system makes the system survive in the only component of the bound state in the long-time limit. Thus, the coherence contained in the bound state is preserved forever. Different from the bound state in a static system, the FBS formed in the quasienergy spectrum itself is temporally periodic. Such periodic driving induced dissipation suppression can be seen as a time-dependent analog to the decoherence suppression induced by the structured environment in spatially periodic photonic crystal setting [Citation176], where the spatial periodicity causes the environment to present a bandgap structure and thus the decoherence of the system is suppressed when the system frequency resides in the bandgap region. However, the temporally periodic driving relaxes greatly the experimental difficulties of the static system in fabricating specific spatial periodicity. Such an FBS mechanism is an exact non-Markovian dynamical control to suppress quantum dissipation by using periodic driving. An optical simulation of this mechanism has been proposed in Ref [Citation40].

3.2. Floquet control of dissipative quantum battery

A quantum battery (QB) is a device that stores and converts energy at the quantum level [Citation177]. It holds the promises of higher energy-storing density in large-scale integration and faster-charging power than its classical counterpart. A minimal model of QB is described by two coupled two-level systems with one acting as the charger and another as the QB [Citation178,Citation179]. The Hamiltonian reads

(24) Hˆ0(t)=l=b,cωlσˆlσˆl+κf(t)(σˆbσˆc+H.c.),(24)

where σˆl=|gle|, with |g and |e being the ground and excited states of the two-level systems, are the transition operators of the QB and charger with frequency ωl, and κ is their coupling strength. The time-dependent f(t) is a control parameter for the charging, storing, and discharging process that takes

(25) f(t)=1,nT<tτc+nT0,nT+τc<tnT+(τs+τc)1,nT+(τs+τc)<t(n+1)TnN.(25)

Here τc, τs, and τd are the time durations of charging, storing, and discharging, respectively. The initial state of the QB-charger setup is set to |ψ(0)=|gb,ec. The average energy E(t)=ωbψ(t)|σˆbσˆb|ψ(t) of the QB is used to characterize its power, where |ψ(t)Tei0tHˆ0(τ)dτ|ψ(0) is the evolved state of the QB-charger setup. For the resonant case ωb=ωc, with the optimal control time, i.e. τc=τs=τd=2πκ [Citation42], the QB-charger setup shows the perfect cyclic evolution: E(t) reaches its maximal value at the end of each charging step, and empties its total energy at the end of each discharging step. However, in the realistic condition, an invertible power loss is present due to the interactions of the QB and the charger with their respective dissipative environment. It was found that the dissipation causes that energy of QB drop to zero after finite charing-storing-discharging cycles, which is called the aging of the QB [Citation180,Citation181]. The temporally periodical nature of the cyclic process of the QB inspires us that the detrimental effect of the dissipation on the QB-charger setup can be inhibited by engineering the formation of the FBSs. Then, we review how to reactive the QB under the influences of the dissipative environments by manipulating the FBSs [Citation42].

Figure 7. (a) Evolution of the energy of the QB and (b) quasienergy spectrum of the total system in different coupling strength κ. (c) Evolution of E(t) in long-time limit and the quasienergy difference Δ0 of two FBSs in different κ. The energy in storage time duration is highlighted by red segments in (c). Other parameters are N=30, ωb=ωc=2ϖ, g=0.5ϖ, q=0.5ϖ and τc=τs=τd=π/(2κ). Reproduced figures from [Citation42]

Figure 7. (a) Evolution of the energy of the QB and (b) quasienergy spectrum of the total system in different coupling strength κ. (c) Evolution of E(t) in long-time limit and the quasienergy difference Δ∈0 of two FBSs in different κ. The energy in storage time duration is highlighted by red segments in (c). Other parameters are N=30, ωb=ωc=2ϖ, g=0.5ϖ, q=0.5ϖ and τc=τs=τd=π/(2κ). Reproduced figures from [Citation42]

Consider that the QB and the charger are coupled to two independent environments consisting of an N×N nearest-neighbor coupled bosonic mode on the square lattice with the mode frequency ϖ and hopping rate q [Citation182]. Their couplings follow EquationEq. (1) with oˆσˆb,(c), ωkϖ2q(coskx+cosky), and the system-environment coupling strength being gk=gN. , b) shows the evolution of E(t) in different coupling strength κ and the corresponding quasienergy spectrum. We can see that the regions where different numbers of FBSs are formed match well with the ones where E(t) shows different behaviors, i.e. complete decay, energy trapping, and persistent oscillation. When no FBS is formed, E(t) asymptotically decays to zero. However, when one or two FBSs are formed, E(t) approaches finite value or shows a persistent oscillation in the long-time limit. According to the Floquet theory, such anomalous dynamical behaviors can be understood by decomposing the evolution state of the total system as |Ψ(t)=j=1Mcjei0jt|ϕ0j(t)+βCBdβeiβt|ϕβ(t), where and |ϕ(t) satisfy EquationEq. (19), M is the number the FBSs, cjϕ0j(0)|Ψ(0) and dβϕβ(0)|Ψ(0). In the long-time limit, the contribution from the continuous energy band to E(t) approaches zero due to the out-of-phase interference. Then, E() reads

(26) E()ω0=jj =1Mcjcj ei(0j0j )tϕ0j (t)|σˆbσˆb|ϕ0j(t),(26)

which indicates that the steady behavior of E() strongly dependent on the number of FBSs, as shown in , b). In particular, when M=2, E() shows the persistent oscillation with multiple frequencies jointly determined by the driven frequency ωT and the quasienergy difference of two FBSs Δ0|0101|. With the bound-state mechanism at hands, it is shown that the QB influenced by the environment can be reactivated by controlling the energy of two FBSs. As shown in ), with the increasing of κ, the quasienergy difference asymptotically tends to zero, which reduces the undesired energy oscillation contributed from the interference between two FBSs and thus the QB returns to its near-ideal cyclic stage.

The combination between the decoherence analysis of QB and the Floquet theory supplies an efficient way to overcome the aging problem of the QB via the versatile Floquet engineering. It opens an avenue to build an aging-free QB in practice.

3.3. Floquet control of topological phases

The distinguished role of the periodic driving in engineering bound states in the quasienergy spectrum can also be used to artificially synthesize exotic topological phases [Citation47–53]. Parallel to the topological phases in the static system, the topological phases in periodically driven systems are called Floquet topological phases. Floquet topological phases defined in the quasienergy spectrum are characterized by a full insulating gap in the bulk and gapless bound states at the edges or the surfaces [Citation183]. The versatility of periodic driving schemes enables us not only to realize topological phases not accessible for the static system in the same setting, but also to explore novel topological phases totally absent in the static system.

Consider a two-band topological system with its parameters periodically driven by a pairwise function with the Hamiltonians read H1 and H2 in the respective time durations T1 and T2. The coefficient matrices of the Hamiltonians in the operator basis of the momentum space take the form Hj(k)=hj(k)σ (j=1,2) with σ being the Pauli matrices. The one-periodic evolution operator reads UT=εI2×2+irσ which yield the effective Hamiltonian Heff(k):=iTlnUT=heff(k)σ with heff(k)=arccos(ε)r_/T and [Citation48]

ε=cos|T1h1(k)|cos|T2h2(k)|h_1h_2sin|T1h1(k)|sin|T2h2(k)|,
r=h_1×h_2sin|T1h1(k)|sin|T2h2(k)|h_2cos|T1h1(k)|sin|T2h2(k)|
(27) h_1cos|T2h2(k)|sin|T1h1(k)|.(27)

Here T=T1+T2 and v_v/|v| is the unit vector of v. The topological phase transition is associated with the closing and reopening of the quasi-energy bands. One can find from EquationEqs. (27) that the band gap closing occurs at ε=±1, which is obtained when

(28) Tj|hj(k)|=njπ,njZ,(28)
(29) orh_1h_2=±1T1|h1(k)|±T2|h2(k)|=nπ,nZ(29)

at the quasienergy zero (or π/T) if n is even (or odd). They give the sufficient condition for judging the occurrence of the Floquet topological phase transition.

3.3.1. Floquet Majorana modes

Figure 8. (a) Phase diagram characterized by the winding number when μ=10|Δ1| and |Δ2|=2.5|Δ1|. (b) Quasienergy spectrum for t1=|Δ1|. The dashed blue line and solid black line stand for two degenerate pairs and a single pair of MMs, respectively. Reproduced figures from [Citation47]

Figure 8. (a) Phase diagram characterized by the winding number when μ=−10|Δ1| and |Δ2|=2.5|Δ1|. (b) Quasienergy spectrum for t1=|Δ1|. The dashed blue line and solid black line stand for two degenerate pairs and a single pair of MMs, respectively. Reproduced figures from [Citation47]

Realizing Majorana modes (MMs) in condensed-matter systems are of vast experimental and theoretical interests due to their potential applications in quantum computation [Citation184–188]. The zero-bias conductance peaks are regarded as a signature of MMs in spin-orbit coupled semiconductor nanowires [Citation189]. However, they can also be induced by other reasons, e.g. the strong disorder in the nanowire [Citation190,Citation191] and smooth confinement potential at the wire end [Citation191]. Therefore, more ways to double confirm the formation of MMs in these systems are desired. Studies show that periodic driving can create a widely tunable number of Floquet MMs, which, on the one hand, could dramatically enhance the experimental signal strength, on the other hand, could supply a novel way to identify whether the signal originates from MFs or others by observing its response to the tuning of the driving coefficients [Citation47].

Consider the Kitaev model Hamiltonian for a 1D spinless p-wave superconductor

H=μl=1Nclcll=1N1(t1clcl+1+Δ1clcl+1+h.c.)l=1N2(t2clcl+2
(30) +Δ2clcl+2+h.c.),(30)

where μ is the chemical potential, ta and Δa=Δaeiϕa with a=1 (a=2) describe the nearest- (next-nearest-)-neighbor hopping amplitude and pairing potential, respectively, and ϕa is the associated superconducting phases. Its topological characters are dependent on the relative phase ϕ=ϕ1ϕ2 [Citation192]. For ϕ=0 and π, H has time-reversal and particle-hole symmetries and its topology is characterized by the winding number, where two pairs of Majorana modes can be generated at most. For other values of ϕ, H has no time-reversal and particle-hole symmetries and its topology is characterized by a Z2 topological invariant, where one pair of Majorana modes is achievable at most.

The periodic driving protocol is designed as follows: in the first half period, H1=H(ϕ1,ϕ2) with both ϕ1 and ϕ2 fixed; whereas in the second half period, we swap ϕ1 and ϕ2 so that H2=H(ϕ2,ϕ1). It has been proved that the distinguished action of the periodic driving is that it successfully recovers the time-reversal symmetry and simulates the long-range hopping of the electrons. Thanks to these abilities, an arbitrarily number of MMs is created. ) shows the phase diagram for given driving parameters with the corresponding quasienergy spectrum shown in ). Compared to the static case which supports at most two pairs of MMs, Floquet engineering provides a feasible method to realize tunable topological edge states. It was reported that more MMs can be induced by increasing the driving period T, where longer-range hopping is simulated by the periodic driving [Citation47].

3.3.2. Floquet Chern insulators

The large-Chern-number topological phases have attracted considerable interests [Citation193–199]. Providing more edge channels, these phases are expected to improve certain device performance by lowering the contact resistance [Citation193,Citation195]. They can also support multi-mode one-way waveguides, a feature relevant to realizing reflectionless waveguides splitters, combiners, and one-way photonic circuits in photonic crystals [Citation194,Citation199]. However, a band structure with large Chern numbers is normally hard to be generated in natural material. Special long-range interactions and multilayered structure of the materials are needed to support such types of exotic phases [Citation196,Citation200,Citation201].

Periodic driving widely used in graphene [Citation172], optical lattices [Citation167], and photonic crystals [Citation170] supplies a useful way to create the large-Chern-number topological phases [Citation48]. Consider the Haldane model with up to the third-neighbor hopping included. In the operator basis Ck=(cak,cbk)T, its static Hamiltonian can be written as [. The associated Bloch vectors h(k) are

(31) hx(k)=t1[1+cos(ka1)+cos(ka2)]+t3{2cos[k(a1a2)]+cos[k(a1+a2)]},hy(k)=t1[sin(ka1)+sin(ka2)]+t3sin[k(a1+a2)],hz(k)=2t2sinϕ{sin(ka1)sin(ka2)sin[k(a1a2)]}+M,(31)

where tj are the jth--neighbor hopping amplitude, ϕ is the phase gained by t2 hoppings, and M is the mass term. The primitive vectors are a1=(3/2,3/2) and a2=(3/2,3/2) in the unit of the hexagonal lattice constant [Citation202]. shows the phase diagram in the T1 and T2 plane by periodically switching t3=0.75t1 and ϕ=π/6 for duration T1 to t3=0.75t1 and ϕ=π/2 for duration T2. A widely tunable Chern numbers with the equal numbers of edge states are generated [see )]. Due to the high controllability of the cold atom system [Citation203,Citation204], a periodic driving scheme was proposed to realize large-topological-number phases via periodically changing the optical lattices [Citation49].

Figure 9. (a) Phase diagram in the T1-T2 plane obtained by calculating Chern numbers. The phase-boundary lines are analytically obtained from EquationEqs. (28) and (Equation29). Quasienergy spectra of the periodically Haldane model with C=4 when (T1,T2)=(0.9,1.2) in (b) and with C=7 when (T1,T2)=(1.3,1.2)] in (c). Parameters t2=0.8t1 and M=0 are used. Reproduced figures from [Citation48]

Figure 9. (a) Phase diagram in the T1-T2 plane obtained by calculating Chern numbers. The phase-boundary lines are analytically obtained from EquationEqs. (28)(28) Tj|hj(k)|=njπ,nj∈Z,(28) and (Equation29(29) orh_1⋅h_2=±1T1|h1(k)|±T2|h2(k)|=nπ,n∈Z(29) ). Quasienergy spectra of the periodically Haldane model with C=−4 when (T1,T2)=(0.9,1.2) in (b) and with C=−7 when (T1,T2)=(1.3,1.2)] in (c). Parameters t2=0.8t1 and M=0 are used. Reproduced figures from [Citation48]

3.3.3. Non-Hermitian Floquet topological phases

Inspired by the recent experimental realization of topological insulators and the rapid development of non-Hermitian photonics, non-Hermitian topological phases have attracted much attention. From the fundamental physics aspect, people desire to know can the well-developed topological theory in Hermitian systems be generalized to the non-Hermitian cases. This is nontrivial because the skin effect of the non-Hermitian system could cause the breakdown of the bulk-boundary correspondence, which lays the foundation in the topological description in Hermitian systems. From the application aspect, non-Hermitian topological phases have exhibited significant advantages in diverse applications, e.g. laser, invisible media, and sensing [Citation205]. In order to prompt the application of non-Hermitian topological phases, the creation of novel non-Hermitian topological phases in a controllable way via Floquet engineering the system by periodic driving is highly promising. However, the interplay between the non-Hermitian skin effect and periodic driving makes it very hard to characterize the non-Hermitian Floquet topological phases. A recent study has established a complete description of the non-Hermitian Floquet topological phases [Citation53].

Consider the non-Hermitian Su-Schrieffer-Heeger model

(32) H=l=1L[(t1+γ2)albl+(t1γ2)blal+t2(albl1+h.c.)],(32)

where al (bl) are the annihilation operators on the sublattice A (B) of the lth- lattice, and L is lattice length. In momentum space and the operator basis (a˜k,b˜k)T with a˜k (b˜k) being the Fourier transform of al (bl), it reads H(k)=dxσx+(dy+iγ/2)σy, where dx=t1+t2cosk and dy=t2sink. The bands close at k=π (or 0) when t1=t2±γ/2 (or t2±γ/2). It is in conflict with the result under the open-boundary condition, where the bands close when t1=t22+γ2/4. It is called the non-Hermiticity caused breakdown of bulk-boundary correspondence [Citation206,Citation207]. The problem for the static system with chiral symmetry σz1H(k)σz=H(k) is cured via replacing eik by β=|t1γ/2t1+γ/2|eik. Then the Hamiltonian is converted into H(β)=n=±Rn(β)σn with σ±=(σx±iσy)/2 and R±(β)=t1±γ2+β1t2. Here β defines a generalized Brillouin zone (GBZ). Its topological property is described by the winding number W=(W+W)/2, where W±=12π[argR±(β)]Cβ with [argR±(β)]Cβ are the phase change of R± as β counterclockwisely goes along the GBZ Cβ [Citation208]. When |t1|<t22+γ2/4, W=1 and a pair of edge states is formed.

If the periodic driving

(33) t2(t)=f,tmT,mT+T1)qf,tmT+T1,(m+1)T),mZ,(33)

is applied to the system, then the chiral symmetry is broken, which makes the GBZ method well developed for the chiral symmetric system not directly applicable. To resolve this problem, two similarity transformations Gj=ei(1)jHjTj/2 are introduced to covert the one-period evolution operator UT into U˜T,1=eiH1T1/2eiH2T2eiH1T1/2 and U˜T,2=eiH2T2/2eiH1T1eiH2T2/2, from which two effective Hamiltonian H˜eff,j=iTlnU˜T,j can be defined. It can be analytically proved that the chiral symmetry is recovered in H˜eff,j. Then after introducing the GBZ β, two winding numbers Wj associated with H˜eff,j can be calculated. The numbers of zero- and π/T-mode edge sates relate to Wj as N0=|W1+W2|/2 and Nπ/T=|W1W2|/2 [Citation209]. ) shows the quasienergy spectrum under the open-boundary conditions. It is different from the quasienergy spectrum under the periodic-boundary condition, which is called the breakdown of bulk-boundary condition. Although qualitatively capturing the exceptional points of the quasienergy under the periodic-boundary condition, the ill-defined topological numbers from the conventional BZ nonphysically take half integers [see the dashed lines in , c)]. However, the ones from the GBZ correctly count the number of the edge states obtained in the open-boundary condition [see the solid lines in , c)]. , e) show the phase diagram in the T1-T2 plane. A widely tunable number of non-Hermitian edge states are induced by changing the driving parameters.

Figure 10. (a) Quasienergy spectra with the change of the driving amplitude under the open (blue lines) and periodic (gray lines) boundary conditions. Numbers of 0-mode (b) and π/T-mode (c) edge states defined in the conventional (red dashed) and generalized (cyan solid) BZ. Parameters t1=2.0γ, T1=T2=0.6γ1, q=3.0, and L=80 are used in (a), (b), and (c). Phase diagram characterized by W1 (d) and W2 (e). The phase boundary lines are obtained from EquationEqs. (28) and (Equation29). Parameters t1=1.5γ, f=2γ, and q=0 are used in (d) and (e). Reproduced figures from [Citation53]

Figure 10. (a) Quasienergy spectra with the change of the driving amplitude under the open (blue lines) and periodic (gray lines) boundary conditions. Numbers of 0-mode (b) and π/T-mode (c) edge states defined in the conventional (red dashed) and generalized (cyan solid) BZ. Parameters t1=2.0γ, T1=T2=0.6γ−1, q=3.0, and L=80 are used in (a), (b), and (c). Phase diagram characterized by W1 (d) and W2 (e). The phase boundary lines are obtained from EquationEqs. (28)(28) Tj|hj(k)|=njπ,nj∈Z,(28) and (Equation29(29) orh_1⋅h_2=±1T1|h1(k)|±T2|h2(k)|=nπ,n∈Z(29) ). Parameters t1=1.5γ, f=2γ, and q=0 are used in (d) and (e). Reproduced figures from [Citation53]

All the above results reveal that, without changing the intrinsic parameters in the static system, the periodic driving supplies us another control dimension to adjust the topological edge states. This is useful in the application of topological physics.

4. Conclusion and outlook

In summary, we have reviewed the recent progress on quantum control in open and periodically driven systems. Particular attention has been paid to the distinguished role of bound states in suppressing decoherence of open systems and realizing exotic topological phases in periodically driven systems. It has been revealed that the decoherence of the open system connects tightly with the feature of the energy spectrum of the total system consisting of the open system and its environment. With the formation of a bound state, the decoherence of the open system would be suppressed. It has a profound impact on entanglement preservation, quantum speedup, noncanonical thermalization, quantum metrology, and quantum plasmonics. A parallel result can be established in a periodically driven open system, where the formation of FBSs in the quasienergy spectrum can suppress the dissipation of the open system and the aging of the dissipative quantum battery. It is interesting to see that the controllability of periodic driving on engineering the FBSs can be applied to realize diverse exotic topological phases, e.g. the widely tunable numbers of Majorana modes, Chern insulators, and non-Hermitian topological insulators. The merit of such types of topological phases induced by periodic driving resides in that their controllability is dramatically enhanced because time as a novel control dimension is introduced to the system by periodic driving. It might inspire the exploration of their application via tuning the number of edge/surface states by the periodic driving. The bound state and its distinguished role in the static open-system dynamics have been observed in circuit QED [Citation210,Citation211] and ultracold atom systems [Citation212]. These progresses provide strong support to the bound-state mechanism and indicate that the profound consequences of the bound reviewed in this article is realizable in the current state of the art. Such experimental feasibility paves the way to beat the decoherence in diverse protocols in quantum technology.

The studies presented in this article hopefully open a number of research lines that deserve to be explored in the future. First, the combination of quantum optics and topological physics has emerged a field called topological quantum optics [Citation213–216], where the environments interacting with the quantum emitters have topologically nontrivial energy bands. It has potential application in a quantum network. Interesting questions include: Can the system-environment bound states be formed and how the interplay of these bound states with the intrinsic edge states of the environment affects the dynamics of the quantum emitters? Second, parallel to the bound states in the bandgap, the bound states in the continuum are also widely studies [Citation217,Citation218]. A recent study has showed that the bound states in the continuum can be used to realize the higher-order topological phases [Citation219]. Can they be efficietly manipulated by periodic driving such that more colorful higher-order topological phases are generated in a controllable manner? Third, time crystal is a novel state of matter showing the spontaneous symmetry breaking in the time domain of a quantum system under periodic driving [Citation220,Citation221]. It arises from the competition of many-body localization due to the disorder against Floquet thermalization under periodic driving [Citation222,Citation223]. Inspired by the dominated role of bound states in noncanonical thermalization in static systems, a natural question is: Does the avoidance of Floquet thermalization in order to form time crystal is due to the presence of FBSs in quasienergy spectrum? The answer to this question might give the intrinsic physical mechanism to the presence of time crystal from the quasienergy-spectrum feature.

Acknowledgments

The work is supported by the National Natural Science Foundation of China (Grant Nos. 11875150 and 11834005).

Disclosure statement

No potential conflict of interest was reported by the authors.

Additional information

Funding

This work was supported by the National Natural Science Foundation of China [11834005,11875150].

References

  • Feynman RP. Simulating physics with computers. Int J Theor Phys. 1982 Jun;21:467–40.
  • Ekert A, Jozsa R. Quantum computation and shor’s factoring algorithm. Rev Mod Phys. 1996 Jul;68:733–753.
  • Law J. Quantum computation and quantum information. SIGSOFT Softw Eng Notes. 2001 July;26:91.
  • DiVincenzo DP. The physical implementation of quantum computation. Fortschritte der Physik. 2000;48:771–783.
  • Giovannetti V, Lloyd S, Maccone L. Quantum-enhanced measurements: beating the standard quantum limit. Science. 2004;306:1330–1336.
  • Leibfried D, Barrett MD, Schaetz T, et al. Toward Heisenberg-limited spectroscopy with multiparticle entangled states. Science. 2004;304:1476–1478.
  • Giovannetti V, Lloyd S, Maccone L. Advances in quantum metrology. Nat Photonics. 2011;5:222–229.
  • Escher BM, de Matos Filho RL, Davidovich L. General framework for estimating the ultimate precision limit in noisy quantum-enhanced metrology. Nat Phys. 2011;7:406–411.
  • Georgescu IM, Ashhab S, Nori F. Quantum simulation. Rev. Mod. Phys. 2014 Mar;86:153–185.
  • Bennett CH, Wiesner SJ. Communication via one- and two-particle operators on Einstein-Podolsky-Rosen states. Phys. Rev. Lett. 1992 Nov;69:2881–2884.
  • Mattle K, Weinfurter H, Kwiat PG, et al. Dense coding in experimental quantum communication. Phys. Rev. Lett. 1996 Jun;76:4656–4659.
  • Zurek WH. Decoherence, einselection, and the quantum origins of the classical. Rev. Mod. Phys. 2003 May;75:715–775.
  • Bennett CH, Bernstein HJ, Popescu S, et al. Concentrating partial entanglement by local operations. Phys Rev A. 1996 Apr;53:2046–2052.
  • Zhao Z, Pan J-W, Zhan MS. Practical scheme for entanglement concentration. Phys Rev A. 2001 Jun;64:14301.
  • Viola L, Lloyd S. Dynamical suppression of decoherence in two-state quantum systems. Phys Rev A. 1998 Oct;58:2733–2744.
  • de Lange G, Wang ZH, Ristè D, et al. Universal dynamical decoupling of a single solid-state spin from a spin bath. Science. 2010;330:60–63.
  • Lloyd S. Coherent quantum feedback. Phys Rev A. 2000 Jul;62:22108.
  • Tombesi P, Vitali D. Macroscopic coherence via quantum feedback. Phys Rev A. 1995 Jun;51:4913–4917.
  • John S, Wang J. Quantum electrodynamics near a photonic band gap: photon bound states and dressed atoms. Phys. Rev. Lett. 1990 May;64:2418–2421.
  • Ogawa S, Imada M, Yoshimoto S, et al. Control of light emission by 3d photonic crystals. Science. 2004;305:227–229.
  • Fujita M, Takahashi S, Tanaka Y, et al. Simultaneous inhibition and redistribution of spontaneous light emission in photonic crystals. Science. 2005;308:1296–1298.
  • Jorgensen MR, Galusha JW, Bartl MH. Strongly modified spontaneous emission rates in diamond-structured photonic crystals. Phys Rev Lett. 2011 Sep;107:143902.
  • Chang DE, Sørensen AS, Hemmer PR, et al. Quantum optics with surface plasmons. Phys Rev Lett. 2006;97:1–4.
  • Tame MS, McEnery KR, Özdemir K, et al. Quantum plasmonics. Nat Phys. 2013;9:329–340.
  • Yang CJ, AnJ-H. Suppressed dissipation of a quantum emitter coupled to surface plasmon polaritons. Phys Rev B. 2017;95:5–9.
  • Yang C-J, An J-H, Lin H-Q. Signatures of quantized coupling between quantum emitters and localized surface plasmons. Phys Rev Res. 2019;1:23027.
  • Miyamoto M. Bound-state eigenenergy outside and inside the continuum for unstable multilevel systems. Phys Rev A. 2005 Dec;72:63405.
  • Tong Q-J, An J-H, Luo H-G, et al. Mechanism of entanglement preservation. Phys Rev A. 2010 May;81:52330.
  • Lü Y-Q, An J-H, Chen X-M, et al. Frozen gaussian quantum discord in photonic crystal cavity array system. Phys Rev A. 2013 Jul;88:12129.
  • Yang WL, An J-H, Zhang C, et al. Preservation of quantum correlation between separated nitrogen-vacancy centers embedded in photonic-crystal cavities. Phys Rev A. 2013 Feb;87:22312.
  • Yang C-J, An J-H, Luo H-G, et al. Canonical versus noncanonical equilibration dynamics of open quantum systems. Phys Rev E. 2014 Aug;90:22122.
  • Liu HB, Yang WL, AnJ-H, et al. Mechanism for quantum speedup in open quantum systems. Phys Rev A. 2016 Feb;93:20105.
  • Wang Y-S, Chen C,An J-H. Quantum metrology in local dissipative environments. New J Phys. 2017 nov;19:113019.
  • Bai K, Peng Z, Luo H-G, et al. Retrieving ideal precision in noisy quantum optical metrology. Phys Rev Lett. 2019 Jul;123:40402.
  • Wei W, Bai S-Y,An J-H, Non-markovian sensing to a quantum reservoir. arXiv:2005.08553.
  • Poyatos JF, Cirac JI, Zoller P. Quantum reservoir engineering with laser cooled trapped ions. Phys. Rev. Lett. 1996 Dec;77:4728–4731.
  • Metelmann A, Clerk AA. Nonreciprocal photon transmission and amplification via reservoir engineering. Phys Rev X. 2015 Jun;5:21025.
  • Oka T, Kitamura S. Floquet engineering of quantum materials. Ann Rev Condens Matter Phys. 2019;10:387–408.
  • Chen C, An J-H, Luo H-G, et al. Floquet control of quantum dissipation in spin chains. Phys Rev A. 2015 May;91:52122.
  • Chao M, Wang Y-S,An J-H. Floquet engineering of localized propagation of light in a waveguide array. Phys Rev A. 2018 Feb;97:23808.
  • Yang WL, Song WL, An J-H, et al. Floquet engineering to entanglement protection of distant nitrogen vacancy centers. New J Phys. 2019 jan;21:013007.
  • Bai S-Y,An J-H, Floquet engineering to reactivate a dissipative quantum battery. Phys Rev A. 2020 Dec;102:060201(R).
  • Grifoni M, Peter H. Driven quantum tunneling. Phys Rep. 1998;304:229–354.
  • Thorwart M, Hartmann L, Goychuk I, et al. Controlling decoherence of a two-level atom in a lossy cavity. null. 2000 November;47:2905–2919.
  • Kohler S, Lehmann J, Peter H. Driven quantum transport on the nanoscale. Phys Rep. 2005;406:379–443.
  • Lindner NH, Refael G, Galitski V. Floquet topological insulator in semiconductor quantum wells. Nat Phys. 2011;7:490–495.
  • Tong Q-J, An J-H, Gong J, et al. Generating many majorana modes via periodic driving: a superconductor model. Phys Rev B. 2013 May;87:201109.
  • Xiong T-S, Gong J,An J-H. Towards large-chern-number topological phases by periodic quenching. Phys Rev B. 2016 May;93:184306.
  • Liu H, Xiong T-S, Zhang W, et al. Floquet engineering of exotic topological phases in systems of cold atoms. Phys Rev A. 2019 Aug;100:23622.
  • Rudner MS, Lindner NH, Berg E, et al. Anomalous edge states and the bulk-edge correspondence for periodically driven two-dimensional systems. Phys Rev X. 2013 Jul;3:31005.
  • Arijit Kundu HAF, Seradjeh B. Effective theory of floquet topological transitions. Phys Rev Lett. 2014 Dec;113:236803.
  • Lababidi M, Satija II, Zhao E. Counter-propagating edge modes and topological phases of a kicked quantum hall system. Phys Rev Lett. 2014 Jan;112:26805.
  • Hong W,An J-H. Floquet topological phases of non-hermitian systems. Phys Rev B. 2020 Jul;102:41119.
  • Cayssol J, Dóra B, Simon F, et al. Floquet topological insulators. Phys Status Solidi (RRL) Rapid Res Lett. 2013;7:101–108.
  • Mukherjee S, Spracklen A, Valiente M, et al. Experimental observation of anomalous topological edge modes in a slowly driven photonic lattice. Nat Commun. 2017;8:13918.
  • Scully MO, Zubairy MS. Quantum optics. New York: Cambridge University Press; 1997.
  • Khaetskii AV, Loss D, Glazman L. Electron spin decoherence in quantum dots due to interaction with nuclei. Phys Rev Lett. 2002 Apr;88:186802.
  • Merkulov IA, Efros AL, Rosen M. Electron spin relaxation by nuclei in semiconductor quantum dots. Phys Rev B. 2002 Apr;65:205309.
  • Bouchiat V, Vion D, Joyez P, et al. Quantum coherence with a single cooper pair. Phys Scr. 1998;T76:165.
  • Senitzky IR. Dissipation in quantum mechanics. the harmonic oscillator. Phys. Rev. 1960 Jul;119:670–679.
  • Senitzky IR. Dissipation in quantum mechanics. the harmonic oscillator. ii. Phys. Rev. 1961 Nov;124:642–648.
  • de Vega I, Alonso D. Dynamics of non-markovian open quantum systems. Rev Mod Phys. 2017 Jan;89:015001.
  • Kossakowski A. On quantum statistical mechanics of non-hamiltonian systems. Rep Math Phys. 1972;3:247–274.
  • Gardiner C and Zoller P, Quantum Noise (Springer-Verlarg, Berlin, 2004): a handbook of markovian and non-markovian quantum stochastic methods with applications to quantum optics. 2nd enlarged ed. Springer series in synergetics. Springer; 2004.
  • Ackerhalt JR, Knight PL, Eberly JH. Radiation reaction and radiative frequency shifts. Phys. Rev. Lett. 1973 Mar;30:456–460.
  • Ithier G, Collin E, Joyez P, et al. Decoherence in a superconducting quantum bit circuit. Phys Rev B. 2005 Oct;72:134519.
  • Breuer H-P, Laine E-M, Piilo J, et al. Colloquium: non-markovian dynamics in open quantum systems. Rev Mod Phys. 2016 Apr;88:21002.
  • Feynman RP, Vernon FL. The theory of a general quantum system interacting with a linear dissipative system. Ann Phys. 1963;24:118–173.
  • An J-H, Zhang W-M. Non-markovian entanglement dynamics of noisy continuous-variable quantum channels. Phys Rev A. 2007 Oct;76:42127.
  • Leggett AJ, Chakravarty S, Dorsey AT, et al. Dynamics of the dissipative two-state system. Rev. Mod. Phys. 1987 Jan;59:1–85.
  • Einstein A, Podolsky B, Rosen N. Can quantum-mechanical description of physical reality be considered complete? Phys. Rev. 1935 May;47:777–780.
  • Horodecki R, Horodecki P, Horodecki M, et al. Quantum entanglement. Rev Mod Phys. 2009 Jun;81:865–942.
  • Bennett CH, Brassard G. Quantum cryptography: public key distribution and coin tossing. Theor Comput Sci. 2014;560:7–11. Theoretical Aspects of Quantum Cryptography: celebrating 30 years of BB84.
  • Ekert AK. Quantum cryptography based on bell’s theorem. Phys. Rev. Lett. 1991 Aug;67:661–663.
  • Gisin N, Ribordy G, Tittel W, et al. Quantum cryptography. Rev. Mod. Phys. 2002 Mar;74:145–195.
  • Ting Y, Eberly JH. Finite-time disentanglement via spontaneous emission. Phys Rev Lett. 2004 Sep;93:140404.
  • Almeida M P, de Melo F, Hor-Meyll M O, et al. Experimental observation of environment-induced sudden death of entanglement. Proc. SPIE 6603, Noise and Fluctuations in Photonics, Quantum Optics, and Communications, 660318 (8 June 2007).
  • Ting Y, Eberly JH. Sudden death of entanglement. Science. 2009;323:598–601.
  • Bellomo B, Lo Franco R, Compagno G. Entanglement dynamics of two independent qubits in environments with and without memory. Phys Rev A. 2008 Mar;77:32342.
  • Maniscalco S, Francica F, Zaffino RL, et al. Protecting entanglement via the quantum zeno effect. Phys Rev Lett. 2008 Mar;100:90503.
  • Xu J-S, Li C-F, Gong M, et al. Experimental demonstration of photonic entanglement collapse and revival. Phys Rev Lett. 2010 Mar;104:100502.
  • Bellomo B, Franco RL, Maniscalco S, et al. Entanglement trapping in structured environments. Phys Rev A. 2008 Dec;78:60302.
  • Tong Q-J, An J-H, Luo H-G, et al. Decoherence suppression of a dissipative qubit by the non-markovian effect. J Phys B: Atomic Mol Opt Phys. 2010 jul;43:155501.
  • Wootters WK. Entanglement of formation of an arbitrary state of two qubits. Phys. Rev. Lett. 1998 Mar;80:2245–2248.
  • Hildebrand R. Concurrence revisited. J Math Phys. 2007;48:102108.
  • An J-H, Yeo Y, Zhang W-M, et al. Entanglement oscillation and survival induced by non-markovian decoherence dynamics of the entangled squeezed state. J Phys A. 2008 nov;42:015302.
  • Anandan J, Aharonov Y. Geometry of quantum evolution. Phys. Rev. Lett. 1990 Oct;65:1697–1700.
  • Mandelstam L., Tamm I. (1991) The Uncertainty Relation Between Energy and Time in Non-relativistic Quantum Mechanics. In: Bolotovskii B.M., Frenkel V.Y., Peierls R. (eds) Selected Papers. Springer, Berlin, Heidelberg.
  • Vaidman L. Minimum time for the evolution to an orthogonal quantum state. Am J Phys. 1992;60:182–183.
  • Lloyd S. Ultimate physical limits to computation. Nature. 2000;406:1047–1054.
  • Luo S. How fast can a quantum state evolve into a target state? Phys D. 2004;189:1–7.
  • Margolus N, Levitin LB. The maximum speed of dynamical evolution. Phys D. 1998;120:188–195. Proceedings of the Fourth Workshop on Physics and Consumption.
  • Taddei MM, Escher BM, Davidovich L, et al. Quantum speed limit for physical processes. Phys Rev Lett. 2013 Jan;110:50402.
  • Del Campo A, Egusquiza IL, Plenio MB, et al. Quantum speed limits in open system dynamics. Phys Rev Lett. 2013 Jan;110:50403.
  • Pires DP, Marco Cianciaruso LC, Céleri GA, et al. Generalized geometric quantum speed limits. Phys Rev X. 2016 Jun;6:21031.
  • Deffner S, Lutz E. Generalized clausius inequality for nonequilibrium quantum processes. Phys Rev Lett. 2010 Oct;105:170402.
  • Alipour S, Mehboudi M, Rezakhani AT. Quantum metrology in open systems: dissipative cramér-rao bound. Phys Rev Lett. 2014 Mar;112:120405.
  • Miroslav Gajdacz KK, Das JA, Sherson JF, et al. Time-limited optimal dynamics beyond the quantum speed limit. Phys Rev A. 2015 Dec;92:62106.
  • Tran MC, Chen C-F, Ehrenberg A, et al. Hierarchy of linear light cones with long-range interactions. Phys Rev X. 2020 Jul;10:31009.
  • Yung M-H. Quantum speed limit for perfect state transfer in one dimension. Phys Rev A. 2006 Sep;74:30303.
  • Deffner S, Lutz E. Quantum speed limit for non-markovian dynamics. Phys Rev Lett. 2013 Jul;111:10402.
  • Cimmarusti AD, Yan Z, Patterson BD, et al. Environment-assisted speed-up of the field evolution in cavity quantum electrodynamics. Phys Rev Lett. 2015 Jun;114:233602.
  • Breuer H-P, Laine E-M, Piilo J. Measure for the degree of non-markovian behavior of quantum processes in open systems. Phys Rev Lett. 2009 Nov;103:210401.
  • Popescu S, Short AJ, Winter A. Entanglement and the foundations of statistical mechanics. Nat Phys. 2006;2:754–758.
  • Linden N, Popescu S, Short AJ, et al. Quantum mechanical evolution towards thermal equilibrium. Phys Rev E. 2009 Jun;79:61103.
  • Tasaki H. From quantum dynamics to the canonical distribution: general picture and a rigorous example. Phys. Rev. Lett. 1998 Feb;80:1373–1376.
  • Sheldon Goldstein JL, Lebowitz RT, Zangh N. Canonical typicality. Phys Rev Lett. 2006 Feb;96:50403.
  • Reimann P. Canonical thermalization. New J Phys. 2010 may;12:55027.
  • Lychkovskiy O. Necessary condition for the thermalization of a quantum system coupled to a quantum bath. Phys Rev E. 2010 Jul;82:11123.
  • Lee CK, Cao J, Gong J. Noncanonical statistics of a spin-boson model: theory and exact monte carlo simulations. Phys Rev E. 2012 Aug;86:21109.
  • Genway S, Ho AF, Lee DKK. Dynamics of thermalization and decoherence of a nanoscale system. Phys Rev Lett. 2013 Sep;111:130408.
  • Rigol M, Dunjko V, Olshanii M. Thermalization and its mechanism for generic isolated quantum systems. Nature. 2008;452:854–858.
  • Ponomarev AV, Denisov S, Hänggi P. Thermal equilibration between two quantum systems. Phys Rev Lett. 2011 Jan;106:10405.
  • Bañuls MC, Cirac JI, Hastings MB. Strong and weak thermalization of infinite nonintegrable quantum systems. Phys Rev Lett. 2011 Feb;106:50405.
  • Polkovnikov A, Sengupta K, Silva A, et al. Colloquium Nonequilibrium dynamics of closed interacting quantum systems. Rev. Mod. Phys. 2011 Aug;83:863–883.
  • Geva E, Rosenman E, Tannor D. On the second-order corrections to the quantum canonical equilibrium density matrix. J Chem Phys. 2000;113:1380–1390.
  • Mori T, Miyashita S. Dynamics of the density matrix in contact with a thermal bath and the quantum master equation. J Phys Soc Jpn. 2008;77:124005.
  • Subasi Y, Fleming CH, Taylor JM, et al. Equilibrium states of open quantum systems in the strong coupling regime. Phys Rev E. 2012 Dec;86:61132.
  • Pagel D, Alvermann A, Fehske H. Equilibration and thermalization of the dissipative quantum harmonic oscillator in a nonthermal environment. Phys Rev E. 2013 Jan;87:12127.
  • Rançon A, Bonart J. Exact dynamics and thermalization of an open bosonic quantum system in the presence of a quantum phase transition induced by the environment. Europhys Lett. 2013 dec;104:50010.
  • Hoeppe U, Wolff C, Küchenmeister J, et al. Direct observation of non-markovian radiation dynamics in 3d bulk photonic crystals. Phys Rev Lett. 2012 Jan;108:43603.
  • Tahara H, Ogawa Y, Minami F. Non-markovian dynamics of spectral narrowing for excitons in the layered semiconductor gase observed using optical four-wave mixing spectroscopy. Phys Rev Lett. 2011 Jul;107:37402.
  • Galland C, Högele A, Türeci HE, et al. Non-markovian decoherence of localized nanotube excitons by acoustic phonons. Phys Rev Lett. 2008 Aug;101:67402.
  • Liu B-H, Li L, Huang Y-F, et al. Experimental control of the transition from markovian to non-markovian dynamics of open quantum systems. Nat Phys. 2011 December;7:931–934.
  • Madsen KH, Ates S, Lund-Hansen T, et al. Observation of non-markovian dynamics of a single quantum dot in a micropillar cavity. Phys Rev Lett. 2011 Jun;106:233601.
  • Liu H-B, An J-H, Chen C, et al. Anomalous decoherence in a dissipative two-level system. Phys Rev A. 2013 May;87:52139.
  • Giovannetti V, Lloyd S, Maccone L. Quantum metrology. Phys Rev Lett. 2006 Jan;96:10401.
  • Braunstein SL, Caves CM. Statistical distance and the geometry of quantum states. Phys. Rev. Lett. 1994 May;72:3439–3443.
  • Degen CL, Reinhard F, Cappellaro P. Quantum sensing. Rev Mod Phys. 2017 Jul;89:35002.
  • Pezzè L, Augusto Smerzi MK, Oberthaler RS, et al. Quantum metrology with nonclassical states of atomic ensembles. Rev Mod Phys. 2018 Sep;90:35005.
  • Tse M, Yu H, Kijbunchoo N, et al. Quantum-enhanced advanced ligo detectors in the era of gravitational-wave astronomy. Physical Review Letters. 2019 Dec;123:231107.
  • Acernese et al. (Virgo Collaboration). Increasing the astrophysical reach of the advanced virgo detector via the application of squeezed vacuum states of light. Phys Rev Lett. 2019;1230:231108.
  • Andrew DL, Boyd MM, Ye J, et al.Optical atomic clocks. Rev Mod Phys. 2015 Jun;87:637–701.
  • Kruse I, Lange K, Peise J, et al. Improvement of an atomic clock using squeezed vacuum. Phys Rev Lett. 2016 Sep;117:143004.
  • Huelga SF, Macchiavello C, Pellizzari T, et al. Improvement of frequency standards with quantum entanglement. Phys Rev Lett. 1997 Nov;79:3865–3868.
  • Matsuzaki Y, Benjamin SC, Fitzsimons J. Magnetic field sensing beyond the standard quantum limit under the effect of decoherence. Phys Rev A. 2011 Jul;84:12103.
  • Chin AW, Huelga SF, Plenio MB. Quantum metrology in non-markovian environments. Phys Rev Lett. 2012 Dec;109:233601.
  • Macieszczak K. Zeno limit in frequency estimation with non-markovian environments. Phys Rev A. 2015 Jul;92:10102.
  • Caves CM. Quantum-mechanical noise in an interferometer. Phys Rev D. 1981 Apr;23:1693–1708.
  • Kabashin AV, Evans P, Pastkovsky S, et al. Plasmonic nanorod metamaterials for biosensing. Nat Mater. 2009;8:867–871.
  • Atwater HA, Polman A. Plasmonics for improved photovoltaic devices. Nat Mater. 2010;9:205–213.
  • Giannini V, Fernández-Domínguez AI, Heck SC, et al. Plasmonic nanoantennas: fundamentals and their use in controlling the radiative properties of nanoemitters. Chem Rev. 2011 June;111:3888–3912.
  • Lee YH, Shi W, Lee HK, et al. Nanoscale surface chemistry directs the tunable assembly of silver octahedra into three two-dimensional plasmonic superlattices. Nat Commun. 2015;6:6990.
  • Pitarke JM, Silkin VM, Chulkov EV, et al. Theory of surface plasmons and surface-plasmon polaritons. Rep Prog Phys. 2006 dec;70:1–87.
  • Cacciola A, Omar Di S, Stassi R, et al. Ultrastrong coupling of plasmons and excitons in a nanoshell. ACS Nano. 2014 November;8:11483–11492.
  • Aberra Guebrou S, Symonds C, Homeyer E, et al. Coherent emission from a disordered organic semiconductor induced by strong coupling with surface plasmons. Phys Rev Lett. 2012 Feb;108:66401.
  • Törmä P, Barnes WL. Strong coupling between surface plasmon polaritons and emitters: a review. Rep Prog Phys. 2014 dec;78:013901.
  • González-Tudela A, Huidobro PA, Martn-Moreno L, et al. Theory of strong coupling between quantum emitters and propagating surface plasmons. Phys Rev Lett. 2013 Mar;110:126801.
  • Barnes WL. Surface plasmon–polariton length scales: a route to sub-wavelength optics. J Opt A:Pure Appl Opt. 2006 mar;8:S87–S93.
  • Aspnes D E, in Handbook of Optical Constants of Solids, edited by Palik E, (Academic, Boston, 1985).
  • Barnes WL, Dereux A, Ebbesen TW. Surface plasmon subwavelength optics. Nature. 2003;424:824–830.
  • Trügler A, Hohenester U. Strong coupling between a metallic nanoparticle and a single molecule. Phys Rev B. 2008 Mar;77:115403.
  • Wersäll M, Jorge Cuadra TJ, Antosiewicz SB, et al. Observation of mode splitting in photoluminescence of individual plasmonic nanoparticles strongly coupled to molecular excitons. Nano Lett. 2017 January;17:551–558.
  • Baranov DG, Wersäll M, Cuadra J, et al. Novel nanostructures and materials for strong light-matter interactions. ACS Photonics. 2018 January;5:24–42.
  • Chikkaraddy R, de Nijs B, Felix Benz SJ, et al. Single-molecule strong coupling at room temperature in plasmonic nanocavities. Nature. 2016;535:127–130.
  • Matsuzaki K, Vassant S, Liu H-W, et al. Strong plasmonic enhancement of biexciton emission: controlled coupling of a single quantum dot to a gold nanocone antenna. Sci Rep. 2017;7:42307.
  • Kewes G, Binkowski F, Burger S, et al. Heuristic modeling of strong coupling in plasmonic resonators. ACS Photonics. 2018 Oct;5:4089–4097.
  • Vasa P, Lienau C. Strong light-matter interaction in quantum emitter/metal hybrid nanostructures. ACS Photonics. 2018 January;5:2–23.
  • Gruner T, Welsch D-G. Green-function approach to the radiation-field quantization for homogeneous and inhomogeneous kramers-kronig dielectrics. Phys Rev A. 1996 Mar;53:1818–1829.
  • Waks E, Sridharan D. Cavity qed treatment of interactions between a metal nanoparticle and a dipole emitter. Phys Rev A. 2010 Oct;82:043845.
  • Shah RA, Scherer NF, Pelton M, et al. Ultrafast reversal of a fano resonance in a plasmon-exciton system. Phys Rev B. 2013 Aug;88:075411.
  • Delga A, Feist J, Bravo-Abad J, et al. Quantum emitters near a metal nanoparticle: strong coupling and quenching. Phys Rev Lett. 2014 Jun;112:253601.
  • Rong-Chun G, Hughes S. Quantum dynamics of two quantum dots coupled through localized plasmons: an intuitive and accurate quantum optics approach using quasinormal modes. Phys Rev B. 2015 Nov;92:205420.
  • Li R-Q, Hernángomez-Pérez D, Garca-Vidal FJ, et al. Transformation optics approach to plasmon-exciton strong coupling in nanocavities. Phys Rev Lett. 2016 Aug;117:107401.
  • Peng P, Liu YC, Xu D, et al. Enhancing coherent light-matter interactions through microcavity-engineered plasmonic resonances. Phys Rev Lett. 2017 Dec;119:233901.
  • Gonzalez-Tudela A, Martin-Cano D, Moreno E, et al. Entanglement of two qubits mediated by one-dimensional plasmonic waveguides. Phys Rev Lett. 2011 Jan;106:20501.
  • Eckardt A. Colloquium: atomic quantum gases in periodically driven optical lattices. Rev Mod Phys. 2017 Mar;89:11004.
  • Meinert F, Mark MJ, Lauber K, et al. Floquet engineering of correlated tunneling in the bose-hubbard model with ultracold atoms. Phys Rev Lett. 2016 May;116:205301.
  • Rechtsman MC, Zeuner JM, Plotnik Y, et al. Photonic floquet topological insulators. Nature. 2013;496:196–200.
  • Cheng Q, Pan Y, Wang H, et al. Observation of anomalous π modes in photonic floquet engineering. Phys Rev Lett. 2019 May;122:173901.
  • Roushan P, Neill C, Megrant A, et al. Chiral ground-state currents of interacting photons in a synthetic magnetic field. Nat Phys. 2017;13:146–151.
  • McIver JW, Schulte B, Stein F-U, et al. Light-induced anomalous hall effect in graphene. Nat Phys. 2020;16:38–41.
  • Shirley JH. Solution of the schrödinger equation with a hamiltonian periodic in time. Phys Rev. 1965 May;138:B979–B987.
  • Sambe H. Steady states and quasienergies of a quantum-mechanical system in an oscillating field. Phys Rev A. 1973 Jun;7:2203–2213.
  • Russomanno A, Silva A, Santoro GE. Periodic steady regime and interference in a periodically driven quantum system. Phys Rev Lett. 2012 Dec;109:257201.
  • Leistikow MD, Mosk AP, Yeganegi E, et al. Inhibited spontaneous emission of quantum dots observed in a 3d photonic band gap. Phys Rev Lett. 2011 Nov;107:193903.
  • Alicki R, Fannes M. Entanglement boost for extractable work from ensembles of quantum batteries. Phys Rev E. 2013 Apr;87:42123.
  • Ferraro D, Campisi M, Andolina GM, et al. High-power collective charging of a solid-state quantum battery. Phys Rev Lett. 2018 Mar;120:117702.
  • Andolina GM, Keck M, Mari A, et al. Extractable work, the role of correlations, and asymptotic freedom in quantum batteries. Phys Rev Lett. 2019 Feb;122:47702.
  • Pirmoradian F, Klaus M. Aging of a quantum battery. Phys Rev A. 2019 Oct;100:43833.
  • Kamin FH, Tabesh FT, Salimi S, et al. Non-markovian effects on charging and self-discharging process of quantum batteries. New J Phys. 2020 aug;22:083007.
  • González-Tudela A, Cirac JI. Markovian and non-markovian dynamics of quantum emitters coupled to two-dimensional structured reservoirs. Phys Rev A. 2017 Oct;96:43811.
  • Kitagawa T, Berg E, Rudner M, et al. Topological characterization of periodically driven quantum systems. Phys Rev B. 2010 Dec;82:235114.
  • Liang F, Kane CL. Superconducting proximity effect and majorana fermions at the surface of a topological insulator. Phys Rev Lett. 2008 Mar;100:096407.
  • Sau JD, Lutchyn RM, Tewari S, et al. Generic new platform for topological quantum computation using semiconductor heterostructures. Phys Rev Lett. 2010 Jan;104:40502.
  • Sato M, Takahashi Y, Fujimoto S. Non-abelian topological order in s-wave superfluids of ultracold fermionic atoms. Phys Rev Lett. 2009 Jul;103:20401.
  • Gong M, Chen G, Jia S, et al. Searching for majorana fermions in 2d spin-orbit coupled fermi superfluids at finite temperature. Phys Rev Lett. 2012 Sep;109:105302.
  • Shi-Liang Zhu L-B, Shao ZD, Wang L-MD. Probing non-abelian statistics of majorana fermions in ultracold atomic superfluid. Phys Rev Lett. 2011 Mar;106:100404.
  • Williams JR, Bestwick AJ, Gallagher P, et al. Unconventional josephson effect in hybrid superconductor-topological insulator devices. Phys Rev Lett. 2012 Jul;109:56803.
  • Liu J, Potter AC, Law KT, et al. Zero-bias peaks in the tunneling conductance of spin-orbit-coupled superconducting wires with and without majorana end-states. Phys Rev Lett. 2012 Dec;109:267002.
  • Kells G, Meidan D, Brouwer PW. Near-zero-energy end states in topologically trivial spin-orbit coupled superconducting nanowires with a smooth confinement. Phys Rev B. 2012 Sep;86:100503.
  • Ryu S, Schnyder AP, Furusaki A, et al. Topological insulators and superconductors: tenfold way and dimensional hierarchy. New J Phys. 2010 jun;12:065010.
  • Wang J, Lian B, Zhang H, et al. Quantum anomalous hall effect with higher plateaus. Phys Rev Lett. 2013 Sep;111:136801.
  • Skirlo SA, Ling L, Marin S. Multimode one-way waveguides of large chern numbers. Phys Rev Lett. 2014 Sep;113:113904.
  • Fang C, Gilbert MJ, Andrei Bernevig B. Large-chern-number quantum anomalous hall effect in thin-film topological crystalline insulators. Phys Rev Lett. 2014 Jan;112:046801.
  • Möller G, Cooper NR. Fractional chern insulators in harper-hofstadter bands with higher chern number. Phys Rev Lett. 2015 Sep;115:126401.
  • Scaffidi T, Simon SH. Large chern number and edge currents in sr2ruo4. Phys Rev Lett. 2015 Aug;115:87003.
  • Röntynen J, Ojanen T. Topological superconductivity and high chern numbers in 2d ferromagnetic shiba lattices. Phys Rev Lett. 2015 Jun;114:236803.
  • Skirlo SA, Ling L, Igarashi Y, et al. Experimental observation of large chern numbers in photonic crystals. Phys Rev Lett. 2015 Dec;115:253901.
  • Jiang H, Qiao Z, Liu H, et al. Quantum anomalous hall effect with tunable chern number in magnetic topological insulator film. Phys Rev B. 2012 Jan;85:45445.
  • Yang S, Zheng-Cheng G, Sun K, et al. Topological flat band models with arbitrary chern numbers. Phys Rev B. 2012 Dec;86:241112.
  • Haldane FDM. Model for a quantum hall effect without landau levels: condensed-matter realization of the “parity anomaly”. Phys Rev Lett. 1988 Oct;61:2015–2018.
  • Jünemann J, Piga A, Ran S-J, et al. Exploring interacting topological insulators with ultracold atoms: the synthetic creutz-hubbard model. Phys Rev X. 2017 Sep;7:31057.
  • Potirniche I-D, Potter AC, Schleier-Smith M, et al. Floquet symmetry-protected topological phases in cold-atom systems. Phys Rev Lett. 2017 Sep;119:123601.
  • Ashida Y, Gong Z, Ueda M. Non-hermitian physics, 2020.
  • Lee TE. Anomalous edge state in a non-hermitian lattice. Phys Rev Lett. 2016 Apr;116:133903.
  • Yao S, Wang Z. Edge states and topological invariants of non-hermitian systems. Phys Rev Lett. 2018 Aug;121:86803.
  • Yokomizo K, Murakami S. Non-bloch band theory of non-hermitian systems. Phys Rev Lett. 2019 Aug;123:66404.
  • Asbóth JK, Tarasinski B, Delplace P. Chiral symmetry and bulk-boundary correspondence in periodically driven one-dimensional systems. Phys Rev B. 2014 Sep;90:125143.
  • Hood JD, Goban A, Asenjo-Garcia A, et al. Atom–atom interactions around the band edge of a photonic crystal waveguide. Proc Nat Acad Sci2. 2016;113:10507–10512.
  • Liu Y, Houck AA. Quantum electrodynamics near a photonic bandgap. Nat Phys. 2017;13:48–52.
  • Krinner L, Stewart M, Pazmiño A, et al. Spontaneous emission of matter waves from a tunable open quantum system. Nature. 2018 July;559:589–592.
  • Perczel J, Borregaard J, Chang DE, et al. Topological quantum optics using atomlike emitter arrays coupled to photonic crystals. Phys Rev Lett. 2020 Feb;124:83603.
  • Barik S, Karasahin A, Flower C, et al. A topological quantum optics interface. Science. 2018;359:666–668.
  • Bello M, Platero G, Cirac JI, et al. Unconventional quantum optics in topological waveguide qed. Sci Adv. 2019;5:eaaw0297.
  • Garca-Elcano I, González-Tudela A, Bravo-Abad J. Tunable and robust long-range coherent interactions between quantum emitters mediated by weyl bound states. Phys Rev Lett. 2020 Oct;125:163602.
  • Plotnik Y, Peleg O, Dreisow F, et al. Experimental observation of optical bound states in the continuum. Phys Rev Lett. 2011 Oct;107:183901.
  • Calajó G, Fang Y-L-L, Baranger HU, et al. Exciting a bound state in the continuum through multiphoton scattering plus delayed quantum feedback. Phys Rev Lett. 2019 Feb;122:73601.
  • Cerjan A, Marius Jürgensen WA, Benalcazar SM, et al. Observation of a higher-order topological bound state in the continuum. Phys Rev Lett. 2020 Nov;125:213901.
  • Yao NY, Potter AC, Potirniche I-D, et al. Discrete time crystals: rigidity, criticality, and realizations. Phys Rev Lett. 2017 Jan;118:30401.
  • Choi S, Choi J, Landig R, et al. Observation of discrete time-crystalline order in a disordered dipolar many-body system. Nature. 2017;543:221–225.
  • Regnault N, Nandkishore R. Floquet thermalization: symmetries and random matrix ensembles. Phys Rev B. 2016 Mar;93:104203.
  • Haldar A, Moessner R, Das A. Onset of floquet thermalization. Phys Rev B. 2018 Jun;97:245122.