Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Quantum state engineering by periodical two-step modulation in an atomic system

Open Access Open Access

Abstract

By periodical two-step modulation, we demonstrate that the dynamics of a multilevel system can evolve even in a multiple large detunings regime and provide the effective Hamiltonian (of interest) for this system. We then illustrate this periodical modulation in quantum state engineering, including achieving direct transition from the ground state to the Rydberg state or the desired superposition of two Rydberg states without satisfying the two-photon resonance condition, switching between the Rydberg blockade regime and the Rydberg antiblockade regime, stimulating distinct atomic transitions by the same laser field, and implementing selective transitions in the same multilevel system. Particularly, it is robust against perturbation of control parameters. Another advantage is that the waveform of the laser field has a simple square-wave form, which is readily implemented in experiments. Thus, it offers us a novel method of quantum state engineering in quantum information processing.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Perfect coherent control of quantum state is a critical technology in quantum information processing [1, 2]. The traditional way usually adopts (near-) resonance driving fields. For instance, with regard to the transition from ground state to Rydberg state, two laser fields should satisfy so-called two-photon resonance condition with the help of intermediate state [3, 4]. One disadvantage of resonance condition is that system dynamics seriously suffers from the amplitude noise of driving fields [5]. Furthermore, the resonance condition in fact is rigorous from experimental viewpoint, since it requires to accurately match driving field frequency with transition frequency. This is also the essential reason why the same laser field cannot stimulate different atomic transitions. But if this unexplored problem is resolved, i.e., not requiring resonance condition in dynamics any longer, it can afford us a novel avenue to implement coherent manipulations by driving fields with arbitrary frequencies in the same quantum system.

On the other hand, there is increasingly interested in periodic driving systems [6–17], to a large extent, due to the fact that periodic driving fields provide the possibility of controlling and changing some physical properties of system, such as tunable spin-orbit coupling in ultracold-atom systems [18, 19] and the appearance of Floquet Majorana fermions [20–25]. For periodic driving systems, they possess well-defined Floquet eigenstates and corresponding eigenvalues (i.e., so-called quasienergies), which can be readily calculated in an extend Hilbert space [26, 27]. By modulating Floquet eigenstates and quasienergies, the dynamical behaviors of periodic driving systems would appear more interesting than that of static systems. Recently, the periodic driving field has been employed to study the Stückelberg interference phenomenon [28–30] and the dynamics of three-level system [31].

To our knowledge, there are several ways such as the chirped adiabatic passage [32–35] that can be employed to implement population transfer when two-photon resonance is not satisfied well. However, it is scarcely ever studied the transition of multilevel system in multiple large detunings (MLDs) regime (i.e., beyond two (multi)-photon resonance). One of the reasons may be that the analytical solutions of eigenvalues and eigenstates are hard to obtain in multilevel system. Thus, it is a very fresh attempt of applying periodic driving fields in MLDs regime. In this work, we demonstrate that quantum state can still evolve by periodical two-step modulation in MLDs regime, however large the detunings are. Interestingly, the system dynamics would be strikingly different when taking different time intervals of two-step modulation. Due to the existence of multiple single-photon detunings, some levels can be safely eliminated so that the system is effectively reduced to two-level model. Particularly, it takes full advantage of large detuning conditions, i.e., the system is robust against perturbation of control parameters and the dynamics is frozen again when removing periodical modulation [1]. We then illustrate its applications in quantum state engineering, including the direct transition from the ground state to the Rydberg state (or the desired superposition of two Rydberg states). The choice of control parameters is flexibility in Rydberg atom system, e.g., the laser frequencies and modulation periods can be arbitrary in principle. Other applications are to switch between Rydberg blockade regime [36–39] and Rydberg antiblockade regime [40–42], stimulate different atomic transitions by the same laser field, and implement selective transitions in the same multilevel system.

2. Two-step modulation in general three-level system

Consider the toy model that a general three-level atomic system interacts with two laser fields, where the level transition |k|k+1 (k = 1, 2) is coupled by the k-th laser field with coupling strength Ωk. In the lab frame, the system Hamiltonian reads (=1 hereafter)

H0=k=13ωk|kk|+k=12Ωkeiωklt|kk+1|+H.c.,

where ωk and ωkl are the frequency of level |k and the k-th laser field, respectively. For convenience, it is instructive to adopt interaction picture and the system Hamiltonian becomes

H0=k=12Δk|k+1k+1|+Ωk|kk+1|+H.c.,

where Δ1=ω2ω1ω1l, Δ2=ω3ω1ω1lω2l. One of the most attracting points in this work is that there is no restriction with detuning Δk (k=1,2), while the two-photon (near-) resonance condition (i.e., Δ20) should be always satisfied during evolution process in previous work. Without loss of generality, we assume the system is in MLDs regime, i.e., ΔkΩk, and there are no other restrictions on the values of Δk and Ωk (k=1,2).

To achieve the evolution operator of system, we need to know the system eigenstates and eigenvalues. In mathematic, by solving a cubic equation, one can analytically calculate the eigenstates and eigenvalues of Hamiltonian H0 in Eq. (2), then achieve the evolution operator subsequently. However, it is helpless for us in this work due to the intricate expressions of eigenstates and eigenvalues. In the following, we adopt alternative method to approximatively achieve its expressions. Before elaborating this method, it is instructive to adopt the matrix form of Hamiltonian H0 in the basis {|1,|2,|3}, which reads

H0=(0Ω10Ω1Δ1Ω20Ω2Δ2).

At first, by imposing the unitary transformation S=(1000sin αcos α0cos αsin α) on the system, the Hamiltonian becomes

0=S+H0S=(0Ω1sin αΩ1cos αΩ1sin αξ10Ω1cos α0ξ2),

where tan 2α=2Ω2Δ2Δ1 and ξ1,2=12(Δ1+Δ2(Δ1Δ2)2+4Ω22). Note that the Hilbert space is composed by the basis {|1,|ξ1,|ξ2} now, where |ξ1=sin α|2+cos α|3 and |ξ2=cos α|2sin α|3. When Ω1{ξ1,ξ2,|ξ1ξ2|}, according to the standard perturbation theory, the eigenvalues Ek and eigenstates |Ek (k=1,2,3) of 0 approximatively read

E1ξ1x12ξ2x22,  |E1|1x1|ξ1x2|ξ2,E2ξ1+ξ1x12,     |E2|ξ1+x1|1,E3ξ2+ξ2x22,     |E3|ξ2+x2|1.
where x1=Ω1sin αξ11, x2=Ω1cos αξ21, and the coefficient of eigenstates is not normalization. As a result, the evolution operator of system reads
U(t)=ei0t=(1x1(eiΘ11)x2(eiΘ21)x1(eiΘ11)eiΘ1x1x2x2(eiΘ21)x1x2eiΘ2),

where we have ignored the global phase and the higher-order terms (x12,x22), Θ1=(E2E1)t, and Θ2=(E3E1)t. Naturally, the expression of evolution operator in the basis {|1,|2,|3} reads

 U(t)=SU(t)S+ (1(eiΘ21)x1sinα+(eiΘ11)x2cosα(eiΘ21)x1cosα+(1eiΘ1)x2sinα(eiΘ21)x1sinα+(eiΘ11)x2cosαeiΘ2sin2α+eiΘ1cos2α(eiΘ2eiΘ1)cosαsinα(eiΘ21)x1cosα+(1eiΘ1)x2sinα(eiΘ2eiΘ1)cosαsinαeiΘ2cos2α+eiΘ1sin2α).

It is easily observed from Eq. (9) that the populations of each level are almost frozen under static Hamiltonian H0 in Eq. (2) in MLDs regime.

From the above derivation procedure, we can find that only employing static Hamiltonian H0 is utterly helpless for quantum state engineering. However, The situation is quite different when we adopt periodical two-step modulation of Hamiltonian, i.e.,

H(t)={H0=k=12Δk|k+1k+1|+Ωk|kk+1|+H.c., t[mT,mT+τ1),H0 '=k=12Δk '|k+1k+1|+Ωk '|kk+1|+H.c., t[mT+τ1,(m+1)T),

where T is the period of two-step modulation, m. Note that the system is still in MLDs regime if only employing static Hamiltonian H0 '. All control parameters of Hamiltonian H0 ' are added the label “” to distinguish the Hamiltonian H0, and we set τ1=Tτ1 for brevity hereafter. Here, the remarkable difference with other periodic works [43–45] is that we cannot employ the Baker-Campbell-Hausdorff expansion to calculate the effective Hamiltonian Heff, since the driving frequency ω=2πT is demanded for the same magnitude to the energy gap in this periodic system. The method we adopt is to directly calculate the evolution operator within a period T, then achieve the effective Hamiltonian Heff by definition U(T)eiHeffT [46]. The detail derivation process is as follows. At first, we suppose the time interval τ1() satisfying the condition: Θ1()=(E2()E1())τ1()=(2n+1)π, n, and Θ2()=(E3()E1())τ1()=ϕ(). Obviously, the time interval τ1(') would be small when the detuning Delta1 is large in MLDs regime. According to Eq. (9), the evolution operator of system in a period T reads,

U(T)=U(τ1 ')U(τ1)=eiH0 'τ1 'eiH0τ1(1z1z2z41z3z5z6ei(ϕ+ϕ)),
where
z1=y1y1 +y2 y3, z2=y2+eiϕy2 +y1 y3, z3=y1 y2y3+eiϕy3 ,z4=y1 y1+y2y3 , z5=y2 +eiϕy2+y1y3 , z6=y1y2 y3 +eiϕy3,y1()=(eiϕ()1)x1()sin α()2x2()cos α(), y2()=(eiϕ()1)x1()cos α()2x2()sin α(),y3()=(eiϕ()+1)cos α()sin α().

Note that there are many ways to choose control parameters to perform two-step modulation in principle. For instance, one can only modulate the detunings Δk and Δk ' (k = 1 or k = 2) while keeping other control parameter fixed, or one can only modulate the coulping strengths Ωk and Ωk ' while keeping other control parameter fixed, and so on. Here, we exemplify one of them: Δk '=Δk, Ω2 '=Ω2, and Ω1 '=Ω1, i.e., employing two-step modulation of the coupling strength {Ω1,Ω1} (other situations of two-step modulation are demonstrated by numerical simulations in Fig. 2 later). As a result, the evolution operator in a period T approximatively reads

U(T)(cos (ϕ1)sin (ϕ1)0sin (ϕ1)cos (ϕ1)000eiφ1),

where ϕ1=4Ω1ξ1sin2α+4Ω1ξ2cos2α. By reversely solving the matrix equation U(T)=eiHeffT, the effective Hamiltonian of two-step modulation reads

Heff=Ωeff|12|+Ωeff*|21|,

where Ωeff=iϕ1T. Therefore, the dynamics of system by two-step modulation is reduced to “Rabi oscillation” between |1 and |2 with the effective “Rabi frequency” Ωeff, without exciting the high level |3.

On the other hand, if we set the time interval τ1() satisfying the condition: Θ2()=(E3()E1())τ1()=(2n+1)π, n, and Θ1()=(E2()E1())τ1()=ϕ(), similar above derivation process, the evolution operator in a period T reads,

U(T)=U(τ1 ')U(τ1)=eiH0 'τ1 'eiH0τ1(1z1z2z1ei2ϕcos2α+sin2αz3z2z3ei2ϕsin2α+cos2α),

where z1=4x1sin α+e2iϕ(eiϕ1)2x2cos α, z2=4x1cos αe2iϕ(eiϕ1)2x2sin α, and z3=(1ei2ϕ)cos αsin α. When eiϕ=1, i.e., E2=(2n+3)E3,n, we find that z1=4x1sin α+4x2cos α, z2=4x1cos α4x2sin α0, and z3=0. Thus, it is invalid for quantum state engineering between |1 and |3 by two-step modulation. However, if eiϕ=1, i.e., E2=2(n+1)E3,n, we find that z1=4x1sin α0, z2=4x1cos α, and z3=0. In this case, by reversely solving the matrix equation U(T)=eiHeffT, we achieve the effective Hamiltonian of two-step modulation:

Heff '=Ωeff '|13|+Ωeff'*|31|,

where Ωeff '=iϕ1 T, ϕ1 '=2Ω1ξ1sin 2α. Therefore, the dynamics of system by two-step modulation is reduced to “Rabi oscillation” between |1 and |3 with the effective “Rabi frequency” Ωeff ', without exciting the intermediate level |2. In fact, except for the narrow spike region: E2/E3(2n+3),n, the effective Hamiltonian of two-step modulation can be always represented by Eq. (20), which is verified by numerical simulations in Fig. 1 (the maximum population P3max of level |3 can mainly reach unit except for this narrow spike region). In particular, the pink-dash line in Fig. 1(b) and Fig. 1(d) also represent the time evolution of population of level |1, but they are plotted by the effective Hamiltonian Heff ' in Eq. (20). The highly consistent between blue-solid line and pink-dash line demonstrates that the effective Hamiltonian Heff ' is valid for describing the system dynamics by two-step modulation.

 figure: Fig. 1

Fig. 1 (a) The maximum population Pkmax of level |k (k=2,3) during the whole evolution versus E2/E3 by periodically modulating the coupling strength {Ω1,Ω1 }, where the initial state is |1 and Δ1/Ω1=60, Ω2()/Ω1=2, Ω1=Ω1, Δk =Δk (k=1,2), τ1()=πE3()E1(). (b-d) The time evolution of populations Pk (k=1,2,3) of each levels with different E2/E3. Except for pink-dash line, all system dynamics are simulated by using Hamiltonian (11).

Download Full Size | PDF

Finally, it is worth mentioning that two-step modulation of other control parameters can also be used to implement quantum state engineering in MLDs regime. Since the derivation process is similar to the process of periodically modulating coupling strength {Ω1,Ω1}, we just present the numerical results in periodically modulating other control parameters by choosing the time interval τ1()=πE3()E1(), which implement the transition between |1 and |3. Figures 2(a)-2(d) respectively denote the time evolution of population P3 by two-step modulation of coupling strength {Ω1,Ω1 '}, coupling strength {Ω2,Ω2 '}, detuning {Δ1,Δ1 '}, and detuning {Δ2,Δ2 '}, witnessing the feasibility of two-step modulation in three-level system. To be more specific, an inspection of Fig. 2(a) shows that the period of Rabi oscillation is long when increasing the coupling strength Ω1 '. In particular, the transition process become showly when the coupling strength Ω1 ' approaches to Ω1. This consequence is not surprising since two-step modulation reverts to one-step modulation if Ω1 '=Ω1 and the system dynamics is frozen in this regime. Similar results are also achieved for two-step modulation of coupling strength {Ω2,Ω2 '} and detuning {Δ1,Δ1 '}, as shown in Figs. 2(b)-2(c). However, as shown in Fig. 2(d), it is quite different from the two-step modulation of detuning {Δ2,Δ2 '}, where the period of Rabi oscillation remains unchange if Δ2 '>0 and increases with the decreasing of Δ2 ' ifΔ2 '<0.

 figure: Fig. 2

Fig. 2 (a) The population P3 versus evolution time and coupling strength Ω1  under two-step modulation of the coupling strength {Ω1,Ω1 }, where Ω2/Ω1=2, Δ1/Ω1=60, Δ2/Ω1=30, Ω2 =Ω2, Δk =Δk (k=1,2), τ1()=πE3()E1(). (b) The population P3 versus evolution time and coupling strength Ω2  under two-step modulation of the coupling strength {Ω2,a2 }, where Ω2/Ω1=2, Δ1/Ω1=60, Δ2/Ω1=20, Ω1 =Ω1, Δk =Δk (k=1,2), τ1()=πE3()E1(). (c) The population P3 versus evolution time and detuning Δ1  under two-step modulation of the detuning {Δ1,Δ1 }, where Ω2/Ω1=2, Δ1/Ω1=60, Δ2/Ω1=30, Ωk =Ωk (k=1,2), Δ2 =Δ2, τ1()=πE3()E1(). (d) The population P3 versus evolution time and detuning Δ2  under two-step modulation of the detuning {Δ2,Δ2 }, where Ω2/Ω1=2, Δ1/Ω1=60, Δ2/Ω1=20, Ωk =Ωk (k=1,2), Δ1 =Δ1, τ1()=πE3()E1().

Download Full Size | PDF

3. Applications in Rydberg systems

In this section, we take Rydberg state transitions as first example. As shown in Fig. 3(a), a Rydberg atom is coupled by two laser fields, where the coupling strengths are Ω1 and Ω2 respectively. The states |5S1/2, |5P1/2, and |62D3/2 correspond to the levels |1, |2, and |3, respectively. To stimulate the Rydberg state |62D3/2 from the ground state |5S1/2, one usually adopts stimulated Raman adiabatic passage (STIRAP) and it requires two-photon resonance condition (i.e., Δ2=0) between two laser fields in experiments [47, 48]. In other words, two laser frequencies must be chosen specially, e.g., the wavelengths of laser fields are 795 nm and 474 nm in Fig. 3(a) [49], respectively. But what happen when the two-photon resonance condition cannot be satisfied well? Clearly, as shown in Fig. 4(a), the population of Rydberg state |62D3/2 sharply drops with the increase of Δ2/Ω1, verifying that STIRAPis invalid even the detuning Δ2 is small (also cf. the “narrow spikes” of Fig. 13 in [50]). To solve this issue, we employ two-step modulation, so that one can safely ignore two-photon resonance condition. The resulting benefits is that two laser frequencies can be chosen arbitrarily now.

 figure: Fig. 3

Fig. 3 (a) The structure of Rydberg atom coupled by two laser fields and a microwave field. (b) The structure of two identical atoms coupled by laser fields and Rydberg-Rydberg interaction.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 (a) The population P3 of level |3 versus Δ2 in STIRAP, where Ω1(t)=Ω1e(tτ)2τ2, Ω2(t)=Ω1et2τ2, τ = 200, and Δ1/Ω1=30. (b) The time evolution of populations Pk (k=1,2,3) of each levels by periodically modulating coupling strength {Ω1,Ω1} in three-level system, where Ω1/Ω1=1, Δ1()/Ω1=60, Δ2()/Ω1=30, Ω2()/Ω1=2, τ1()=πE3()E1(). (c-d) The time evolution of populations Pk (k=1,2,3,4) of each levels by periodically modulating coupling strength {Ω1,Ω1} in four-level system, where Ω1/Ω1=1, Δ1()/Ω1=60, Δ2()/Ω1=30, Ω2()/Ω1=2, Δ3()/Ω1=28.8 , Ω3()/Ω1=2. The time interval satisfies τ1()=πE3()E1() in panel (c) while the time interval satisfies τ1()=πE4()E1() in panel (d). All system dynamics are simulated by Hamiltonian (11).

Download Full Size | PDF

At the start we can arbitrarily choose two laser frequencies to ensure the system in MLDs regime. Note that different laser frequencies only affect the period of two-step modulation. Then we set the system in the ground state |5S1/2 initially, and the population of each levels are frozen due to MLDs regime. Next we periodically modulate coupling strength to implement “Rabi oscillation” between ground state |5S1/2gle (=|1) and Rydberg state |62D3/2 (=|3). Figure 4(b) demonstrates the system dynamics when the time interval satisfies τ1()=πE3()E1() in two-step modulation of coupling strength {Ω1,Ω1}, and the Rydberg state is achieved at specific time ts=37.2/Ω1. In practice, this process is easily realized by modulating the phase of laser field with square-wave generator to produce π-phase difference [51–55]. After achieving the Rydberg state, we need to remove phase modulator since the populations are frozen again in MLDs regime. Note that if two different coupling strengths are adopted in two-step modulation, it needs an attenuation modulator rather than phase modulator.

The second application of two-step modulation is to prepare desired superposition of two Rydberg states: |ψ=cos ϑ|62D3/2+sin ϑ|63P1/2, which is exploited for fast Rydberg quantum gate [55]. In the experiment [49], |ψ is achieved by several operation steps, including switching off and on the laser fields and the microwave field in sequence. In particular this experiment also requires Δ2=0 and δ = 0 in Fig.3(a). However, the two-step modulation offers a quite simple way to achieve this goal, which only requires periodically modulating the coupling strength {Ω1,Ω1}. In Figs. 4(c)-4(d), one directly drives the ground state to the desired superposition of two Rydberg states by choosing specific time ts, where the angle ϑ is determined by the coupling strength Ω3 and detuningδ, i.e., ϑ12tan12Ω3δ. To be more specific, if we choose the coupling strengths Ω1=100MHz, Ω2=Ω3=200MHz, the detunings Δ1=6GHz, Δ2=3GHz, δ = 120MHz, the period of phase modulator would be T = 2ns. Hence, the superposition of two Rydberg states is achieved at time ts = 446ns by two-step modulation of coupling strength {Ω1,Ω1}.

 figure: Fig. 5

Fig. 5 The time evolution of populations Pm (m=gg,T,rr) of each levels by periodically modulating coupling strength {Ωeff,Ωeff}, where Ωeff =0.5Ωeff, Δ1/Ωeff=23, V/Ωeff=39. (a) τ1()=πET()Egg(), (b) τ1()=πErr()Egg(). (c) The time evolution of populations Pm (m=gg,T,rr) of each levels without two-step modulation, Δ1/Ωeff=23, V/Ωeff=39.

Download Full Size | PDF

Another application of two-step modulation in Rydberg atoms is that we can control the transition between Rydberg blockade regime and Rydberg antiblockade regime. As shown in Fig. 3(b), two identical atoms have ground state |g and Rydberg state |r. The |g|r transition is coupled by laser fields with the effective coupling strength Ωeff and the detuning Δ1. At the same time, there exists Rydberg-Rydberg interaction between two atoms, where the interaction strength is V. Thus, the system Hamiltonian reads

H0=ΩeffeiΔ1t(|g11r|I2+I1|g22r|+H.c.)+V|rrrr|,
where Ik denotes the identity operator of the k-th atom (k=1,2), and |mn is the abbreviation of |m1|n2 (m,n=g,r). In the rotation frame defined by R=ei(Δ1|TT|+2Δ1|rrrr|)t, the system Hamiltonian becomes
H0=Δ1|TT|+(V2Δ1)|rrrr|+2Ωeff(|Tgg|+|Trr|+H.c.),

where |T=12(|gr+|rg). When Δ1=0, the system is in the Rydberg blockade regime, i.e., the doubly excited Rydberg state |rr cannot be stimulated from |T directly. However, when Δ1=V2, the system is in the Rydberg antiblockade regime, i.e., the doubly excited Rydberg states |rr can be stimulated from |gg. In other words, it requires different laser fields to make the system in Rydberg blockade regime or Rydberg antiblockade regime. Here, we demonstrate that the Rydberg blockade regime and Rydberg antiblockade regime can be switched by merely regulating the period of two-step modulation in the same laser fields. Figure 5 presents the system dynamics by periodically modulating coupling strength {Ωeff,Ωeff}. As shown in Fig. 5(a), when we choose the time interval τ1() satisfy: τ1()=πET()Egg(), it emerges Rabi oscillation between |gg and |T and hinders the transition to |rr. That is, the system is in the Rydberg blockade regime. However, as shown in Fig. 5(b), when we choose the time interval τ1() satisfy: τ1()=πErr()Egg(), it emerges Rabi oscillation between |gg and |rr, which means the system is in the Rydberg antiblockade regime. Note that the slight oscillation in the populations Pm can be effectively eliminated by decreasing the time interval τ1() of two-step modulation. When removing the two-step modulation, the system is in neither Rydberg blockade regime nor Rydberg antiblockade regime, as shown in Fig. 5(c). Therefore, we can regulate the time interval τ1() of two-step modulation to determine the system in Rydberg blockade regime or Rydberg antiblockade regime, or neither of them, which does not require different laser fields now.

 figure: Fig. 6

Fig. 6 (a) The structure of three-level system coupled by a laser field, where the detuning Δ1 exactly matches with the transition frequency ω23. (b) The structure of Rubidium atom driven by single laser field with large detunings. (c-d) The structure of Ne* atom coupled by laser fields ε1 and ε2, where the degeneracy of sublevels are removed by magnetic field B.

Download Full Size | PDF

4. Implementation of selective transitions in a multilevel system

In most situations, multilevel nature of quantum system needs to be considered, since it may make invalid for the hypothesis that the laser fields only interact with two levels. We first investigate a counterintuitive phenomenon in two-step modulation. The physical model consists of a three-level system coupled by single laser field with coupling strength Ω1, where we artificially add large detuning Δ1 that exactly matches with the transition frequency ω23, as shown in Fig. 6(a). In the rotating frame, the Hamiltonian reads

H0=Δ1|22|+Ω1|12|+Ω1|13|+H.c.

Naturally, if we do not employ two-step modulation, the laser field would only stimulate the |1|3 transition due to the resonance condition, as shown in Fig. 7(a). However, we verify in Fig. 7(b) that, regardless of resonance condition, the laser field would stimulate the |1|2 transition by two-step modulation, sharply suppressing the transition to the level |3. As a result, even though the laser field resonantly drives the transition between |1 and |3, the |1|3 transition still cannot occur by two-step modulation.

 figure: Fig. 7

Fig. 7 The time evolution of populations Pm (m=1,2,3) of each levels (a) without, (b) with, periodically modulating coupling strength {Ω1,Ω1 }, where Ω1 =Ω1, Δ1/Ω1=48, τ1=πE2E1, u1 =πE3 E1 .

Download Full Size | PDF

This counterintuitive phenomenon can be applied to implement different atomic transitions with single laser field by artificially adding MLDs in multilevel systems. As shown in Fig. 6(b), suppose that the multilevel atom is coupled by a single laser field, which drives |1|2 transition with large detuning Δ1, drives the |1|3 transition with large detuning Δ2, and drives the |1|4 transition with large detuning Δ3. In the rotating frame, the Hamiltonian reads

H0=k=24Δk1|kk|+Ω1|1k|+Ω1|k1|.

There is no doubt that the laser field cannot drive any transition without any modulations in system due to MLDs condition. However, this situation is changed by two-step modulation. Figure 8 demonstrates different atomic transition is achieved by two-step modulation of the same laser field. As a concrete example, with regard to the Rubidium 85 D2 transition hyperfine structure [56], one can stimulate the 52S1/2|F=252P3/2|F=2 transition with the time interval τ1()=πE2()E1() in Fig. 8(a), stimulate the 52S1/2|F=252P3/2|F=3 transition with the time interval τ1()=πE3()E1() in Fig. 8(b), or stimulate the 52S1/2|F=252P3/2|F=4 transition with the time interval τ1()=πE4()E1() in Fig. 8(c) by two-modulation of coupling strength {Ω1,Ω1}. In other words, different periods of two-step modulation would determine different atomic transitions in the same system.

 figure: Fig. 8

Fig. 8 The time evolution of populations Pm (m=1,2,3) of each levels by periodically modulating coupling strength {Ω1,Ω1} with different time interval τ1(), where Ω1=Ω1, Δ1/Ω1=30, τ2/Ω1=53, Δ3/Ω1=100. (a) τ1()=πE2()E1(), (b) τ1()=πE3()E1(), (c) τ1()=πE4()E1().

Download Full Size | PDF

Other applications of two-step modulation can be found in the Ne * atom system. As shown in Fig. 6(c), when the laser field ε1 is parallel to the magnetic field B while the laser field ε2 is perpendicular to the magnetic field B, it would be reduced to five-level system. To be specific, the laser field ε1 drives the transitions |1|2 and |1|2 with the coupling strength Ω1, while the laser field ε2 drives the transitions |2|3 and |2|3 with the coupling strength Ω2. In the rotating frame, the system Hamiltonian reads

H0=Δ1|22|+Δ3|22|+(Δ1+Δ2)|33|+(Δ3+Δ4)|33|+Ω1|12|+Ω1|12|+Ω2|23|+Ω2|23|+H.c.

With the help of two-step modulation, the transition paths are selective by choosing different time interval τ1() in this system. For instance, if τ1()=πE3()E1(), as shown in Fig. 9(a), we achieve the level |3 through transition path ‘|1|2|3’. If τ1()=πE3()E1(), as shown in Fig. 9(b), we achieve the level |3 through transition path ‘|1|2|3’. In practice, we just change the period of square-wave on phase modulator to realize two different transition paths.

 figure: Fig. 9

Fig. 9 The time evolution of populations Pm (m=1,2,3,2,3) of each levels by periodically modulating coupling strength {Ω1,Ω1} in five-level system, where Ω1=Ω1, Ω2/Ω1=2, Δ1/Ω1=33, lta2/Ω1=9, Δ3/Ω1=36, Δ4/Ω1=6. (c) τ1()=πE3()E1(), (d) τ1()=πE3()E1().

Download Full Size | PDF

However, if the laser field ε1 is perpendicular to the magnetic field B while the laser field ε2 is parallel to the magnetic field B, only four levels are coupled by laser fields in this system, as shown in Fig. 6(d). To be specific, the laser field ε1 drives the |1|2 transition with Rabi frequency Ω1 and detuning Δ1. The laser field ε2 drives the |2|3 and |2|3 transition with Rabi frequency Ω2 and detunings (Δ1+Δ2) and (Δ1+Δ3), respectively. Thus, the system Hamiltonian reads

H0=Δ1|22|+Δ2|33|+Δ3|33|+Ω1|12|+Ω2|23|+Ω2|23|+H.c.

We also adopt two-step modulation of coupling strength {Ω1,Ω1 '} while all other physical parameters remain unchanged. If the time interval τ1() satisfies τ1()=(2n+1)πE3()E1(), Rabi oscillation between levels |1 and |3 occurs, as shown in Fig. 10(a). However, if the time interval τ1() satisfies τ1()=(2n+1)πE3()E1(), Rabi oscillation between levels |1 and |3 occurs, as shown in Fig. 10(b). That is, we can control selective transition by choosing different periods of two-step modulation.

 figure: Fig. 10

Fig. 10 The time evolution of populations Pk (k=1,2,3,3) of each levels by periodically modulating coupling strength {Ω1,Ω1}, where Ω2/Ω1=2, Δ1/Ω1=60, Δ2/Ω1=30, Δ3/Ω1=28, Ω1=Ω1, Ω2=Ω2, Δk =Δk,(k=1,2,3). (a) τ1()=(2n+1)πE3()E1(). (b) τ1()=(2n+1)πE3()E1().

Download Full Size | PDF

5. Robust against perturbation of control parameters

Until now, we have studied the physical model in ideal case, i.e., all control parameters are known accurately. However, in practice, the control parameters of system unavoidably exist some unknown perturbations so that it would affect the evolution process. As a result, it is very essential to examine whether or not the two-step modulation is valid in the presence of perturbations. At first, we exemplify the influence of perturbations in laser fields on Rydberg state preparation by periodically modulating coupling strength {Ω1,Ω1 '}, where the system Hamiltonian is rewritten as

H(t)={Δ1|22|+Δ2|33|+(Ω1+δΩ1)|12|+(Ω2+δΩ2)|23|+H.c.,t[mT,mT+τ),Δ1|22|+Δ2|33|+(Ω1 '+δΩ1)|12|+(Ω2+δΩ2)|23|+H.c.,t[mT+τ,(m+1)T).

δΩk (k=1,2) denotes the strength of unknown perturbations in coupling strength Ωk. Figure 11 represents the population P3 of Rydberg state versus perturbations δΩ1 and δΩ2 at the evolution time ts=37.2/Ω1. We can find that the population of Rydberg state is still high (0.988) even though there exist the perturbation (|δΩ2/Ω1|0.05) in the coupling strength Ω2. Particularly, it is almost immune to the perturbation δΩ1, which stems from the fact that the energy gap of system is hardly affected by the perturbation δΩ1 when periodically modulating coupling strength Ω1.

 figure: Fig. 11

Fig. 11 The population P3 of Rydberg state versus the perturbations δΩ1 and δΩ2 in two-step modulation of coupling strength {Ω1,Ω1}, where Ω1/Ω1=1, Δ1()/Ω1=60, Δ2()/Ω1=30, Ω2()/Ω1=2, τ1()=πE3()E1().

Download Full Size | PDF

On the other hand, the square-wave produced by waveform generator might not be perfect in experiments. In the following, we study the influence of the square-wave deformation of coupling strength Ω1(t) on Rydberg state preparation, where the expression now reads

Ω1(t)={Ω1+Ω1Ω11+eγmod(t/T),mod(t/T)<τ2,Ω1+Ω1Ω11+eγ[mod(t/T)τ],τ2mod(t/T)Tτ2,Ω1+Ω1Ω11+eγ[mod(t/T)T],mod(t/T)>Tτ2.

The dimensionless parameter γ represents the hardness of square-wave. Specifically, the shape of coupling strength Ω1(t) in Eq. (29) gradually approaches to square-wave when γ is large, and it becomes perfect square-wave if amma. As illustrations, Figs. 12(b)-12(d) represent some concrete shapes of coupling strength Ω1(t) with different γ. In Fig. 12(a), the blue-dash line denotes the population P3 of Rydberg state as a function of γ at evolution time ts=37.2/Ω1. One observes that the population of Rydberg state maintains a relatively high value for most γ, i.e., the deformation of square-wave has little influence on Rydberg state preparation. Note that only when the square-wave is heavy deformation, e.g., γ = 50 in Fig. 12(b), it would affect the preparation process of Rydberg state. In fact, this shape is not square-wave any more, and the amplitude of coupling strength Ω1(t) cannot reach unit. Nevertheless, the heavy deformation problem can be effectively solved by properly shifting the evolution time of two-step modulation. In Fig. 12(a), the blue-solid line denotes the maximum population P3 of Rydberg state as a function of γ during the evolution process, and the orange-dot line denotes the evolution time ts ' when reaching the maximum population P3(ts '). We can easily find that the population P3 of Rydberg state is almost unchanged at evolution time ts even though there exists heavy deformation in square-wave.

 figure: Fig. 12

Fig. 12 (a) The population P3 of Rydberg state (the left-blue vertical axis) and the evolution time ts achieving the maximum population P3 (the right-orange vertical axis) as a function of γ in two-step modulation of coupling strength {Ω1,Ω1}, where Ω1/Ω1=1, Δ1()/Ω1=60, Δ2()/Ω1=30, Ω2()/Ω1=2, τ1()=πE3()E1(). (b-d) The shapes of coupling strength Ω1(t) with different γ. (b) γ = 50. (c) γ = 100. (d) γ = 1000.

Download Full Size | PDF

6. Conclusion

We have studied a fantastic phenomenon that quantum states can still evolve by periodic modulation even in MLDs regime. The purpose of artificially adding MLDs is to take full advantage of large detuning conditions and freeze system dynamics when removing periodic modulation. By regulating the time interval and the period of two-step modulation, we have demonstrated that the multilevel system can be equivalent into distinct two-level systems. In particular, the atomic system does not need to satisfy resonance condition and the laser frequencies can be chosen arbitrarily in two-step modulation. Additionally, this modulation is robust against perturbation of control parameters and is easily implemented in experiments due to simple square-pulse form.

For its applications, we have shown: achieving direct transition from the ground state to the Rydberg state or the desired superposition of two Rydberg states; switching between Rydberg blockade regime and Rydberg antiblockade regime; stimulating distinct atomic transitions by the same laser field; implementing selective transitions in multilevel system. Recently, the transition of two Rydberg states is controlled by using an addressing beam to produce detuning in experiments [57]. We find that this process can also be controlled by two-step modulation of microwave fields without requiring extra addressing beam. In a word, this periodic modulation would offer us a powerful tool for quantum state engineering as well as implementing a variety of high-fidelity quantum logic gates.

Funding

National Natural Science Foundation of China (11575045, 11674060, 11747011, 11805036, 11534002, 11775048, 61475033); Major State Basic Research Development Program of China (2012CB921601); Fund of Fujian Education Department (JAT170086); Natural Science Foundation of Fujian Province of China (2018J01413).

References

1. M. O. Scully and M. S. Zubairy, Quantum Optics(Cambridge University, 1997). [CrossRef]  

2. M. Nielsen and I. Chuang, Quantum Computation and Quantum Information(Cambridge University, 2010). [CrossRef]  

3. H. Schempp, G. Günter, S. Wüster, M. Weidemüller, and S. Whitlock, “Correlated exciton transport in Rydberg-dressed-atom spin chains,” Phys. Rev. Lett. 115, 093002 (2015). [CrossRef]   [PubMed]  

4. N. Thaicharoen, A. Schwarzkopf, and G. Raithel, “Measurement of the van der Waals interaction by atom trajectory imaging,” Phys. Rev. A 92, 040701(2015). [CrossRef]  

5. C. W. Gardiner and P. Zoller, Quantum Noise(Springer-Verlag, 1999).

6. T. Kitagawa, E. Berg, M. Rudner, and E. Demler, “Topological characterization of periodically driven quantum systems,” Phys. Rev. B 82, 235114 (2010). [CrossRef]  

7. L. Zhou and L. M. Kuang, “Zeno-anti-Zeno crossover via external fields in a one-dimensional coupled-cavity waveguide,” Phys. Rev. A 82, 042113 (2010). [CrossRef]  

8. M. S. Rudner, N. H. Lindner, E. Berg, and M. Levin, “Anomalous edge states and the bulk-edge correspondence for periodically driven two-dimensional systems,” Phys. Rev. X 3, 031005 (2013).

9. Q. J. Tong, J. H. An, L. C. Kwek, H. G. Luo, and C. H. Oh, “Simulating Zeno physics by a quantum quench with superconducting circuits,” Phys. Rev. A 89, 060101 (2014). [CrossRef]  

10. P. M. Perez-Piskunow, G. Usaj, C. A. Balseiro, and L. E. F. Foa Torres, “Floquet chiral edge states in graphene,” Phys. Rev. B 89, 121401 (2014). [CrossRef]  

11. C. Chen, J. H. An, H. G. Luo, C. P. Sun, and C. H. Oh, “Floquet control of quantum dissipation in spin chains,” Phys. Rev. A 91, 052122 (2015). [CrossRef]  

12. P. Ponte, Z. Papić, F. Huveneers, and D. A. Abanin, “Many-body localization in periodically driven systems,” Phys. Rev. Lett. 114, 140401 (2015). [CrossRef]   [PubMed]  

13. K. Iwahori and N. Kawakami, “Long-time asymptotic state of periodically driven open quantum systems,” Phys. Rev. B 94, 184304 (2016). [CrossRef]  

14. Z. C. Shi, W. Wang, and X. X. Yi, “Population transfer driven by far-off-resonant fields,” Opt. Express , 24, 21971–21985 (2016). [CrossRef]   [PubMed]  

15. Y. C. Yang, S. N. Coppersmith, and M. Friesen, “Achieving high-fidelity single-qubit gates in a strongly driven silicon-quantum-dot hybrid qubit,” Phys. Rev. A 95, 062321(2017). [CrossRef]  

16. D. Pagel and H. Fehske, “Non-Markovian dynamics of few emitters in a laser-driven cavity,” Phys. Rev. A 96, 041802 (2017). [CrossRef]  

17. R. Desbuquois, M. Messer, F. Görg, K. Sandholzer, G. Jotzu, and T. Esslinger, “Controlling the Floquet state population and observing micromotion in a periodically driven two-body quantum system,” Phys. Rev. A 96, 053602 (2017). [CrossRef]  

18. K. Jiménez-García, L. J. LeBlanc, R. A. Williams, M. C. Beeler, C. Qu, M. Gong, C. Zhang, and I. B. Spielman, “Tunable spin-orbit coupling via strong driving in ultracold-atom systems,” Phys. Rev. Lett. 114, 125301 (2015). [CrossRef]   [PubMed]  

19. X. Luo, L. Wu, J. Chen, Q. Guan, K. Gao, Z. F. Xu, L. You, and R. Wang, “Tunable atomic spin-orbit coupling synthesized with a modulating gradient magnetic field,” Sci. Rep. 6, 18983 (2016). [CrossRef]   [PubMed]  

20. N. H. Lindner, G. Refael, and V. Galitski, “Floquet topological insulator in semiconductor quantum wells,” Nat. Phys. 7, 490 (2011). [CrossRef]  

21. L. Jiang, T. Kitagawa, J. Alicea, A. R. Akhmerov, D. Pekker, G. Refael, J. I. Cirac, E. Demler, M. D. Lukin, and P. Zoller, “Majorana fermions in equilibrium and in driven cold-atom quantum wires,” Phys. Rev. Lett. 106, 220402(2011). [CrossRef]   [PubMed]  

22. D. E. Liu, A. Levchenko, and H. U. Baranger, “Floquet Majorana fermions for topological qubits in superconducting devices and cold atom systems,” Phys. Rev. Lett. 111, 047002 (2013). [CrossRef]  

23. Y. T. Katan and D. Podolsky, “Modulated Floquet topological insulators,” Phys. Rev. Lett. 110, 016802 (2013). [CrossRef]  

24. L. E. F. Foa Torres, P. M. Perez-Piskunow, C. A. Balseiro, and G. Usaj, “Multiterminal conductance of a floquet topological insulator,” Phys. Rev. Lett. 113, 266801 (2014). [CrossRef]  

25. M. Benito, A. Gómez-León, V. M. Bastidas, T. Brandes, and G. Platero, “Floquet engineering of long-range p-wave superconductivity,” Phys. Rev. B 90, 205127 (2014). [CrossRef]  

26. J. H. Shirley, “Solution of the Schrödinger equation with a Hamiltonian periodic in time,” Phys. Rev. 138, B979 (1965). [CrossRef]  

27. H. Sambe, “Steady states and quasienergies of a quantum-mechanical system in an oscillating feld,” Phys. Rev. A 7, 2203 (1973). [CrossRef]  

28. S. N. Shevchenko, S. Ashhab, and F. Nori, “Landau-Zener-Stückelberg interferometry,” Phys. Rep. 492, 1–30 (2010). [CrossRef]  

29. P. Huang, J. Zhou, F. Fang, X. Kong, X. Xu, C. Ju, and J. Du, “Landau-Zener-Stückelberg interferometry of a single electronic spin in a noisy environment,” Phys. Rev. X 1, 011003 (2011).

30. M. P. Silveri, K. S. Kumar, J. Tuorila, J. Li, A. Vepsäläinen, E. V. Thuneberg, and G. S. Paraoanu, “Stückelberg interference in a superconducting qubit under periodic latching modulation,” New J. Phys. 17, 043058 (2015). [CrossRef]  

31. M. B. Kenmoe and L. C. Fai, “Periodically driven three-level systems,” Phys. Rev. B , 94, 125101 (2016). [CrossRef]  

32. S. A. Malinovskaya and V. S. Malinovsky, “Chirped-pulse adiabatic control in coherent anti-Stokes Raman scattering for imaging of biological structure and dynamics,” Opt. Lett. 32, 707–709 (2007). [CrossRef]  

33. S. Malinovskaya, “Chirped pulse control methods for imaging of biological structure and dynamics,” Int. J. Quant. Chem. 107, 3151–3158 (2007). [CrossRef]  

34. E. Kuznetsova, G. Liu, and S. A Malinovskaya, “Adiabatic rapid passage two-photon excitation of a Rydberg atom,” Phys. Scr. T160, 014024 (2014). [CrossRef]  

35. S. A. Malinovskaya, “Design of many-body spin states of Rydberg atoms excited to highly tunable magnetic sublevels,” Opt. Lett. 42, 314–317 (2017). [CrossRef]  

36. D. Jaksch, J. I. Cirac, P. Zoller, S. L. Rolston, R. Côté, and M. D. Lukin, “Fast quantum gates for neutral atoms,” Phys. Rev. Lett. 85, 2208 (2000). [CrossRef]   [PubMed]  

37. M. D. Lukin, M. Fleischhauer, R. Cote, L. M. Duan, D. Jaksch, J. I. Cirac, and P. Zoller, “Dipole blockade and quantum information processing in mesoscopic atomic ensembles,” Phys. Rev. Lett. 87, 037901 (2001). [CrossRef]  

38. A. Gaëtan, Y. Miroshnychenko, T. Wilk, A. Chotia, M. Viteau, D. Comparat, P. Pillet, A. Browaeys, and P. Grangier, “Observation of collective excitation of two individual atoms in the Rydberg blockade regime,” Nat. Phys. 5, 115 (2009). [CrossRef]  

39. E. Urban, T. A. Johnson, T. Henage, L. Isenhower, D. D. Yavuz, T. G. Walker, and M. Saffman, “Observation of Rydberg blockade between two atoms,” Nat. Phys. 5, 110 (2009). [CrossRef]  

40. C. Ates, T. Pohl, T. Pattard, and J. M. Rost, “Antiblockade in Rydberg Excitation of an Ultracold Lattice Gas,” Phys. Rev. Lett. 98, 023002 (2007). [CrossRef]   [PubMed]  

41. T. Amthor, C. Giese, C. S. Hofmann, and M. Weidemüller, “Evidence of Antiblockade in an Ultracold Rydberg Gas,” Phys. Rev. Lett. 104, 013001 (2010). [CrossRef]  

42. S. L. Su, Y. Gao, E. Liang, and S. Zhang, “Fast Rydberg antiblockade regime and its applications in quantum logic gates,” Phys. Rev. A 95, 022319 (2017). [CrossRef]  

43. N. Goldman and J. Dalibard, “Periodically driven quantum systems: Effective Hamiltonians and engineered gauge fields,” Phys. Rev. X 4, 031027 (2014).

44. H. Ribeiro, A. Baksic, and A. A. Clerk, “Systematic Magnus-based approach for suppressing leakage and nonadiabatic errors in quantum dynamics,” Phys. Rev. X 7, 011021 (2017).

45. P. W. Claeys and J. S. Caux, “Breaking the integrability of the Heisenberg model through periodic driving,” arXiv:1708.07324 (2017).

46. M. S. P. Eastham, The Spectral Theory of Periodic Differential Equations (Scottish Academic, 1973).

47. C. Tresp, C. Zimmer, I. Mirgorodskiy, H. Gorniaczyk, A. Paris-Mandoki, and S. Hofferberth, “Single-photon absorber based on strongly interacting Rydberg atoms,” Phys. Rev. Lett. 117, 223001 (2016). [CrossRef]   [PubMed]  

48. N. Thaicharoen, A. Schwarzkopf, and G. Raithel, “Control of spatial correlations between Rydberg excitations using rotary echo,” Phys. Rev. Lett. 118, 133401 (2017). [CrossRef]  

49. D. Barredo, H. Labuhn, S. Ravets, T. Lahaye, A. Browaeys, and C. S. Adams, “Coherent excitation transfer in a spin chain of three Rydberg atoms,” Phys. Rev. Lett. 114, 113002 (2015). [CrossRef]   [PubMed]  

50. K. Bergmann, H. Theuer, and B. W. Shore, “Coherent population transfer among quantum states of atoms and molecules,” Rev. Mod. Phys. 70, 1003 (1998). [CrossRef]  

51. J. Cho, D. G. Angelakis, and S. Bose, “Fractional quantum Hall state in coupled cavities,” Phys. Rev. Lett. 101, 246809 (2008). [CrossRef]   [PubMed]  

52. Z. X. Chen, Z. W. Zhou, X. Zhou, X. F. Zhou, and G. C. Guo, “Quantum simulation of Heisenberg spin chains with next-nearest-neighbor interactions in coupled cavities,” Phys. Rev. A 81, 022303 (2010). [CrossRef]  

53. T. Keating, C. H. Baldwin, Y. Y. Jau, J. Lee, G. W. Biedermann, and I. H. Deutsch, “Arbitrary Dicke-state control of symmetric Rydberg ensembles,” Phys. Rev. Lett. 117, 213601 (2016). [CrossRef]   [PubMed]  

54. Y. C. Zhang, X. F. Zhou, X. X. Zhou, G. C. Guo, and Z. W. Zhou, “Cavity-assisted single-mode and two-mode spin-squeezed states via phase-locked atom-photon coupling,” Phys. Rev. Lett. 118, 083604 (2017). [CrossRef]   [PubMed]  

55. X. F. Shi and T. A. B. Kennedy, “Annulled van der Waals interaction and fast Rydberg quantum gates,” Phys. Rev. A 95, 043429 (2017). [CrossRef]  

56. D. A. Steck, Rubidium 85 D line data, available online at http://steck.us/alkalidata (revision 2.1.6, 20 September 2008).

57. S. de Léséleuc, D. Barredo, V. Lienhard, A. Browaeys, and T. Lahaye, “Optical control of the resonant dipole-dipole interaction between Rydberg atoms,” Phys. Rev. Lett. 119, 053202 (2017). [CrossRef]   [PubMed]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (12)

Fig. 1
Fig. 1 (a) The maximum population P k m a x of level | k ( k = 2 , 3 ) during the whole evolution versus E 2 / E 3 by periodically modulating the coupling strength { Ω 1 , Ω 1   } , where the initial state is | 1 and Δ 1 / Ω 1 = 60 , Ω 2 ( ) / Ω 1 = 2 , Ω 1 = Ω 1 , Δ k   = Δ k ( k = 1 , 2 ), τ 1 ( ) = π E 3 ( ) E 1 ( ) . (b-d) The time evolution of populations Pk ( k = 1 , 2 , 3 ) of each levels with different E 2 / E 3 . Except for pink-dash line, all system dynamics are simulated by using Hamiltonian (11).
Fig. 2
Fig. 2 (a) The population P3 versus evolution time and coupling strength Ω 1   under two-step modulation of the coupling strength { Ω 1 , Ω 1   } , where Ω 2 / Ω 1 = 2 , Δ 1 / Ω 1 = 60 , Δ 2 / Ω 1 = 30 , Ω 2   = Ω 2 , Δ k   = Δ k ( k = 1 , 2 ), τ 1 ( ) = π E 3 ( ) E 1 ( ) . (b) The population P3 versus evolution time and coupling strength Ω 2   under two-step modulation of the coupling strength { Ω 2 , a 2   } , where Ω 2 / Ω 1 = 2 , Δ 1 / Ω 1 = 60 , Δ 2 / Ω 1 = 20 , Ω 1   = Ω 1 , Δ k   = Δ k ( k = 1 , 2 ), τ 1 ( ) = π E 3 ( ) E 1 ( ) . (c) The population P3 versus evolution time and detuning Δ 1   under two-step modulation of the detuning { Δ 1 , Δ 1   } , where Ω 2 / Ω 1 = 2 , Δ 1 / Ω 1 = 60 , Δ 2 / Ω 1 = 30 , Ω k   = Ω k ( k = 1 , 2 ), Δ 2   = Δ 2 , τ 1 ( ) = π E 3 ( ) E 1 ( ) . (d) The population P3 versus evolution time and detuning Δ 2   under two-step modulation of the detuning { Δ 2 , Δ 2   } , where Ω 2 / Ω 1 = 2 , Δ 1 / Ω 1 = 60 , Δ 2 / Ω 1 = 20 , Ω k   = Ω k ( k = 1 , 2 ), Δ 1   = Δ 1 , τ 1 ( ) = π E 3 ( ) E 1 ( ) .
Fig. 3
Fig. 3 (a) The structure of Rydberg atom coupled by two laser fields and a microwave field. (b) The structure of two identical atoms coupled by laser fields and Rydberg-Rydberg interaction.
Fig. 4
Fig. 4 (a) The population P3 of level | 3 versus Δ2 in STIRAP, where Ω 1 ( t ) = Ω 1 e ( t τ ) 2 τ 2 , Ω 2 ( t ) = Ω 1 e t 2 τ 2 , τ = 200, and Δ 1 / Ω 1 = 30 . (b) The time evolution of populations Pk ( k = 1 , 2 , 3 ) of each levels by periodically modulating coupling strength { Ω 1 , Ω 1 } in three-level system, where Ω 1 / Ω 1 = 1 , Δ 1 ( ) / Ω 1 = 60 , Δ 2 ( ) / Ω 1 = 30 , Ω 2 ( ) / Ω 1 = 2 , τ 1 ( ) = π E 3 ( ) E 1 ( ) . (c-d) The time evolution of populations Pk ( k = 1 , 2 , 3 , 4 ) of each levels by periodically modulating coupling strength { Ω 1 , Ω 1 } in four-level system, where Ω 1 / Ω 1 = 1 , Δ 1 ( ) / Ω 1 = 60 , Δ 2 ( ) / Ω 1 = 30 , Ω 2 ( ) / Ω 1 = 2 , Δ 3 ( ) / Ω 1 = 28.8 , Ω 3 ( ) / Ω 1 = 2 . The time interval satisfies τ 1 ( ) = π E 3 ( ) E 1 ( ) in panel (c) while the time interval satisfies τ 1 ( ) = π E 4 ( ) E 1 ( ) in panel (d). All system dynamics are simulated by Hamiltonian (11).
Fig. 5
Fig. 5 The time evolution of populations Pm ( m = g g , T , r r ) of each levels by periodically modulating coupling strength { Ω e f f , Ω e f f } , where Ω e f f   = 0.5 Ω e f f , Δ 1 / Ω e f f = 23 , V / Ω e f f = 39 . (a) τ 1 ( ) = π E T ( ) E g g ( ) , (b) τ 1 ( ) = π E r r ( ) E g g ( ) . (c) The time evolution of populations Pm ( m = g g , T , r r ) of each levels without two-step modulation, Δ 1 / Ω e f f = 23 , V / Ω e f f = 39 .
Fig. 6
Fig. 6 (a) The structure of three-level system coupled by a laser field, where the detuning Δ1 exactly matches with the transition frequency ω23. (b) The structure of Rubidium atom driven by single laser field with large detunings. (c-d) The structure of Ne* atom coupled by laser fields ε 1 and ε 2 , where the degeneracy of sublevels are removed by magnetic field B .
Fig. 7
Fig. 7 The time evolution of populations Pm ( m = 1 , 2 , 3 ) of each levels (a) without, (b) with, periodically modulating coupling strength { Ω 1 , Ω 1   } , where Ω 1   = Ω 1 , Δ 1 / Ω 1 = 48 , τ 1 = π E 2 E 1 , u 1   = π E 3   E 1   .
Fig. 8
Fig. 8 The time evolution of populations Pm ( m = 1 , 2 , 3 ) of each levels by periodically modulating coupling strength { Ω 1 , Ω 1 } with different time interval τ 1 ( ) , where Ω 1 = Ω 1 , Δ 1 / Ω 1 = 30 , τ 2 / Ω 1 = 53 , Δ 3 / Ω 1 = 100 . (a) τ 1 ( ) = π E 2 ( ) E 1 ( ) , (b) τ 1 ( ) = π E 3 ( ) E 1 ( ) , (c) τ 1 ( ) = π E 4 ( ) E 1 ( ) .
Fig. 9
Fig. 9 The time evolution of populations Pm ( m = 1 , 2 , 3 , 2 , 3 ) of each levels by periodically modulating coupling strength { Ω 1 , Ω 1 } in five-level system, where Ω 1 = Ω 1 , Ω 2 / Ω 1 = 2 , Δ 1 / Ω 1 = 33 , l t a 2 / Ω 1 = 9 , Δ 3 / Ω 1 = 36 , Δ 4 / Ω 1 = 6 . (c) τ 1 ( ) = π E 3 ( ) E 1 ( ) , (d) τ 1 ( ) = π E 3 ( ) E 1 ( ) .
Fig. 10
Fig. 10 The time evolution of populations Pk ( k = 1 , 2 , 3 , 3 ) of each levels by periodically modulating coupling strength { Ω 1 , Ω 1 } , where Ω 2 / Ω 1 = 2 , Δ 1 / Ω 1 = 60 , Δ 2 / Ω 1 = 30 , Δ 3 / Ω 1 = 28 , Ω 1 = Ω 1 , Ω 2 = Ω 2 , Δ k   = Δ k , ( k = 1 , 2 , 3 ) . (a) τ 1 ( ) = ( 2 n + 1 ) π E 3 ( ) E 1 ( ) . (b) τ 1 ( ) = ( 2 n + 1 ) π E 3 ( ) E 1 ( ) .
Fig. 11
Fig. 11 The population P3 of Rydberg state versus the perturbations δ Ω 1 and δ Ω 2 in two-step modulation of coupling strength { Ω 1 , Ω 1 } , where Ω 1 / Ω 1 = 1 , Δ 1 ( ) / Ω 1 = 60 , Δ 2 ( ) / Ω 1 = 30 , Ω 2 ( ) / Ω 1 = 2 , τ 1 ( ) = π E 3 ( ) E 1 ( ) .
Fig. 12
Fig. 12 (a) The population P3 of Rydberg state (the left-blue vertical axis) and the evolution time t s achieving the maximum population P3 (the right-orange vertical axis) as a function of γ in two-step modulation of coupling strength { Ω 1 , Ω 1 } , where Ω 1 / Ω 1 = 1 , Δ 1 ( ) / Ω 1 = 60 , Δ 2 ( ) / Ω 1 = 30 , Ω 2 ( ) / Ω 1 = 2 , τ 1 ( ) = π E 3 ( ) E 1 ( ) . (b-d) The shapes of coupling strength Ω 1 ( t ) with different γ. (b) γ = 50. (c) γ = 100. (d) γ = 1000.

Equations (22)

Equations on this page are rendered with MathJax. Learn more.

H 0 = k = 1 3 ω k | k k | + k = 1 2 Ω k e i ω k l t | k k + 1 | + H . c . ,
H 0 = k = 1 2 Δ k | k + 1 k + 1 | + Ω k | k k + 1 | + H . c . ,
H 0 = ( 0 Ω 1 0 Ω 1 Δ 1 Ω 2 0 Ω 2 Δ 2 ) .
0 = S + H 0 S = ( 0 Ω 1 sin  α Ω 1 cos  α Ω 1 sin  α ξ 1 0 Ω 1 cos  α 0 ξ 2 ) ,
E 1 ξ 1 x 1 2 ξ 2 x 2 2 ,    | E 1 | 1 x 1 | ξ 1 x 2 | ξ 2 , E 2 ξ 1 + ξ 1 x 1 2 ,       | E 2 | ξ 1 + x 1 | 1 , E 3 ξ 2 + ξ 2 x 2 2 ,       | E 3 | ξ 2 + x 2 | 1 .
U ( t ) = e i 0 t = ( 1 x 1 ( e i Θ 1 1 ) x 2 ( e i Θ 2 1 ) x 1 ( e i Θ 1 1 ) e i Θ 1 x 1 x 2 x 2 ( e i Θ 2 1 ) x 1 x 2 e i Θ 2 ) ,
  U ( t ) = S U ( t ) S +   ( 1 ( e i Θ 2 1 ) x 1 sin α + ( e i Θ 1 1 ) x 2 cos α ( e i Θ 2 1 ) x 1 cos α + ( 1 e i Θ 1 ) x 2 sin α ( e i Θ 2 1 ) x 1 sin α + ( e i Θ 1 1 ) x 2 cos α e i Θ 2 sin 2 α + e i Θ 1 cos 2 α ( e i Θ 2 e i Θ 1 ) cos α sin α ( e i Θ 2 1 ) x 1 cos α + ( 1 e i Θ 1 ) x 2 sin α ( e i Θ 2 e i Θ 1 ) cos α sin α e i Θ 2 cos 2 α + e i Θ 1 sin 2 α ) .
H ( t ) = { H 0 = k = 1 2 Δ k | k + 1 k + 1 | + Ω k | k k + 1 | + H . c . ,   t [ m T , m T + τ 1 ) , H 0   ' = k = 1 2 Δ k   ' | k + 1 k + 1 | + Ω k   ' | k k + 1 | + H . c . ,   t [ m T + τ 1 , ( m + 1 ) T ) ,
U ( T ) = U ( τ 1   ' ) U ( τ 1 ) = e i H 0   ' τ 1   ' e i H 0 τ 1 ( 1 z 1 z 2 z 4 1 z 3 z 5 z 6 e i ( ϕ + ϕ ) ) ,
z 1 = y 1 y 1   + y 2   y 3 ,   z 2 = y 2 + e i ϕ y 2   + y 1   y 3 ,   z 3 = y 1   y 2 y 3 + e i ϕ y 3   , z 4 = y 1   y 1 + y 2 y 3   ,   z 5 = y 2   + e i ϕ y 2 + y 1 y 3   ,   z 6 = y 1 y 2   y 3   + e i ϕ y 3 , y 1 ( ) = ( e i ϕ ( ) 1 ) x 1 ( ) sin  α ( ) 2 x 2 ( ) cos  α ( ) ,   y 2 ( ) = ( e i ϕ ( ) 1 ) x 1 ( ) cos  α ( ) 2 x 2 ( ) sin  α ( ) , y 3 ( ) = ( e i ϕ ( ) + 1 ) cos  α ( ) sin  α ( ) .
U ( T ) ( cos  ( ϕ 1 ) sin  ( ϕ 1 ) 0 sin  ( ϕ 1 ) cos  ( ϕ 1 ) 0 0 0 e i φ 1 ) ,
H e f f = Ω e f f | 1 2 | + Ω e f f * | 2 1 | ,
U ( T ) = U ( τ 1   ' ) U ( τ 1 ) = e i H 0   ' τ 1   ' e i H 0 τ 1 ( 1 z 1 z 2 z 1 e i 2 ϕ cos 2 α + sin 2 α z 3 z 2 z 3 e i 2 ϕ sin 2 α + cos 2 α ) ,
H e f f   ' = Ω e f f   ' | 1 3 | + Ω e f f '* | 3 1 | ,
H 0 = Ω e f f e i Δ 1 t ( | g 11 r | I 2 + I 1 | g 22 r | + H . c . ) + V | r r r r | ,
H 0 = Δ 1 | T T | + ( V 2 Δ 1 ) | r r r r | + 2 Ω e f f ( | T g g | + | T r r | + H . c . ) ,
H 0 = Δ 1 | 2 2 | + Ω 1 | 1 2 | + Ω 1 | 1 3 | + H . c .
H 0 = k = 2 4 Δ k 1 | k k | + Ω 1 | 1 k | + Ω 1 | k 1 | .
H 0 = Δ 1 | 2 2 | + Δ 3 | 2 2 | + ( Δ 1 + Δ 2 ) | 3 3 | + ( Δ 3 + Δ 4 ) | 3 3 | + Ω 1 | 1 2 | + Ω 1 | 1 2 | + Ω 2 | 2 3 | + Ω 2 | 2 3 | + H . c .
H 0 = Δ 1 | 2 2 | + Δ 2 | 3 3 | + Δ 3 | 3 3 | + Ω 1 | 1 2 | + Ω 2 | 2 3 | + Ω 2 | 2 3 | + H . c .
H ( t ) = { Δ 1 | 2 2 | + Δ 2 | 3 3 | + ( Ω 1 + δ Ω 1 ) | 1 2 | + ( Ω 2 + δ Ω 2 ) | 2 3 | + H . c . , t [ m T , m T + τ ) , Δ 1 | 2 2 | + Δ 2 | 3 3 | + ( Ω 1   ' + δ Ω 1 ) | 1 2 | + ( Ω 2 + δ Ω 2 ) | 2 3 | + H . c . , t [ m T + τ , ( m + 1 ) T ) .
Ω 1 ( t ) = { Ω 1 + Ω 1 Ω 1 1 + e γ mod ( t / T ) , mod ( t / T ) < τ 2 , Ω 1 + Ω 1 Ω 1 1 + e γ [ mod ( t / T ) τ ] , τ 2 mod ( t / T ) T τ 2 , Ω 1 + Ω 1 Ω 1 1 + e γ [ mod ( t / T ) T ] , mod ( t / T ) > T τ 2 .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.