10
Effective Landau-Zener transitions in circuit dynamical Casimir effect with time-varying modulation frequency A. V. Dodonov, 1, 2 B. Militello, 2, 3 A. Napoli, 2, 3 and A. Messina 2, 3 1 Institute of Physics and International Center for Condensed Matter Physics, University of Brasilia, 70910-900, Brasilia, Federal District, Brazil 2 Dipartimento di Fisica e Chimica, Universit`a degli Studi di Palermo, Via Archirafi 36, I-90123 Palermo, Italy 3 I.N.F.N., Sezione di Catania (Dated: October 22, 2018) We consider the dissipative single-qubit circuit QED architecture in which the atomic transition frequency undergoes a weak external time-modulation. For sinusoidal modulation with linearly varying frequency we derive effective Hamiltonians that resemble the Landau-Zener problem of fi- nite duration associated to a two- or multi-level systems. The corresponding off-diagonal coupling coefficients originate either from the rotating or the counter-rotating terms in the Rabi Hamiltonian, depending on the values of the modulation frequency. It is demonstrated that in the dissipation- less case one can accomplish almost complete transitions between the eigenstates of the bare Rabi Hamiltonian even for relatively short duration of the frequency sweep. To assess the experimental feasibility of our scheme we solved numerically the phenomenological and the microscopic quantum master equations in the Markovian regime at zero temperature. Both models exhibit qualitatively similar behavior and indicate that photon generation from vacuum via effective Landau-Zener tran- sitions could be implemented with the current technology on the timescales of a few microseconds. Moreover, unlike the harmonic dynamical Casimir effect implementations, our proposal does not re- quire the precise knowledge of the resonant modulation frequency to accomplish meaningful photon generation. PACS numbers: 42.50.Pq, 42.50.Ct, 42.50.Hz, 32.80-t, 03.65.Yz Keywords: nonstationary circuit QED; dynamical Casimir effect; Landau-Zener transition; Rabi Hamilto- nian; dressed-picture master equation I. INTRODUCTION The area of circuit Quantum Electrodynamics (circuit QED) offers unprecedented possibilities to manipulate in situ the properties of mesoscopic systems composed of superconducting artificial atoms interacting with the Electromagnetic field inside the waveguide resonator on a chip [1–3]. Along with practical application for Quantum Information Processing (QIP) this solid-state architec- ture also allows for the experimental study of some of the most fundamental physical processes, such as the light– matter interaction at the level of a few photons. Due to the relative ease to achieve the strong coupling regime, in which the coherent light–matter coupling rate is much larger than the system damping and dephasing rates [1, 4], this setup can probe novel phenomena with tiny coupling rates. One example is the implementation of the dynamical Casimir effect (DCE) and associated phenom- ena using actively controlled artificial atoms, which may serve both as the source and the detector of modulation- induced radiation [5–11]. We recall that DCE is a com- mon name ascribed to the processes in which photons are generated from vacuum due to the external time- variation of boundary conditions for some field [12–15]. For the usual Electromagnetic case this corresponds to the fast motion of a mirror or modulation of the dielec- tric properties of the mirror / intracavity medium [16]. Analogs of DCE were recently verified experimentally in the setups resembling a single-mirror [17] and a lossy cavity [18], where the modulation of the boundary condi- tions was achieved by threading a time-dependent mag- netic flux through SQUIDs (superconductive quantum interference devices) located inside the coplanar waveg- uide. These experiments stimulated new theoretical re- search on role of dynamical Casimir physics in quantum- information processing, quantum simulations and engi- neering of nonclassical states of light and matter [19–24]. Recent studies have indicated that DCE could be im- plemented even using a single two-level atom (qubit) with time-dependent parameters, such as the transition fre- quency or the atom–field coupling strength [5, 25–28]. Generation of excitation from vacuum occurs due to the counter-rotating terms in the Rabi Hamiltonian, which for many years had been neglected under the Rotat- ing Wave Approximation (RWA). Moreover, single-atom DCE carries some new characteristics, such as the atom– field entanglement, saturation in the number of created photons (due to intrinsic nonlinearities associated with non-harmonic spectrum of the composite system) and emergence of additional resonant modulation frequencies [27]. Since the ultra-strong coupling regime [29–32] (for which the atom–field coupling rate is comparable to the cavity and atomic frequencies) is experimentally demand- ing, especially in nonstationary configurations, here we assume moderate values of the coupling strength, which in turn imply small transition rates for the phenomena originating from the counter-rotating terms. Hence the dissipation must be sufficiently weak and the frequency of modulation must be tuned with a high precision, typi- arXiv:1603.04651v1 [quant-ph] 15 Mar 2016

arXiv:1603.04651v1 [quant-ph] 15 Mar 2016 · E ective Landau-Zener transitions in circuit dynamical Casimir e ect with time-varying modulation frequency A. V. Dodonov,1,2 B. Militello,

Embed Size (px)

Citation preview

Effective Landau-Zener transitions in circuit dynamical Casimir effect withtime-varying modulation frequency

A. V. Dodonov,1, 2 B. Militello,2, 3 A. Napoli,2, 3 and A. Messina2, 3

1Institute of Physics and International Center for Condensed Matter Physics,University of Brasilia, 70910-900, Brasilia, Federal District, Brazil

2Dipartimento di Fisica e Chimica, Universita degli Studi di Palermo, Via Archirafi 36, I-90123 Palermo, Italy3I.N.F.N., Sezione di Catania

(Dated: October 22, 2018)

We consider the dissipative single-qubit circuit QED architecture in which the atomic transitionfrequency undergoes a weak external time-modulation. For sinusoidal modulation with linearlyvarying frequency we derive effective Hamiltonians that resemble the Landau-Zener problem of fi-nite duration associated to a two- or multi-level systems. The corresponding off-diagonal couplingcoefficients originate either from the rotating or the counter-rotating terms in the Rabi Hamiltonian,depending on the values of the modulation frequency. It is demonstrated that in the dissipation-less case one can accomplish almost complete transitions between the eigenstates of the bare RabiHamiltonian even for relatively short duration of the frequency sweep. To assess the experimentalfeasibility of our scheme we solved numerically the phenomenological and the microscopic quantummaster equations in the Markovian regime at zero temperature. Both models exhibit qualitativelysimilar behavior and indicate that photon generation from vacuum via effective Landau-Zener tran-sitions could be implemented with the current technology on the timescales of a few microseconds.Moreover, unlike the harmonic dynamical Casimir effect implementations, our proposal does not re-quire the precise knowledge of the resonant modulation frequency to accomplish meaningful photongeneration.

PACS numbers: 42.50.Pq, 42.50.Ct, 42.50.Hz, 32.80-t, 03.65.YzKeywords: nonstationary circuit QED; dynamical Casimir effect; Landau-Zener transition; Rabi Hamilto-nian; dressed-picture master equation

I. INTRODUCTION

The area of circuit Quantum Electrodynamics (circuitQED) offers unprecedented possibilities to manipulatein situ the properties of mesoscopic systems composedof superconducting artificial atoms interacting with theElectromagnetic field inside the waveguide resonator on achip [1–3]. Along with practical application for QuantumInformation Processing (QIP) this solid-state architec-ture also allows for the experimental study of some of themost fundamental physical processes, such as the light–matter interaction at the level of a few photons. Due tothe relative ease to achieve the strong coupling regime, inwhich the coherent light–matter coupling rate is muchlarger than the system damping and dephasing rates[1, 4], this setup can probe novel phenomena with tinycoupling rates. One example is the implementation of thedynamical Casimir effect (DCE) and associated phenom-ena using actively controlled artificial atoms, which mayserve both as the source and the detector of modulation-induced radiation [5–11]. We recall that DCE is a com-mon name ascribed to the processes in which photonsare generated from vacuum due to the external time-variation of boundary conditions for some field [12–15].For the usual Electromagnetic case this corresponds tothe fast motion of a mirror or modulation of the dielec-tric properties of the mirror / intracavity medium [16].Analogs of DCE were recently verified experimentally inthe setups resembling a single-mirror [17] and a lossy

cavity [18], where the modulation of the boundary condi-tions was achieved by threading a time-dependent mag-netic flux through SQUIDs (superconductive quantuminterference devices) located inside the coplanar waveg-uide. These experiments stimulated new theoretical re-search on role of dynamical Casimir physics in quantum-information processing, quantum simulations and engi-neering of nonclassical states of light and matter [19–24].

Recent studies have indicated that DCE could be im-plemented even using a single two-level atom (qubit) withtime-dependent parameters, such as the transition fre-quency or the atom–field coupling strength [5, 25–28].Generation of excitation from vacuum occurs due to thecounter-rotating terms in the Rabi Hamiltonian, whichfor many years had been neglected under the Rotat-ing Wave Approximation (RWA). Moreover, single-atomDCE carries some new characteristics, such as the atom–field entanglement, saturation in the number of createdphotons (due to intrinsic nonlinearities associated withnon-harmonic spectrum of the composite system) andemergence of additional resonant modulation frequencies[27]. Since the ultra-strong coupling regime [29–32] (forwhich the atom–field coupling rate is comparable to thecavity and atomic frequencies) is experimentally demand-ing, especially in nonstationary configurations, here weassume moderate values of the coupling strength, whichin turn imply small transition rates for the phenomenaoriginating from the counter-rotating terms. Hence thedissipation must be sufficiently weak and the frequencyof modulation must be tuned with a high precision, typi-

arX

iv:1

603.

0465

1v1

[qu

ant-

ph]

15

Mar

201

6

2

cally of the order of 10−100 kHz for modulation frequen-cies in the GHz range, that ultimately should be deter-mined experimentally or numerically. For the harmonicmodulation of system parameters the observable quan-tities, e. g., the average photon number or the atomicexcitation probability, exhibit a oscillating behavior asfunction of time [5, 28]. Therefore the duration of mod-ulation must also be minutely adjusted to grasp a mean-ingful amount of excitations, since the detection is typi-cally carried out when the modulation has ceased.

In this paper we propose a simple scheme to implementthe phenomena induced by the counter-rotating termsthat does not require an accurate knowledge of the reso-nance frequencies and is little sensitive to the duration ofthe perturbation. Our method is based on the adiabaticvariation of the modulation frequency of the atomic levelsplitting through the expected resonance, when one cansteadily create excitations from vacuum and other initialstates without the aforementioned sporadic comebacks ofthe system to the initial state. This phenomenon can beunderstood in terms of some effective two- or multi-levelHamiltonians with constant nondiagonal couplings, de-pendent on the modulation depths, and time-dependentdiagonal terms, which are responsible for Landau-Zenerprocesses.

By the way, Landau-Zener transitions [33–36] are avery fundamental phenomenon, since they occur everytime a two-level system is subjected to a time-dependentHamiltonian whose bare energies change with time andcross at some instant of time. In its original form, LZmodel is characterized by bare energies changing lin-early in time. Over the years, the original scheme hasbeen extended in several directions: non linear time-dependence of the diagonal matrix elements of the Hamil-tonian [37, 38], multilevel systems, even with n-fold cross-ings [39, 40], and, of course, including environmental ef-fects inducing dissipation and decoherence [41–44]. Thisubiquitous model has been applied to describe similardynamical situations in several physical scenarios. Forexample, it has been useful in such systems as spino-rial Bose-Einstein condensates [45], Josephson junctions[46, 47], in optical lattices with cold atoms [48], andeven — in its modified version know as ‘Hidden Crossingmodel’ [49] — in classical optics [50]. Also in the contextof Circuit QED, Landau-Zener tunneling has been exten-sively used to suitably manipulate the state of a singlequbit [41, 51, 52]. In our case, Landau-Zener processesnaturally rise from the fact that the frequency sweep al-lows realizing a series of resonances between the (time-varying frequency) external perturbation and a series ofcouples of the Rabi-Hamiltonian eigenstates.

One possible downside of our method is the relativelylong implementation times. Nonetheless, we demonstratethat it can work with the current dissipative parametersprovided the modulation depth of the atomic transitionfrequency is of the order of a few percent of its bare value.Besides, we show that the states generated from vacuumcan be quite different from the squeezed vacuum state

(SVS) that is typically generated for strictly harmonicmodulations [16, 53], so our scheme may find new appli-cations in QIP.

This paper is organized as follows. In section II wededuce the effective Hamiltonians similar to the Landau-Zener physics in the resonant and dispersive regimes ofatom–field interaction. In section III we describe ourapproach to realistically account for the dissipation, withthorough numerical analysis carried out in section IV.Finally, in section V we discuss the obtained results andpresent our conclusions.

II. EFFECTIVE LANDAU-ZENER SWEEPS

We consider a single qubit in nonstationary circuitQED. Our starting point is the Hamiltonian (we set~ = 1)

HR(t) = ω0n+Ω(t)

2σz + g0

(a+ a†

)(σ+ + σ−) , (1)

where a and a† are cavity annihilation and creation op-erators and n = a†a is the photon number operator;σ+ = |e〉〈g|, σ− = |g〉〈e| and σz = |e〉〈e| − |g〉〈g| arethe atomic ladder operators, where |g〉 (|e〉) denotes theatomic ground (excited) state. ω0 is the cavity frequency,Ω is the atomic transition frequency and g0 is the atom–field coupling strength. For the sake of simplicity here weconsider the external modulation of the atomic transitionfrequency as Ω(t) = Ω0 + εΩ sin[η(t)t + φΩ], although itwas shown that for g0 ω0,Ω0 the weak modulation ofany system parameter produces similar results. In thiswork we suppose that the modulation frequency η(t) mayalso slowly change as function of time. It will be shownthat in specific regimes we obtain the two-level or multi-level effective Landau-Zener physics, so initially unpop-ulated system states can be excited without undergoingoscillations back to the initial state.

For εΩ Ω0 and εΩ . g0 we can treat the externalmodulation as a perturbation that drives the transitionsbetween the bare eigenstates of the Rabi Hamiltonian,

H(0)R |Ri〉 = Ei|Ri〉, where

H(0)R = ω0n+

Ω0

2σz + g0

(a+ a†

)(σ+ + σ−) (2)

and Ei increases with the index i. Expanding the wave-function corresponding to the Hamiltonian (1) as

|ψ(t)〉 =∑i

Ai(t)e−itEi |Ri〉 (3)

the probability amplitudes obey the coupled differentialequations

iAj(t) =εΩ sin (ηt+ φΩ)

2

∑i

Ai(t)e−it(Ei−Ej)〈Rj |σz|Ri〉.

(4)

3

Thus one can induce the coherent coupling between theRabi dressed states |Ri〉, |Rj〉 by setting the modula-tion frequency roughly equal to |Ei−Ej |. The remainingrapidly oscillating terms can be neglected according tothe RWA approach, which sets the criteria for the valid-ity of the resulting effective equations. In addition, theRWA method introduces small intrinsic frequency shiftsin the final equations, which slightly alter the resonantmodulation frequency [26, 27]. One of the advantages ofthe present scheme is that such frequency shifts are com-pletely irrelevant for the experimental implementation.

For the weak qubit–field coupling considered here,g0 ω0, we can find the approximate spectrum of theRabi Hamiltonian by performing the unitary transforma-tion [54]

UR = exp[Λ(aσ− − a†σ+

)+ ξ

(a2 − a†2

)σz], (5)

where Λ ≡ g0/∆+, ξ ≡ Λg0/2ω0 and ∆± = ω0 ± Ω0.To the first order in Λ we get the Bloch-Siegert Hamilto-nian [? ]

HBS = U†RH(0)R UR = (ω0 + δ+σz) n

+Ω0 + δ+

2σz + g0

(aσ+ + a†σ−

). (6)

Hence the approximate eigenvalues of H(0)R are

E0 = − (Ω0 + δ+) /2 (7)

En>0,± = ω0n−ω0 + δ+

2± 1

2

√[∆−(n)]2 + 4g2

0n , (8)

where δ± ≡ g20/∆±, ∆−(n) ≡ ∆−−2δ+n and the integer

n is the number of total excitations associated to HBS .The approximate eigenstates of (2) are |Ri〉 = UR|Υi〉,where |Υi〉 are the eigenstates of HBS :

|Υ0〉 = |g, 0〉 , (9)

|Υn>0,+〉 = sin θn|g, n〉+ cos θn|e, n− 1〉 , (10)

|Υn>0,−〉 = cos θn|g, n〉 − sin θn|e, n− 1〉 , (11)

θn>0 = arctan

∆−(n) +√

[∆−(n)]2 + 4g20n

2g0√n

. (12)

Contrary to the standard studies on DCE and relatedresonant effects, where the precise knowledge of spec-trum is necessary to accomplish the desired transitions[25, 27, 28, 55], in this work only an approximate knowl-edge of the spectrum is enough to achieve the phenom-ena of interest. Therefore the lowest order eigenvaluesand eigenstates derived above are sufficient for our pur-poses. The essence of our proposal is most clearly seen inspecific regimes of the system parameters, as illustratedbelow for the resonant and dispersive regimes.

A. Resonant regime

First we consider the resonant regime, ∆− = 0, andassume a small number of excitations, Λ2n 1. Forthe initial system ground state |R0〉 and the modulationfrequency

η = 2ω0 ± g0

√2− ν (t) (13)

one can show [by neglecting the rapidly oscillating termsin equation (4)] that to the lowest order in Λ the dynam-ics is described by the effective Hamiltonian

Hi = E0|R0〉〈R0|+ E2,±|R2,±〉〈R2,±| (14)

±

(ig0

√2

4

ε′Ω∆+

eit(E2,±−E0−ν(t))|R0〉〈R2,±|+ h.c.

).

Here we defined the complex modulation depth ε′Ω ≡εΩe

iφΩ and ν (t) is a small time-dependent function,|ν (t) | g0, to be specified later. This Hamiltonian canalso be obtained after cumbersome calculations using themethod of [26, 27], though in this paper it is derived justin a few lines [? ]. Performing the time-dependent uni-tary transformation

S (t) = exp

−it

[(E0 +

ν (t)

2

)|R0〉〈R0|

+

(E2,± −

ν (t)

2

)|R2,±〉〈R2,±|

](15)

we obtain the ‘interaction-picture’ effective Hamiltonian

Hf ≡ −iS†dS

dt+ S†HiS (16)

=V (t)

2(|R2,±〉〈R2,±| − |R0〉〈R0|)± (iβ|R0〉〈R2,±|+ h.c.)

V (t) ≡ ν (t) + tν (t) , β =1√2g0

ε′Ω2∆+

. (17)

Notice that this Hamiltonian only holds for the modula-tion frequency (13), and β is nonzero due to the presenceof the counter-rotating terms in the Rabi Hamiltonian(2).

When ν (t) is a linear function of time the equation(16) describes the standard two-level Landau-Zener prob-lem, so for adiabatic variation of V (t) across the avoided-crossing one can transfer steadily the population fromthe initial state |R0〉 to the final one |R2,±〉. For the lin-ear variation of V (t) from −∞ to +∞ the probability ofsuch transition is 1 − exp

(−πβ2/|ν|

), which is close to

one provided we have |ν| πβ2. However, the equation(16) is only valid for small values of |ν|, so our schemecorresponds to the finite duration Landau-Zener sweepsin which we have |ν(t)| ≤ K|β|, where K is of the orderof 10. Nonetheless, as shown later, in this way we canachieve the desired transition with sufficiently high prob-ability and compensate for the ignorance in the knowl-edge of the exact eigenvalues of the system Hamiltonian.

4

B. Dispersive regime

The dispersive regime is defined as g0√n |∆−| /2 for

all relevant values of n, and we also assume the standardcondition |∆−| ω0. Repeating the above reasoning onefinds that for the initial ground state and the modulationfrequency

η = ∆+ − 2 (δ− − δ+) + 4α− ν(t) , α ≡ g40

∆3−

(18)

the interaction-picture effective Hamiltonian reads

Hf =V (t)

2(|R2,−D〉〈R2,−D| − |R0〉〈R0|)

−D (iβ|R0〉〈R2,−D|+ h.c.) (19)

β = g0ε′Ω

2∆+, (20)

where D = ±, being the sign of ∆−/ |∆−|. Thus onecan achieve the steady population transfer from |R0〉 to|R2,−D〉, which corresponds approximately to the tran-sition |g, 0〉 → |e, 1〉. For ν(t) = 0 this behavior waspreviously named Anti-Jaynes-Cummings regime [5, 6]or the blue-sideband transition [54].

On the other hand, for

η = 2ω0 + 2 (δ− − δ+)− 4α− ν(t) (21)

we obtain an analog of the DCE Hamiltonian in the pres-ence of the Kerr nonlinearity [27]

Hf =

nmax∑n=0

[(V − 2α (n− 2)

2

)n|Rn,D〉〈Rn,D| (22)

+

(i

√(n+ 1) (n+ 2)

2β|Rn,D〉〈Rn+2,D|+ h.c.

)]

β = δ−ε′Ω√2∆+

. (23)

Here we defined |R0,D〉 ≡ |R0〉, and nmax denotes thelimiting value for the validity of the dispersive regime. If|β| & |α| we get the ‘multi-level’ Landau-Zener prob-lem, in which one can couple several dressed states|R2n,D〉 ≈ |g, 2n〉 at each avoided-crossing (although theavoided crossings for different pairs of levels do not co-incide exactly), so one can asymptotically create severalphotons from the initial vacuum state |g, 0〉.

Our approach is not limited to the initial ground state.For example, for the initial state with a finite number ofexcitations one can apply the low modulation frequency

η = |∆− − 2δ+m|+ 2 |δ−|m− 2 |α|m2 − ν(t) . (24)

For εΩ . ∆− we find the interaction-picture effectiveHamiltonian

Hf =

nmax∑n=1

[(DV − 2 (δ− − δ+) (m− n) + 2α

(m2 − n2

)2

)× (|Rn,D〉〈Rn,D| − |Rn,−D〉〈Rn,−D|)

+

(i

√n

2β|Rn,−D〉〈Rn,D|+ h.c.

)](25)

β = g0ε′(D)Ω

∆−, (26)

where ε′(+) ≡ ε′ and ε′(−) ≡ ε′∗. The nondiagonal termsin the Hamiltonian (25) rely only on the rotating termsin the Rabi Hamiltonian, so the parameter Λ does notappear in the coupling coefficient β. For εΩ

√m g0

and |ν|max . 10|β| we achieve the coupling only betweenthe states |Rm,D〉, |Rm,−D〉, that corresponds approxi-mately to the steady transition |g,m〉 → |e,m − 1〉. Forlarger variations of |ν| several dressed states may becomesuccessively coupled during the frequency sweep.

III. ACCOUNT OF DISSIPATION

For open quantum systems the dynamics must be de-scribed by the master equation (ME) for the system den-sity operator

dρ/dt = −i[HR(t), ρ] + Lρ, (27)

where L is the Liouvillian superoperator whose form de-pends on the details of system-reservoir interaction. Inthis work we do not aim to develop a microscopic ab initiomodel for dissipation in nonstationary systems, insteadwe assess whether the effective Landau-Zener transitionsdiscussed in the previous section could be implementedin real circuit QED architectures. So we use the simplestconsistent dissipation approaches available in the liter-ature to evaluate numerically the dynamics during thetimescales of interest.

We consider independent reservoirs for different pro-cesses, such as dissipation and pure dephasing. Moreover,we assume that their correlation times are much shorterthan the system relevant timescales, virtually zero, sothat we can treat the noise in the Markovian limit. Ap-plying the Davies-Spohn theory [56], we can consider thatin a given time window the bath sees the system as if itwas governed by a time-independent Hamiltonian, and ifsuch time window is larger than the typical correlationtime of the bath, then the bath acts on the system as if itwas governed by a time-independent Hamiltonian, timeinterval after time interval. If the transition frequenciesof the system are all different for any pair of eigenstatesof HR(t), which is true for the examples discussed below,

5

then at zero temperature the master equation reads [54]

LR• = D

[∑l

Φl|l〉〈l|

]•+

∑l,k 6=l

Γlkφ D [|l〉〈k|] •

+∑l,k>l

(Γlkκ + Γlkγ

)D [|l〉〈k|] • , (28)

where D[O]ρ ≡ 12 (2OρO†−O†Oρ−ρO†O) is the Lindbla-

dian superoperator and we use the shorthand notation|l〉 to denote the time-dependent eigenstates of HR(t),where the index l increases with the eigenenergy λl(t).The time-dependent parameters of equation (28) are de-fined as Φl = [γφ(0)/2]1/2σllz , Γlkφ = γφ(∆kl)|σlkz |2/2,

Γlkκ = κ(∆kl)|alk|2 and Γlkγ = γ(∆kl)|σlkx |2. Here κ($),γ($) and γφ($) are the dissipation rates correspond-ing to the resonator and qubit dampings and dephasingnoise spectral densities at frequency $; we also definedthe time-dependent quantities ∆kl ≡ λk(t)−λl(t), σlkz =〈l|σz|k〉, alk = 〈l|(a+ a†)|k〉 and σlkx = 〈l|(σ+ + σ−)|k〉.

From the lowest order approximate expressions for theeigenvalues and the eigenstates [consider equations (7)– (12)], we see that for εΩ maxg0,∆− the time-dependent eigenvalues and eigenstates are very close tothe time-independent ones evaluated at the bare qubitfrequency Ω0. In this work we assume a small modula-tion depth and g0 ω0,Ω0, hence in the ME (28) we canuse the lowest order time-independent Rabi eigenvaluesand eigenstates given by equations (7) – (12) [? ]. More-over, we do not restrict our analysis to a specific modelfor the reservoirs’ spectral densities and make the sim-plest assumption that the dissipation rates are zero for$ < 0 and take on constant values κ, γ and γφ for $ ≥ 0.Since our primary goal is to study the photon generationfrom vacuum, such assumption is the most conservativewith regards to the spurious generation of excitations dueto dephasing [54, 57, 58]. Besides, eventual random fluc-tuations in the modulation frequency may be treated asadditional dephasing noise [54, 57], so the simultaneousinclusion of κ, γ and γφ covers the most common exper-imental situations.

The numeric integration of the master equation (28)with the approximate Rabi eigenstates |Ri〉 is still cum-bersome from the practical viewpoint, so we also evaluatethe kernel LJC in which one uses the time-independentJaynes-Cummings eigenvalues and eigenstates obtainedby setting δ+ = Λ = ξ = 0 in equations (7) - (12). Itwill be shown that in our examples, where Λ ≈ 0.02,these two approaches give almost indistinguishable re-sults, which are qualitatively similar to the predictionof the phenomenological ‘standard master equation’ ofQuantum Optics (in whose microscopic derivation oneassumes that the qubit and resonator do not interact [59])

Lph• = κD[a] •+γD[σ−] •+γφ2D[σz] • (29)

with constant dissipative rates κ, γ and γφ. So the knowl-

edge of regimes in which Lph provides the same results

as LR may be important for future studies where thenumerical evaluation of LR or LJC is prohibitively com-plicated.

IV. NUMERICAL RESULTS

We verified the feasibility of photon generation fromvacuum via Landau-Zener sweeps of the modulation fre-quency in actual circuit QED setups by solving numer-ically the master equation (27) for the kernels LR, LJCand Lph. We considered the standard value ω0/2π =8 GHz for the cavity frequency, the realistic qubit–fieldcoupling strength g0/ω0 = 4 × 10−2 and the currentlyavailable dissipative rates κ = 10−4g0, γ = γφ = 7 ×10−4g0 [60–62].

FIG. 1. (Color online) Time behavior of: a) average photonnumber 〈n〉, b) Mandel’s Q-factor and c) the atomic excita-tion probability Pe in the resonant regime for the modulationfrequency η = 2ω0 +g0

√2−ν(t) and initial state |g, 0〉. Black

lines correspond to the unitary evolutions, while red, blue andgreen lines correspond to the three dissipative models char-acterized by Lph,LJC ,LR, respectively. The three dissipativemodels give almost the same results. The inset in (a) displaysthe photon statistics at the time instant βt = 10 under thedissipative kernel LR.

6

In figure 1 we exemplify the implementation of thetransition |g, 0〉 → |R2,+〉 in the resonant regime forthe parameters: ∆− = 0, εΩ = 0.01 × Ω0, η =2ω0 + g0

√2 − ν(t), where ν(t) = −8β + (β2/2)t and

β ≡ g0εΩ/(2√

2∆+). We plot the average photon number〈n〉, the Mandel’s Q-factor Q = [〈(∆n)2〉−〈n〉]/〈n〉 (thatquantifies the spread of the photon number distribution)and the atomic excitation probability Pe = Tr[|e〉〈e|ρ]. Inthe ideal case the system ends up approximately in thedressed state |R2,+〉. Under realistic conditions the pho-ton generation from vacuum persists for initial times, butlater the system decays to the ground state associated tothe kernel L because the external modulation goes off-resonance. We notice that the predictions of differentdissipation models are quite similar in this example, sofor a estimative of the time behavior one can employ thesimplest phenomenological master equation. In the insetwe show the photon statistics evaluated at the time in-terval βt = 10 for the dissipator LR, which confirms thatone or two photons could be measured with roughly 70 %probability.

In figure 2 we consider the transition |g, 0〉 → |R2,−D〉in dispersive regime for the parameters: ∆− = 9g0,εΩ = 0.04×Ω0, η = ∆+− 2 (δ− − δ+) + 4α− ν(t), whereν(t) = −8β + (β2/2)t and β ≡ g0εΩ/2∆+. Under theunitary evolution one would generate approximately thestate |e, 1〉, but the dissipation alters dramatically suchbehavior due to successive couplings |R1,D〉 → |R3,−D〉,|R2,D〉 → |R4,−D〉 (roughly |g, 1〉 → |e, 2〉 and |g, 2〉 →|e, 3〉) induced by the modulation for large times, whilethe transitions |R2,−D〉 → |R1,D〉 and |R3,−D〉 → |R2,D〉are caused by dissipation. These additional transitionsare testified by plotting the behavior of probabilitiesPe,n ≡ Tr[|e, n〉〈e, n|ρ] obtained from the kernel LJC (fig-ure 2a); they can be avoided by sweeping the modulationfrequency in the opposite direction, ν(t) → −ν(t), sincein this case the transition |R1,D〉 → |R3,−D〉 becomesoff-resonant for larger times. In the inset of 2c we showthe photon statistics obtained from the kernel LJC forthe time instant βt = 32, which proves that two photonscould be observed experimentally. Again we notice thatthe predictions of the kernels LJC and LR are almostindistinguishable, and qualitatively identical to the pre-dictions of Lph. So in the remaining of the paper we shall

only employ the dissipators LJC and Lph, as the evalu-

ation of LR becomes too demanding for a large numberof excitations.

In figures 3 – 4 we consider multiple transitions |g, 0〉 →|R2k,D〉 (k = 1, 2, . . .) in dispersive regime for the pa-rameters of figure 2 and the modulation frequency η =2ω0+2 (δ− − δ+)−4α−Sν(t), where ν(t) = [8β−(β2/2)t],

S = ± and β ≡ δ−εΩ/√

2∆+. For these parameters wehave β/α ≈ 0.9, so from the Hamiltonian (22) we expectLandau-Zener transitions among several dressed statesduring each avoided-crossing. In figure 3 we set S = +,so only the states |R2k,D〉 with k ∼ 1 may become popu-lated from the initial vacuum state, as can be seen fromthe plots. In the ideal case the atom acquires a small

FIG. 2. (Color online) Behavior of: a) 〈n〉, b) Q and c) Pe forthe coupling between the states |g, 0〉 → |R2,−D〉 in the disper-sive regime and modulation frequency η = ∆+−2 (δ− − δ+)+4α− ν(t). For the dissipator LJC we also show the probabil-ities Pe,n ≡ Tr(|e, n〉〈e, n|ρ) (a) and the photon statistics forβt = 32 [inset in (c)].

probability of excitation and up to six photons are cre-ated with non-negligible probability – this is shown in theinset of 3b, where the photon statistics is displayed forthe time instant βt = 25. Besides, the created field stateis nonclassical and quite different from the usual squeezedvacuum state created during standard DCE, as corrobo-rated by the negative values of the Q-factor (recall thatfor the SVS one has Q = 1 + 2〈n〉). In the presence ofdissipation the system tends to the ground state for largetimes, since the modulation becomes off-resonant. Stillit is possible to measure a meaningful average photonnumber 〈n〉 ∼ 2 on the timescales of a few microseconds.

When S = − one can generate a quite large amount ofphotons from vacuum in the ideal case, since the systemundergoes successive Landau-Zener transitions towardsthe higher-energy dressed states. This is illustrated infigure 4. For adiabatic variation of η only a few pho-ton states are populated at a time (see 4c for the photonstatistics at the time instants βt = 15 and 30), which isreflected in the sub-Poissonian photon statistics for large

7

FIG. 3. (Color online) Coupling of the states |g, 0〉 → |R2k,D〉in the dispersive regime for the modulation frequency η =2ω0 + 2 (δ− − δ+)− 4α− ν(t). Inset in (b): photon statisticsunder the unitary evolution for the time instant βt = 25.

times. The amount of created photons can substantiallyexceed the average photon number achievable for the har-monic modulation with constant frequency [27, 28] (greenlines in 4a and 4b, for which η was found numerically tooptimize the photon generation). Our numerical sim-ulation of dissipation is not accurate for large photonnumbers, so we only show the dissipative dynamics forβt < 21. The differences between the predictions of ker-nels LJC and Lph become significant for 〈n〉 & 4 dueto the differences in the available decay channels, butthe overall behavior is qualitatively similar. Our resultsdemonstrate that several photons can be generated onthe timescales of a few microseconds, but the dissipationmodifies the photon statistics to super-Poissonian (seethe inset in 4a, evaluated at the time instant βt = 10 forthe kernel LJC).

Finally, in figure 5 we study the transitions between thedressed states |Rn,+〉 → |Rn,−〉 in the dispersive regimefor n ≤ M , which roughly correspond to the transitions|g, n〉 → |e, n − 1〉. We consider the initial state |g, α〉,where |α〉 denotes the coherent state with α =

√4.5.

Other parameters are: ∆− = 10g0, εΩ/ω0 = 4×10−3, η =

FIG. 4. (Color online) Similar to figure 3 but for the modu-lation frequency η = 2ω0 + 2 (δ− − δ+) − 4α + ν(t). a) andb): time behavior of 〈n〉 and Q under different dissipators,where the green lines labeled ‘η = const’ correspond to har-monic modulation with constant frequency adjusted to max-imize 〈n〉. Inset in (a): photon statistics for βt = 10 underdissipator LJC . c) photon statistics under unitary evolutionfor two time instants: βt = 15 and 30.

∆− + 2M (δ− − δ+)− ν(t), where ν(t) = −7β + (β2/2)t,M = 4 and β ≡ g0εΩ/∆−. As expected, the dissipationstrongly affects the dynamics, since excitations are lostto the environment from the very beginning and the slowmodulation is unable to create additional excitations. Toconfirm the selective Landau-Zener sweeps between thestates |Rn,+〉, |Rn,−〉 in 5c and 5d we show the timeevolution of the probabilities Pg,n ≡ Tr(|g, n〉〈g, n|ρ),which change one at a time from the initial value toroughly zero (Pg,n also undergoes fast oscillations due tothe dispersive exchange of excitations between the fieldand the qubit). For n > M the probabilities Pg,n arenot affected by the perturbation, since the correspond-ing avoided-crossings are not swept during the frequencyvariation (for the frequency change in the opposite di-rections, ν(t) → −ν(t), only the states with n ≥ Mwould be affected). Although under dissipation the oc-currence of the Landau-Zener transitions is almost com-

8

FIG. 5. (Color online) Coupling of the states |Rn,+〉 → |Rn,−〉for n ≤ 4 in the dispersive regime for the initial state |g, α〉 andmodulation frequency η = ∆−+ 2M (δ− − δ+)− ν(t). a) andb): time behavior of 〈n〉 and Pe under different dissipators.c) and d): behavior of probabilities Pg,n ≡ Tr(|g, n〉〈g, n|ρ)under the unitary evolution and the kernel Lph, respectively.

pletely washed out in the behavior of 〈n〉 (figure 5a), theatomic excitation probability Pe still preserves the char-acteristic Landau-Zener plateaus, which are transformedinto peaks due to the damping (figure 5b).

V. DISCUSSION AND CONCLUSIONS

In this work we have discussed a simple scheme toachieve photon generation from vacuum due to thecounter-rotating terms in a time-dependent Rabi Hamil-tonian, where a suitable perturbation characterized by asinusoidal time dependence with a slowly changing fre-quency is present. Because of this frequency changethe perturbation intercepts several resonance frequen-cies of the system (described by the time-independentRabi Hamiltonian) and can induce a single or a seriesof Landau-Zener processes that allow for populating theupper levels starting from the ground state. This schemethen works quite well without the need to know accu-rately the resonant unperturbed frequencies of the sys-

tem nor accurately adjusting the shape of the modulationfrequency. Besides, it is only little sensitive to the dura-tion of the external perturbation. The latter property isrelated to the fact that, provided the diabatic energies(i.e., the diagonal Rabi-basis matrix elements) in the ef-fective Hamiltonians vary in a sufficiently wide range andwith a sufficiently low speed, each Landau-Zener processessentially leads to the same final result, irrespectively ofthe specific range and speed.

It is worth noting that although we considered themodulation of the atomic transition frequency, ourmethod can easily be extended to the time-variation ofthe atom-field coupling strength, or both. Moreover, theeffective Hamiltonians deduced here are valid for arbi-trary variation of the modulation frequency. So our re-sults can be employed to propose protocols that optimize,for instance, the photon production or generation of spe-cific entangled states, although in these cases one wouldneed a precise knowledge of the system spectrum and tocontrol accurately the duration of the perturbation.

We showed that the effective Landau-Zener transitionscould be implemented for current parameters of dissipa-tion provided one maintains the modulation for a time in-terval of the order of microseconds with relative modula-tion strength of a few percent. For consistent descriptionof the experimental results the dissipation must be neces-sarily included because it alters significantly the unitarybehavior due to the additional transitions between thesystem eigenstates induced by the combined action of dis-sipation and coherent perturbation. It is remarkable thatthe predictions of the phenomenological quantum opticalmaster equation are qualitatively similar to the predic-tions of the microscopic dressed-picture master equation,and in many situations both predictions are almost indis-tinguishable (this is typical in the weak damping limit).This means that one can use the simpler phenomenolog-ical master equation to get a crude estimation about theoverall behavior. The effect of random fluctuations ofthe modulation frequency can be incorporated into ourapproach as additional pure dephasing, so one does notexpect major qualitative differences in this case. There-fore we hope that our protocol will facilitate the hithertomissing experimental observation of DCE due to a singlequbit.

ACKNOWLEDGMENTS

A.V.D. would like to thank Dipartimento di Fisica eChimica of Universita degli studi di Palermo for the warmhospitality during the visit in winter of 2016. A.V.D.also acknowledges a support of Brazilian agencies CNPq(Conselho Nacional de Desenvolvimento Cientıfico e Tec-nologico) and FAPDF (Fundacao de Apoio a Pesquisa doDistrito Federal).

9

[1] A. Wallraff, D. I. Schuster, A. Blais, L. Frunzio, R. -S.Huang, J. Majer, S. Kumar1, S. M. Girvin and R. J.Schoelkopf, Nature 431, 162 (2004).

[2] R. J. Schoelkopf and S. M. Girvin, Nature 451, 664(2008).

[3] J. Q. You and F. Nori, Nature 474, 589 (2011).[4] D. I. Schuster, A. A. Houck, J. A. Schreier, A. Wallraff,

J. M. Gambetta, A. Blais, L. Frunzio, J. Majer, B. John-son, M. H. Devoret, S. M. Girvin, and R. J. Schoelkopf,Nature 445, 515 (2007).

[5] A. V. Dodonov, J. Phys.: Conf. Ser. 161, 012029 (2009).[6] A. V. Dodonov, R. Lo Nardo, R. Migliore, A. Messina,

and V. V. Dodonov, J. Phys. B 44, 225502 (2011).[7] A. V. Dodonov and V. V. Dodonov, Phys. Lett. A 375,

4261 (2011).[8] A. V. Dodonov and V. V. Dodonov, Phys. Rev. A 85,

015805 (2012).[9] A. V. Dodonov and V. V. Dodonov, Phys. Rev. A 85,

055805 (2012).[10] A. V. Dodonov and V. V. Dodonov, Phys. Rev. A 85,

063804 (2012).[11] A. V. Dodonov and V. V. Dodonov, Phys. Rev. A 86,

015801 (2012).[12] V. V. Dodonov, Adv. Chem. Phys. 119, 309 (2001).[13] V. V. Dodonov, Phys. Scr. 82, 038105 (2010).[14] D. A. R. Dalvit, P. A. Maia Neto, and F. D. Mazzitelli, in

Casimir Physics (Lecture Notes in Physics 834), editedby D. Dalvit, P. Milonni, D. Roberts, and F. da Rosa(Springer, Berlin, 2011), p. 419.

[15] P. D. Nation, J. R. Johansson, M. P. Blencowe, and F.Nori, Rev. Mod. Phys. 84, 1 (2012).

[16] C. K. Law, Phys. Rev. A 49, 433 (1994).[17] C. M. Wilson, G. Johansson, A. Pourkabirian, M.

Simoen, J. R. Johansson, T. Duty, F. Nori, and P. Dels-ing, Nature 479, 376 (2011).

[18] P. Lahteenmaki, G. S. Paraoanu, J. Hassel, and P. J.Hakonen, Proc. Nat. Acad. Sci. 110, 4234 (2013).

[19] S. Felicetti, C. Sabın, I. Fuentes, L. Lamata, G. Romero,and E. Solano, Phys. Rev. B 92, 064501 (2015).

[20] G. Benenti, A. D’Arrigo, S. Siccardi, and G. Strini, Phys.Rev. A 90, 052313 (2014).

[21] J. R. Johansson, G. Johansson, C. M. Wilson, P. Delsing,and F. Nori, Phys. Rev. A 87, 043804 (2013).

[22] S. Felicetti, M. Sanz, L. Lamata, G. Romero, G. Johans-son, P. Delsing, and E. Solano, Phys. Rev. Lett. 113,093602 (2014).

[23] R. Stassi, S. De Liberato, L. Garziano, B. Spagnolo, andS. Savasta, Phys. Rev. A 92, 013830 (2015).

[24] D. Z. Rossatto, S. Felicetti, H. Eneriz, E. Rico, M. Sanz,and E. Solano, arXiv:1511.03825.

[25] S. De Liberato, D. Gerace, I. Carusotto, and C. Ciuti,Phys. Rev. A 80, 053810 (2009).

[26] A. V. Dodonov, J. Phys. A 47, 285303 (2014).[27] I. M. de Sousa and A. V. Dodonov, J. Phys. A 48, 245302

(2015).[28] D. S. Veloso and A. V. Dodonov, J. Phys. B 48, 165503

(2015).[29] M. Devoret, S. Girvin and R. Schoelkopf, Ann. Phys. 16,

767 (2007).[30] B. Peropadre, P. Forn-Dıaz, E. Solano, and J. J. Garcıa-

Ripoll, Phys. Rev. Lett. 105, 023601 (2010).

[31] J. Casanova, G. Romero, I. Lizuain, J. J. Garcıa-Ripoll,and E. Solano, Phys. Rev. Lett. 105, 263603 (2010).

[32] R. Stassi, A. Ridolfo, O. Di Stefano, M. J. Hartmann,and S. Savasta, Phys. Rev. Lett. 110, 243601 (2013).

[33] L. D. Landau, Phys. Z. Sowjetunion 2, 46 (1932).[34] C. Zener, Proc. R. Soc. Lond. Ser. A 137, 696 (1932).[35] E. Majorana, Nuovo Cimento 9, 43 (1932).[36] E. C. G. Stueckelberg, Helv. Phys. Acta 5, 369 (1932).[37] N. V. Vitanov and K. -A. Suominen, Phys. Rev. A 59,

4580 (1999).[38] T. Chasseur, L. S. Theis, Y. R. Sanders, D. J. Egger, and

F. K. Wilhelm, Phys. Rev. A 91, 043421 (2015).[39] A. V. Shytov, Phys. Rev. A 70, 052708 (2004).[40] C. E. Carroll and F. T. Hioe, J. Phys. A: Math. Gen. 19,

2061 (1986); C. E. Carroll and F. T. Hioe, J. Phys. A:Math. Gen. 19, 1151 (1986).

[41] D. Zueco, P. Hanggi and S. Kohler, New J. Phys. 10,115012 (2008).

[42] K. Saito, M. Wubs, S. Kohler, Y. Kayanuma, and P.Hanggi, Phys. Rev. B 75, 214308 (2007).

[43] P. Ao and J. Rammer, Phys. Rev. B 43, 5397 (1991).[44] P. Ao and J. Rammer, Phys. Rev. Lett. 62, 3004 (1989).[45] J. -N. Zhang, C.-P. Sun, S. Yi, and F. Nori, Phys. Rev.

A 83, 033614 (2011).[46] D. M. Berns, M. S. Rudner, S. O. Valenzuela, K. K.

Berggren, W. D. Oliver, L. S. Levitov, and T. P. Orlando,Nature (London) 455, 51 (2008).

[47] G. Sun, X. Wen, B. Mao, J. Chen, Y. Yu, P. Wu, and S.Han, Nature Commun. 1, 51 (2010).

[48] A. Zenesini, D. Ciampini, O. Morsch, and E. Arimondo,Phys. Rev. A 82, 065601 (2010).

[49] S. Fishman, K. Mullen and E. Ben-Jacob, Phys. Rev. A42, 5181 (1990).

[50] D. Bouwmeester, N. H. Dekker, F. E. v. Dorsselear, C.A. Schrama, P. M. Visser and J. P. Woerdman, Phys.Rev. A 51, 646 (1995).

[51] K. Saito, M. Wubs, S. Kohler, P. Hanggi, and Y.Kayanuma, Europhys. Lett. 76, 22 (2006).

[52] Li Jun-Wang, Wu Chun-Wang, Dai Hong-Yi, Chin. Phys.Lett. 28, 090302 (2011).

[53] V. V. Dodonov and A. B. Klimov, Phys. Rev. A 53, 2664(1996).

[54] F. Beaudoin, J. M. Gambetta and A. Blais, Phys. Rev.A 84, 043832 (2011).

[55] G. Vacanti, S. Pugnetti, N. Didier, M. Paternostro, G.M. Palma, R. Fazio, and V. Vedral, Phys. Rev. Lett. 108,093603 (2012).

[56] E. B. Davies and H. Spohn, J. Stat. Phys. 19, 511 (1978).[57] T. Werlang, A. V. Dodonov, E. I. Duzzioni, and C. J.

Villas-Boas, Phys. Rev. A 78, 053805 (2008).[58] A. V. Dodonov, Phys. Scr. 86, 025405 (2012).[59] W. Vogel and D. -G. Welsch, Quantum Optics (Wiley,

Berlin, 2006).[60] G. Kirchmair, B. Vlastakis, Z. Leghtas, S. E. Nigg, H.

Paik, E. Ginossar, M. Mirrahimi, L. Frunzio, S. M.Girvin, and R. J. Schoelkopf, Nature 495, 205 (2013).

[61] D. Riste, M. Dukalski, C. A. Watson, G. de Lange, M.J. Tiggelman, Ya. M. Blanter, K. W. Lehnert, R. N.Schouten, and L. DiCarlo, Nature 502, 350 (2013).

[62] L. Sun, A. Petrenko, Z. Leghtas, B. Vlastakis, G. Kirch-mair, K. M. Sliwa, A. Narla, M. Hatridge, S. Shankar, J.

10

Blumoff, L. Frunzio, M. Mirrahimi, M. H. Devoret, andR. J. Schoelkopf, Nature 511, 444 (2014).