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

Macroscopically distinct superposition in a spin ensemble coupled to superconducting flux-qubits

Open Access Open Access

Abstract

Large optical nonlinearities can create fancy physics, such as big Schrödinger-cat states and quadrature squeezing. We present the possibility to practically generate macroscopic Schrödinger-cat states, based on a giant Kerr nonlinearity, in a diamond nitrogen-vacancy ensemble interacting with two coupled flux-qubits. The nonlinearity comes from a four-level N-type configuration formed by two coupled flux-qubits under the appropriately driving fields. We discuss the experimental feasibility in the presence of system dissipations using current laboratory technology and our scheme can be easily extended to other ensemble systems.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Over past decades, Schrödinger cat [1] has been the unique example of a paradoxical but fundamental property of quantum mechanics, reflecting that quantum superposition is at the core of quantum theory. Despite experimental evidences in individual atoms [2], photons [3], and molecules [4], this physicists’ favorite feline remains purely hypothetical if its size becomes larger, because objects beyond microscopic regime generally interact strongly with their surroundings, which force them away from superposition.

Great efforts have been devoted to exploring superposition states in mesoscopical systems, e.g., Schrödinger-cat states with six to ten qubits [5, 6], and to generating Schrödinger-cat states in an ultrafast fashion [7]. It is still strongly expected to observe quantum superposition in macroscopic systems for further understanding the peculiar laws in quantum world. Recently, control of vibrational degrees of freedom has been achieved in optomechanics [8] and nanomechanics [9]. Nevertheless, generation of macroscopically distinct superposition states, which are extremely sensitive to decoherence, is still experimentally challenging [10, 11].

In this work, we propose a practical scheme to create macroscopically distinct superposition states, i.e., macroscopic Schrödinger-cat states, in a diamond nitrogen-vacancy ensemble (NVE) coupled to two interacting flux-qubits. The NVE is a promising candidate for quantum information processing, which owns the capability from scalable quantum computing to long-distance quantum information transfer [12–14]. As the other component of our model, the flux-qubit is constructed by a compound Josephson-junction (CJJ) rf-superconducting quantum interference device (rf-SQUID) and we consider two such flux-qubits to be coupled by a coupler [15, 16]. There have been some schemes proposed so far to couple the superconducting flux-qubits to the spin ensembles by a magnetic field [17–20], and also some experimental observations for large couplings between the flux-qubits and the NVEs [21–23]. In contrast, our scheme works based on a giant Kerr nonlinearity appearing in the strong coupling regime of the system under a weak driving field, which is relevant to an N-type level configuration (i.e. the transitions between any two of the levels in the shape of the letter N) [24–27] of the coupled flux-qubits. Compared with the previous method for generating the macroscopic entanglement of NVE [28], the initialization of the quantum data bus into macroscopic cat state is not required in our scheme. Due to the giant Kerr nonlinearity, distinct superposition states could be generated in the NVE within a very short time. The idea can also be applied to generation of large quadrature squeezing states in the NVE and extended to other ensemble systems.

2. Model and Hamiltonian

Consider a superconducting diamond hybrid system as shown in Fig. 1(a) where an NVE, involving M NV-centers, is placed above the two flux-qubits. The latter is of Ising spin glass architecture [15, 16] with coupled CJJ rf-SQUIDs.The system experiences an external magnetic field Bqj generated by the circularly persistent current of the jth flux-qubit that is in parallel with the direction of Sx of the NVE (defined below). Under the co-resonance conditions [29], the system is described in units of =1 by the Hamiltonian as below,

H=Ωq2i=12σxi+Jσz1σz2+i=1MDSz,i2+i=1MSx,i(g1σz1+g2σz2),
where σx,zi are the Pauli spin operators of the ith flux-qubits, and Sk,i (k=x,z) is the spin-1 operator of the ith NV-center with the eigenstates |±1 and |0 [see inset of Fig. 1(a)]. The parameters Ωq and J are the tunnelling energy of the flux-qubit and the coupling strength between the flux-qubits, respectively. Besides, D=2.88 GHz is the zero-field splitting constant regarding the NV-center ground state [21, 30]. For our purpose, the magnetic field Bqj generated by the circularly persistent current of the jth flux-qubit is assumed in parallel with the direction of Sx of the NV-center and in this case the coupled term Sxσzj is dominant. The coupling strengths g1 and g2 between the NVE and the flux-qubits are determined by the persistent current of the flux-qubits and the distance between the flux-qubits and the NVE [21]. A detailed description of the model can be found in Appendix A.

The first two terms in Eq. (1) denoting the flux-qubits energy, as plotted in Fig. 1(b), can be expressed by the N-type level configuration with the eigenenergies [E4,E3,E2,E1]=[J2+Ωq2,J,J,J2+Ωq2]. Transforming H into a rotating frame through Hrot=eiRtHeiRt+R with R=D|11|+(Dω¯)|22|ω¯|44|+Di=1M|0ii0|, the effective Hamiltonian under rotating-wave approximation is

H1=Δ|44|+δ|33|+i=1M[|bii0|(g¯|13|+g˜|24|)+h.c.],
where, the dressed state |b=(|++|)/2, the effective coupling strengths g¯=(g1+g2)sin θ and g˜=(g1g2)sin θ with θ=arccos (J/J2+Ωq2)/2. Δ and δ represent the detunings regarding the states |4 and |3, respectively. Ignoring the impact caused by the strain-induced splitting of the NVE, both the two detunings are equal to ωD. Denoting S¯+=i=1M|bii0|, S¯=i=1M|0iib| and S¯z=i=1M[|biib||0ii0|], we map the collective spin operator into the bosonic operator by the Holstein-Primakoff transformation [31] in the low excitation condition, that is, S¯+=aMaaMa, S¯=MaaaMa and S¯z=2aaM. Besides, two classical fields are applied to the system with one coupling |± with |0 as a probe field to resonantly drive the NVE, and the other coupling |2 with |3 to form the N-type level structure. Thus the final Hamiltonian turns to be
H2=H0+a(λ¯|31|+λ˜|42|)+h.c.,
where H0=Δ|44|+δ|33|+Ω(|23|+|32|)+ϵp(a+a) with ϵp=Mξp, λ¯=Mg¯sin θ and λ˜=Mg˜sinθ.

 figure: Fig. 1

Fig. 1 (a) Schematic diagram of the hybrid quantum system composed of two coupled CJJ rf-SQUID qubits and an NVE. Inset: the level structure of an NV-center [30] under a weak classical field with an amplitude ξp coupling the states |± to |0. (b) The N-type level structure for the coupled flux-qubits with ω=J+J2+Ωq2 and ω¯=J2+Ωq2J. A control field with Rabi frequency Ω and frequency ωc=2J is in resonance with the states |2 and |3. (c) Three-dimensional plot of log10[g2(0)] as functions of Ω and ϵp. The inset denotes the nonlinearity factor χ varying with the amplitude ϵp of the driving field along the gradient direction (red curve). Here the coupling strengths λ¯/κ=400 and λ˜/κ=200. The decay paths are set as γ41=γ32=0.1γ and γ31=γ21=γ42=γ43=γ. Where, γijrepresent the decay rate from the states |i to |j. Other parameters are Δ=δ=κ and γ/κ=0.4.

Download Full Size | PDF

3. Kerr nonlinearity

Due to no selection rule in the flux-qubits, all the decay paths are possible in the N-type level structure. As such, the Hamiltonian in Eq. (3) can be reduced to an effective self-Kerr nonlinearity interaction form Heff=χa2a2 [27, 29]. By adiabatically eliminating the flux-qubit operator [27, 32], the nonlinearity factor χ is given by χ=λ¯2/Ω2[λ˜2Δ/(γ422+Δ2)λ¯2δ/[(γ31+γ32)2+δ2]], which generally holds in the weak coupling limit λ¯2/Ω21 [29] (the detail provided in Appendix A). However, when the system has all operational decay paths, a very large nonlinearity is possible [29] and observable by measuring the second-order correlation function g(2)(τ)=a(t)a(t+τ)a(t+τ)a(t)/|a(t)a(t)|2. Under the weak driving field, the effective self-Kerr nonlinearity χ can be obtained from the analytical result of g(2)(0) [29, 33] by solving ρ˙=0 from the master equation [33],

ρ˙=i[H,ρ]+κ2D[a,ρ]+j<kγj,k2D[σjk,ρ],
where D[A,ρ]=2AρAAAρρAA, κ is the decay rate regarding the bosons, σjk=|jk| and γj,k is the decay rate from the states |k to |j. Fig. 1(c) presents the values of g(2)(0) under the condition of weak driven field, where χ/κ103 is achieved from our calculation. This implies that with an appropriate strength of the control field we can create a huge self-Kerr nonlinearity. The maximal nonlinearity exists along the red curve in Fig. 1(c), where the nonlinearity decreases exponentially with the increase of ϵp due to the fact that the strong drive causes unexpected excitations of the bosons spoiling the nonlinearity term a2a2. In addition, the self-Kerr nonlinearity can also be used to create quadrature squeezing [34, 35]. As detailed in Appendix B, if |ε|2κ2/3 where ε=2iχTr[aaρss] with ρss denoting the steady state solution of Eq. (4), the squeezing spectrum has a single peak at ω=0 with the maximum squeezing S(0)=1/6, and otherwise, it splits into two peaks at ω2=3|ε|2κ2 with the same squeezing maxima.

 figure: Fig. 2

Fig. 2 (a) Time evolution of the fidelity F. (b) Wigner function W(β) at five different time points corresponding to t1, t2, t3, t4 and t5 in (a). (c) Probability distribution of the rotated quadrature operator PX at time points t3 (blue solid curve) and t4 (black dashed curve). Here we choose κ=1, γ=0.4, Δ=δ=1, α=2eiπ/4, ϵp=2, λ˜=300, Ω=50 and decay paths are chosen same as in Fig. 1(c).

Download Full Size | PDF

4. Macroscopic superposition states

Using the self-Kerr nonlinearity, we focus on generating coherent state superpositions of the NVE, such as the Schrödinger-cat state |cat=(eiπ/4|iα+eiπ/4|iα)/2 with the amplitude α [36]. To this end, we start from preparing a coherent state |α of the NVE by the displacement operator D(α)=eαaα*a, and then the superposition of coherent states is created under the government of the effective self-Kerr nonlinearity Heff (see Appendix C for detail). Using Eq. (4), we have evaluated the produced state in the presence of dissipations by the Uhlmann fidelity F=cat|ρcat|cat, as shown in Fig. 2(a). The fidelity oscillates in time and reaches the maximum 0.88 within a very short time. As shown later, the fidelity would not be unit even in absence of any decay effect, which is due to unexpected effects beyond nonlinearity from Heff under the weak decay condition. So we call below the maximum fidelity we could reach as the optimal fidelity, and the time reaching the optimal fidelity as the optimal time τop.

In order to scrutinize the quantum coherence and interference effects during the cat creating process, we have examined the Wigner function W(β)=(2/π)Tr[D(β)ρ^(t)D(β)P] with P=(1)aa [37], as plotted in Fig. 2(b), where the interference patterns implies the coherent superposition of the states. Besides, we also employ the probability distribution PX=eX(θ)|ρe|X(θ)e to portray quantum superposition properties teprl-116-163602,qo, where |X(θ)e is the eigenstate corresponding to the rotated quadrature operator X^(θ)=(aeiθ+aeiθ)/2. In order to maximize the interference, we choose the rotation angle to be θ0=arg [α]. The oscillation in both curves in Fig. 2(c) indicates the quantum interference between the superposition components. In addition, we have also noticed from the values that with the position of the maximal PX approaching X(θ0)=0, the superposition state is closer to the state |cat and quantum interference becomes stronger.

 figure: Fig. 3

Fig. 3 (a) Optimal fidelity of the cat state as a function of the cat’s size. Inset: the dots denote the oscillating period of the fidelity for |α|25 and the line means a constant obtained by fitting these dots. (b) Optimal time ctτop in variation with the cat’s size. Inset: the dots represents Δτop between the nearest-neighbor ladders and the curve denotes an exponential fitting. (c) Dynamical evolution of the fidelity for three sizes of the cat states around the inflection point |α|2=5.05, which is labeled as a red dot in (a). (d) Probability distribution of the rotated quadrature operator PX at time points rotectκτop=0.02 (|α|2=4.8,5) and 0.025 (|α|2=5.1,5.8). Here the parameters used are the same as in Fig. 2.

Download Full Size | PDF

In general, the fidelity of the prepared optimal cat state is sensitive to its size under the dissipative condition. In contrast to a simple asymptotic decay as usual, the decay in our case shows a subtle behavior (Fig. 3(a)). To understand the physics behind the fidelity, we define τop as the time reaching the optimal fidelity of the cat state. We see in Fig. 3(a) that the fidelity decreases rapidly when |α|2<2 and τop decreases accordingly [see Fig. 3(b)]. Then the fidelity has a small rise, followed by a large descent until |α|25, during which τop remains almost unchanged. For |α|25, the fidelity behaves as a periodic oscillation in a linear decay and the corresponding τop increases discretely as stairs. Quantitatively, the laws of the changes for the case of |α|25 can be described as (some details can be seen in Appendix D),

Δ|α|2C,Δτop=a1ea2n+a3,
where n denotes the nth oscillating period, the oscillation length Δ|α|2 and the optimal time spacing Δτop are defined as in Fig. 3. C and a1,2,3 are constants related to the parameters of the system (for the parameters in Fig. 3, they are corresponding to C=1.92, a1=3.4×103, a2=0.26 and a3=1.96×103), but independent of the dissipation. Moreover, Fig. 3(c) shows an inflection point at |α|2=5.05, for which the fidelity of the cat state presents double peaks with the same maxima. For any deviation from this inflection point, the fidelity only owns a single maximum value. Since this change is discontinuous, one may observe the discrete increase of τop with the jump Δτop for the stairs as in Fig. 3(b). The oscillation of curves of PX in Fig. 3(d) also demonstrates distinct quantum interferences between the superposition components. The values around the inflection point present approximate symmetry with respect to X(θ0)=0, implying a π rotation about the original point in phase space.

 figure: Fig. 4

Fig. 4 (a) and (b) Fidelities of the cat state prepared under different flux-qubit dissipation rates and NVE decay rates, where we set u0/2π=0.1 MHz, and for convenience of presentation the parameters below are written in units of u0. We have κ=1 in (a) and γ=1 in (b). The solid, dashed and dotted curves in both (a) and (b) correspond to λ˜=300,200 and 100, respectively. Inset: time evolution of the fidelity with tλ˜=300. (c) Probability distribution of the rotated quadrature operator PX at the time point t4 of Fig. 2. The dotted, dashed and solid curves correspond to the excitation numbers nc=1, 2 and 5. In (c), κ=1 and γ=0.4. Inset: Wigner function W(β) with respect to nc=1 and 5. Other parameters in (a-c) are pha=2eiπ/4, ϵp=2, λ¯=400 and Ω=50, respectively.

Download Full Size | PDF

5. The influence from decoherence and dissipation

To check more seriously the influence from the system dissipations, we have numerically investigated the fidelity variation of the prepared cat states with respect to the system decay. As shown in Figs. 4(a) and 4(b), the fidelity of the cat state is damaged by both the flux-qubit dissipation and the NVE decay. As a comparison, one can find that under the same condition the decay of the NVE bring a more serious dissipation than that of flux-qubit due to the fact that the Schrödinger-cat state is prepared in the NVE and thus the NVE decay directly leads to the fidelity decrease. In contrast, the flux-qubit dissipation is associated with an indirect coupling between the flux-qubit and the NVE.

Moreover, considering some imperfect factors in the NVE, we assume the NVE under a finite-temperature reservoir, which modifies the second term in the right-hand side of Eq. (4) as κ(nc+1)D[a,ρ]/2+κncD[aer,ρ]/2 with nc representing the boson number regarding the noise of the reservoir. To explore this imperfection, we employ the Wigner function and the probability PX in Fig. 4(c). The probability PX demonstrates less evident fringes with increase of the reservoir noise, although quantum interference of the coherent superposition states is still distinct even in the case of nc=5. Besides, the Wigner function presents a different view for the noise influence on the cat state that with the increase of the noise, the two components |iα and |iα themselves turn to be more dominant than the interference in between, i.e., the darker ends for the larger nc [inset of Fig. 4(c)]. This is the reason for the reduced fidelity of the cat state in the more noisy case. Nevertheless, as long as the reservoir noise is not big enough, quantum interference pattern remains existing.

 figure: Fig. 5

Fig. 5 (a) Fidelities of the cat state prepared under different dephasing rates κd. (b) A detail dynamical evolution of the fidelities for three different dephasing rates at a coupling λ˜=300. (c) The Wigner function W(β) atthe optimal cat state position of in (b) for the case κd=0.5. Here κ=1 and γ=0.4 and the other parameters are same as in Fig. 4.

Download Full Size | PDF

In addition, inhomogeneous broadening is another key problem in manipulation of the NVEs, which is caused by magnetic dipolar interactions with the nuclear or excess electron spins in diamond [21, 22] and results in the dephasing effects [19]. Previous studies have found that the magnetic field applied along a special direction [100] of each NV-center can largely suppress the inhomogeneous broadening of the NVE [23]. In our case for effectively coupling two flux qubits via the NVE, we should also consider the inhomogeneous broadening resulted from spatial distribution of the NV-centers, which diversifies the coupling strength g1,2, is very small and thus could be resorted to dephasing, following the solution in [19]. Under the weak field approximation, the dephasing rate κd of the NVE is related to the inhomogeneous width γs of the NVE by the relation γs=2κd. Therefore, a dephasing term κdD[aa,ρ]/2 should be complemented into the master equation (at the right-hand side of Eq. (4)) to quantitatively describe the influence from the inhomogeneous broadening effect. Fig. 5 is the result of numerical simulation of such a master equation.Under the dephasing effect, the interference are still clearly visible in Fig. 5(c). Specially, in the case of (λ˜=300κ), a small inhomogeneous broadening does not distinctly damage the prepared macroscopic superposition states.

6. Discussion and conclusion

Fast generation and long-time remaining of Schrödinger-cat states are essential to deeper understanding of quantum theory and also to some practical applications. Since they can be measured and characterized experimentally by the quantum spectroscopy [38], the cat states can be employed in the quantum metrology, for example, as a sensitive electrometer [39], for improving the measurement accuracy of phases [40], for engineering a remote amplifier [41] and for other tasks of quantum metrology [42]. In addition, NVEs are practical for scalable quantum information processing, such as used as a magnetic field transducer [43], measuring the oscillating field [44], applied as a magnetometer [45] and nanometre positioning meter [46], detecting the molecular radical reaction [47] and a single nuclear spin [48]. As a result, studying Schrödinger-cat states in the NVEs would help keeping quantum coherence in the NVEs when they work as qubits in quantum computing or as nodes in quantum network.

On the other hand, experimentally detecting the Schrödinger-cat states in spin ensembles is not a trivial task. We can visualize the desired Schrödinger cat state by measuring the Wigner function of the NVE. Earlier experimental works have shown mature techniques to determine the density matrix of the resonators and to directly measure Wigner function in ion traps [49] as well as in microwave superconducting circuits [36, 50]. In general, the Wigner function of Schrödinger-cat states could be determined by the parity of the state [51], by the statistical moments of the field operator [52], or by reconstructing the density matrix of the system using a least square fit to each Husimi Q-function [53]. In our case, based on the self-Kerr nonlinearity, we displace the state of the NVE by a displacement pulse D(α) and wait for a variable evolution time. Then, by means of state tomography, we may employ the least-square fit to each Husimi Q-function for reconstructing the density matrix of the NVE, and then obtain the Wigner function by straightforward calculation.

Our scheme is feasible using current laboratory techniques [54, 55]. The available dissipation values are κ/2π=0.1 MHz for the NVE [22] and γ/2π[0,100] kHz for this coupled flux-qubit four-level system [21, 56, 57]. For other parameters, the tunneling energy Ωq/2π of the flux-qubit is from 1 MHz to 10 GHz [15]. The coupling energy J is related to the effective mutual inductance M and the qubit persistent current, implying that J/2π= 2.3 GHz is available if M=1.5 pH and |Iip|=1μA [15]. With these values, the level spacing ω, as defined in Fig. 1(b), could match the zero-field splitting D of the NV-center ground state. The coupling strength between the flux-qubit and the NVE is regarding the flux-qubit persistent current strengths and the number of the NV-center, which could reach 116 MHz (about λ¯,λ˜/κ103) for |Iip|= 1 μA and 3.2×107 NV-centers. As such, the giant effective self-Kerr nonlinearity χ/κ could be 103 under a weak driving field, implying that we can create a Schrödinger-cat state within a time about 200 ns.

In conclusion, we have proposed a practical scheme to create macroscopic Schrödinger-cat states in a hybrid system with an NVE interacting with two coupled CJJ-flux qubits. The key point of our work is to try to generate Schrödinger-cat states in a fast and decoherence-suppressed fashion. Numerical results have demonstrated that our scheme works well in a wide parameter range in the presence of real dissipations of the system. The idea can also be used to demonstrate quadrature squeezing and extended to other ensemble systems.

Appendix A establish the model and construct the Hamiltonian of system

1. Constructing a four-level system

The compound Josephson-junction (CJJ) rf-superconducting quantum interference device (rf-SQUID) is employed as a qubit and two such qubits are coupled by a CJJ rf-SQUID coupler (see Fig. 1). The Hamiltonian for this architecture is a quantum Ising spin glass [15, 16] that is written as Hs=i=1212[ωqiσzi+Ωqiσxi]+Jσz1σz2, where ωqi/2π=2|Iip|Φix and Ωqi are the bias and tunneling strength of the ith flux-qubit, respectively, and J=M|I1p||I2p| is the coupling strength. Here |Iip| is the magnitude of the qubit persistent current, Φix and M denote the external flux bias and mutual inductance, and they are in situ tunable. σzi and σxi are Pauli spin operators. In addition, if the CJJ rf-SQUIDs are identical and the individual tunneling energy can be modulated by local flux tuning at the co-resonance point ωqi=0 [29], we set Ωq1=Ωq2=Ωq and rewrite the Hamiltonian Hs as Hs'=Ωq2i=12σxi+Jσz1σz2. Thus, the eigenenergies of the Hamiltonian Hs' can be solved straightforwardly as [E4,E3,E2,E1]=[J2+Ωq2,J,J,J2+Ωq2], with the corresponding eigenstates denoted by |4, |3a¯ngle, |2 and |1. Defining an angle cos 2θ=J/J2+Ωq2, the above four states are presented as

(|4|3|2|1)=12(cos θsin θsin θcos θ10010110sin θcos θcos θsin θ)(|ee|eg|ge|gg).
Defining the transition frequencies as ωi,j=EiEj, thus we have ω32=2J and ω42=ω31=ω=J+J2+Ωq2 with Kω42/ω31=1. Here we mention the exception, i.e., the case of nonzero ωqi for which the eigenenergies are solved by moving slightly off the co-resonance point through introducing equal-strength Zeeman terms in Hs', i.e., Hs''=12i=12ωqiσzi+Hs' [29]. For this case, K is no longer equal to one and a slight difference will be introduced between ω42 and ω31. Nevertheless, this case does not change the physical essence of our model. As a result, we will proceed following Eq. (6), that is,
(|ee|eg|ge|gg)=12(cos θ10sin θsin θ01cos θsin θ01cos θcos θ10sin θ)(|4|3|2|1).

Eq. (7) implies,

σz1=cos θ|43|+sin θ|42|+sin θ|31|cos θ|21|+h.c.,σz2=cos θ|43|sin θ|42|+sin θ|31|+cos θ|21|+h.c

2. Coupling the four-level system to the NVE

The Hamiltonian for the NVE is given by HNVE=i=1M[DSz,i2+E(Sx,i2Sy,i2)+geμBBz,iSz,i], where ge=2.0028, μB=14 MHzmT1, and D=2.88 GHz is the zero-field splitting (E<10 MHz) [21, 30]. The Zeeman interaction Bz,iSz,i and strain-induced splitting term E(Sx,i2Sy,i2) are negligible. Sk,i (k=x,y,z) are the spin-1 operators of the NV-center with the eigenstates |± and |0. The magnetic field generated by the circularly persistent current of the jth flux-qubit is assumed to be in parallel with the direction of Sx.

In addition, the coupling between the NVE and the flux-qubits is written as Hint=i=1MSx,i(g1σz1+g2σz2), where the coupling strength gk can be estimated by gk=geμBBqk/2, Bqk=μ0Ikp/2R with k=1,2, μ0=4π×107 NA 2 and R is the average distance between the flux-qubits and the individual NV-centers. gk depends on Ik and R [17], for example, with a typical persistent current value in the flux-qubit Ip=0.3 μA and the distance R=1.2 μm the coupling strength is gk/2π4.4 kHz [21]. Therefore, the total Hamiltonian can be written as H1=Ωq2i=12σxi+Jσz1σz2+i=1MDSz,i2+i=1MSx,i(g1σz1+g2σz2). Considering the relations in Eq. (8), we rewrite the Hamiltonian H1 as H2=i=14Ei|ii|+i=1MDSz,i2+i=1MSx,i[G¯(cos θ|4+sin θ|1)3|+G˜(sin θ|4cos θ|1)2|+h.c.], where G¯=g1+g2 and G˜=g1g2. Transforming H2 into a rotating frame via H3=eiRtH2eiRt+R with R=D|11|+(Dω¯)|22|ω¯|44|+Di=1M|0ii0| and ω¯=ω2J, we obtain

H3=Δ|44|+δ|33|+i=1M[eiDt|bii0|+h.c.][G¯(cos θeiω¯t|4+sin θeiDt||1)3|+G˜(sinθeiDt|4cosθeiω¯t|1)2|+h.c.],
where |b=(|++|)/2, Δ=ω42D and δ=ω31D. Because of ω31=ω42=ω in above discussion, we have δ=Δ. Due to the fact of 2D,D+ω¯,Dω¯Δ,e˙lta, under the rotating-wave approximation [17], Eq. (9) is reduced to the effective Hamiltonian as below,
H4=Δ|44|+δ|33|+sin θi=1M[|bii0|(G¯|13|+G˜|24|)+h.c.],
which is the Hamiltonian H1 in the main text by replacing g¯=sin θG¯ and g˜=sin θG˜.

We have to emphasize that the effect regarding E is ignored in above treatment, and our treatment below will follow Eq. (10). But if E is large enough, some modifications are necessary. For example, in the case of |δ|=|ω31DE| and |Δ|=|ω42DE|E, |b should be replaced by |+/2. In contrast, if |δ|=|ω31D+E| and |Δ|=|ω42D+E|E, |b should be written as |/2. Meanwhile, for these two cases, D should be replaced by D+E or DE, respectively. With these modifications, following treatments still apply.

3. Deriving the effective self-Kerr nonlinearity Hamiltonian

We investigate the Hamiltonian H2 (Eq. (3) in the main text) in the dressed states picture. In order to obtain the effective self-Kerr Hamiltonian Heff, we will mainly concentrate on the analysis of H2 excluding the pumping term Hp=εp(a+a) as HK=Δσ44+δσ33+Ω(σ32+σ23)+a(λ¯σ31+λ˜σ42+h.c.). Here, we define σij=|ij|. Actually, the pumping term Hp as a probe field is weak in our work, thus it is assumed to be virtually negligible.

For convenience, we use |bosonnumber,fluxqubitstate as the notation for the bare states. Then, the general dressed state of the Hamiltonian HK can be written as |Φ=n=0m=14cnm|n,m. Here, cnm is the coefficients of the bare states |n,m. On the basis of the level of excitation of the whole system, the Hamiltonian HK naturally separates into different manifolds, without the driving Hp. Obviously, the ground state of the system HK is |g=|0,1. The bare states |1,1, |0,2 and |0,3, formed the first manifold of the system. Thus, according to Hamiltonian HK, the dressed states in the first manifold can be written as |e0=(λ¯|0,2Ω|1,1)/X and |e±=±μ±|0,3+μ(Ω|0,2+λ¯|1,1)/X. Where, X=Ω2+λ¯2, μ±=(Y±δ)/2Y, with Y=δ2+4X2. The corresponding eigenenergies are ε0=0, ε±=(δ±Y)/2. For the nth manifold (n2), the four bare states of the system are |n+1,1, |n,2, |n,3, and |n1,4.

According to the dressed states |ej (j=0,±), we define the polariton creation operators in the first manifold as [58, 59] P0=(λ¯σ21Ωa)/X and P±=±μ±σ31+μ(Ωσ21+λ¯a)/X, such that |ej=Pj|g with j=0,±. These operators satisfy the commute formula, [Pi,Pj]=δij, [Pi,Pj]=[Pi,Pj]=0 with i,j=0,±, which means that these polaritons are bosons.

In addition, according to the polariton creation operators Pj, we obtain σ21=Ω(μP++μ+P)/X+λ¯P0/X and a=λ¯(μP++μ+P)/XΩP0/X.

 figure: Fig. 6

Fig. 6 Squeezing spectrum S(ω) under different conditions, where (a) the solid (dashed) curve denotes Ω=30 (36) and (b) Ω=50 (60) in unit of u0 with u0=0.1 MHz. Here we choose Δ=δ=5 and ϵp=2. Other parameters are the same as in Fig. 1(c) of main text.

Download Full Size | PDF

Besides, for the Hamiltonian HK, we assume that the detuning of levels |3 and |4 are sufficiently large (Δγ42 and δ(γ31+γ32)/2), so we can adiabatically eliminate levels |3 and |4. Then, the Hamiltonian HK can be effectively written as HmΔλ˜2/(γ422+Δ2)aaσ22+δλ¯2/[(γ312+γ322)+δ2]aaσ11, which was described just by the operators {a,σ11,σ12,σ22} (and their adjoints). With the expressions σ21 and a (and their adjoints), the Hamiltonian Hm can be expressed in terms of polariton operators Pj with j=0,±. On the other hand, when the condition |δ|, |Δ|, |λ¯|, |λ˜||ε±| is satisfied, the terms containing the operators P± (P±) will oscillate with much higher frequency than the terms containing only P0 (P0). Therefore, under the rotating wave approximation, the resulting effective Kerr Hamiltonian has the simple form as H˜eff=χ˜(P0)2P02 with χ˜=[λ¯2/(Ω2+λ¯2)]{Δλ˜2/(γ422+Δ2)δλ¯2/[(γ312+γ322)+δ2]}. When the weak coupling limit λ¯2/Ω21 is satisfied, the effective Hamiltonian becomes Heff=χa2a2 with χ=(λ¯2/Ω2){Δλ˜2/(γ422+Δ2)δλ¯2/[(γ312+γ322)+δ2]}.

Appendix B the quadrature squeezing

The quadrature squeezing can also be created by the self-Kerr nonlinearity [34, 35]. Define ε=2iχTr[aaρss] with ρss denoting the steady state solution of Eq. (4). If |epsilon|2κ2/3, the squeezing spectrum has a single peak at ω=0. Otherwise, it splits into two peaks at ω2=3|ε|2κ2 [35]. For the first condition, the squeezing spectrum is written by

S(ω)=2ξ2(13ξ2)2+16ξ2ξ[(ω2/κ2+13ξ2)(13ξ2)+16ξ2](13ξ2)2+16ξ2[(ω2/κ2+13ξ2)2+12ξ2]
with the dimensionless parameter ξ=|ε|/κ. Shown in Fig. 6(a) is the spectrum with its minimum at ω=0. An optimal squeezing spectrum obtained at ξ=1/3 is S(ω)=2ξ2/(ω4/κ4+12ξ2) with the maximal squeezing S(0)=1/6. If the second condition is satisfied, the maximum squeezing spectrum is S(ω)=2ξ2/[(ω2/κ2+13ξ2)2+12ξ2) [see Fig. 6(b)] and the two peaks appear at ω±=±κ3ξ21 with a same squeezing S(ω±)=1/6.

 figure: Fig. 7

Fig. 7 (a) Dynamical evolution of the fidelity for different sizes of the cat states. The curves from left to right correspond to the cat states with the size |α|2 changing form 2 to 6.5. (b) and (c) are zooming-in plots of Figs. 3(a) and 3(b) of the main text, respectively.

Download Full Size | PDF

Appendix C the ideal Kerr state

The Kerr nonlinearity can be observed in the evolution of a coherent state [36]. An ideal Kerr state arises from the evolution of a coherent state |α, which is the eigenstate of the annihilation operator a under the effective Hamiltonian Heff=χa2a2. After evolving for time t, the ideal Kerr state is written as |ψ(α,θ)=e|α|2/2n=0αnn!ein(n1)θ|n with θ=χt. Since n(n1) is an even number, the period for |ψ(α,θ) is T=π that means |ψ(α,θ+T)=|ψ(α,θ). Supposing θ=NMT with the integer M is indivisible with respect to the integer N, we rewrite the state |ψ(α,θ) by a superposition of coherent states as |ψ(α,θ=NMT)=k=02M1ck|eiφkα. Here, φk=kπ/M and ck=1+(1)kN(M1)2Mn=0M1exp {iπM[nk+Nn(n1)]} shows that only M different coherent states exist. For simplicity, we set N = 1 and particularly consider M = 2 that is for the Schrödinger-cat state |α,π2=12(eiπ/4|iα+eiπ/4|iα). The phase difference between two neighboring cat-states is i which could be applied to speculate the effective anharmonicity parameter χ in the evolution based on the master equation.

Appendix D the detail for the dynamical evolution of fidelity in the preparation of different-size cat state

Zooming-in plots around the inflection point |α|2=5.05of Fig. 3 are shown in Fig. 7(a) where the change of the optimal fidelity can be explicitly seen in Fig. 7(a). The first oscillating period in Fig. 3 is amplified in Fig. 7(b) and a sharp change in fidelity can be found around |α|2=5.05. On the other hand, an amplification for the first stair in Fig. 3 is shown in Fig. 7(c) where a sudden jump at the inflection point |α|2=5.05 occurs.

 figure: Fig. 8

Fig. 8 (a) and (c) Optimal fidelity of the cat state as a function of the cat’s size. Inset: the dots denote the oscillating period of the fidelity for |α|25 and the line means a linear fitting of these dots. (b) and (d) The time τop for obtaining the optimal fidelity of the cat state. Inset: the dots represent the corresponding spacing Δτop between the nearest neighbor two ladders and the curve denotes an exponential fitting. The black, red and green data and curves in (a) and (b) correspond to γ=0.2,0.4 and 1 with κ=1, and in (c) and (d) they are corresponding to κ=0.5,1 and 2 with γ=0.4. The other parameters are same in Fig. 3 of main text.

Download Full Size | PDF

Figure 8 presents some details about the fidelity variation under different flux-qubit dissipation rates γ and NVE decay rates κ. It clear proves that Eq. (5) in the main text hold at different dissipation rate and decoherence rate. Particularly, in Figs. 8(a) and 8(b), the parameters of the oscillation length Δ|α|2 and optimal time space Δτop in Eq. (5) are corresponding to C=1.91(1.92,1.92), a1=3.4(3.4,3.4)×103, a2=0.26,0.26,0.26 and a3=1.96(1.96,1.96)×103 for γ=0.2,0.4,1, and they are C=2.03(1.92,1.91), a1=3.4(3.39,3.35)×103, a2=0.26,0.26,0.27 and a3=1.98(1.96,1.99)×103 for κ=0.5,1,2 in Figs. 8(c) and 8(d). Therefore, the property denoted by Eq. (5) is determined by the internal dynamic of flux-qubits and independent on the flux-qubits dissipation and NVE decay.

Funding

National Key Research and Development Program of China (Grant No. 2017YFA0304503); National Natural Science Foundation of China (Grants No. 11835011, No. 11804375, No. 11734018, No. 11674360, No. 11574353, and No. 11404377); Strategic Priority Research Program of the Chinese Academy of Sciences (Grant No. XDB21010100); China Postdoctoral Science Foundation (Grant No. 2018M642956).

References

1. E. Schrödinger, “Die gegenwärtige situation in der quantenmechanik,” Naturwissenschaften 23, 807–812 (1935). [CrossRef]  

2. C. Monroe, D. M. Meekhof, B. E. King, and D. J. Wineland, “A Schrödinger cat superposition state of an atom,” Science 272, 1131–1136 (1996). [CrossRef]   [PubMed]  

3. A. Ourjoumtsev, H. Jeong, R. Tualle-Brouri, and P. Grangier, “Generation of optical Schrödinger cat from photon number states,” Nature 448, 784–786 (2007). [CrossRef]   [PubMed]  

4. M. Arndt, O. Nairz, J. Vos-Andreae, C. Keller, G. van der Zouw, and A. Zeilinger, “Wave-particle duality of C60 molecules,” Nature 401, 680–682 (1999). [CrossRef]  

5. D. Leibfried, E. Knill, S. Seidelin, J. Britton, R. B. Blakestad, J. Chiaverini, D. B. Hume, W. M. Itano, J. D. Jost, C. Langer, R. Ozeri, R. Reichle, and D. J. Wineland, “Creation of a six-atom Schrödinger cat state,” Nature 438, 639–642 (2005). [CrossRef]   [PubMed]  

6. W.-B. Gao, C.-Y. Lu, X.-C. Yao, P. Xu, O. Gühne, A. Goebel, Y.-A. Chen, C.-Z. Peng, Z.-B. Chen, and J.-W. Pan, “Experimental demonstration of a hyper-entangled ten-qubit Schrödinger cat state,” Nat. Phys. 6, 331–335 (2010). [CrossRef]  

7. K. G. Johnson, J. D. Wong-Campos, B. Neyenhuis, J. Mizrahi, and C. Monroe, “Ültrafast creation of large Schrödinger cat states of an atom,” Nat. Commun. 8, 697 (2017). [CrossRef]  

8. M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, “Cavity optomechanics,” Rev. Mod. Phys. 86, 1391–1452 (2014). [CrossRef]  

9. M. Blencowe, “Quantum electromechanical systems,” Phys. Rep. 395, 159–222 (2004). [CrossRef]  

10. O. Romero-Isart, A. C. Pflanzer, F. Blaser, R. Kaltenbaek, N. Kiesel, M. Aspelmeyer, and J. I. Cirac, “Large quantum superpositions and interference of massive nanometer-sized objects,” Phys. Rev. Lett. 107, 020405 (2011). [CrossRef]   [PubMed]  

11. J.-Q. Liao and L. Tian, “Macroscopic quantum superposition in cavity optomechanics,” Phys. Rev. Lett. 116, 163602 (2016). [CrossRef]   [PubMed]  

12. Y. Kubo, C. Grezes, A. Dewes, T. Umeda, J. Isoya, H. Sumiya, N. Morishita, H. Abe, S. Onoda, T. Ohshima, V. Jacques, A. Drau, J.-F. Roch, I. Diniz, A. Auffeves, D. Vion, D. Esteve, and P. Bertet, “Hybrid quantum circuit with a superconducting qubit coupled to a spin ensemble,” Phys. Rev. Lett. 107, 220501 (2011). [CrossRef]   [PubMed]  

13. T. Astner, S. Nevlacsil, N. Peterschofsky, A. Angerer, S. Rotter, S. Putz, J. Schmiedmayer, and J. Majer, “Coherent coupling of remote spin ensembles via a cavity bus,” Phys. Rev. Lett. 118, 140502 (2017). [CrossRef]   [PubMed]  

14. W. L. Yang, Z. Yin, Z. X. Chen, S. Kou, M. Feng, and C. H. Oh, “Quantum simulation of an artificial Abelian gauge field using nitrogen-vacancy-center ensembles coupled to superconducting resonators,” Phys. Rev. A 86, 012307 (2012). [CrossRef]  

15. R. Harris, F. Brito, A. J. Berkley, J. Johansson, M. W. Johnson, T. Lanting, P. Bunyk, E. Ladizinsky, B. Bumble, A. Fung, A. Kaul, A. Kleinsasser, and S. Han, “Synchronization of multiple coupled rf-SQUID flux qubits,” New J. Phys. 11, 123022 (2009). [CrossRef]  

16. R. Harris, T. Lanting, A. J. Berkley, J. Johansson, M. W. Johnson, P. Bunyk, E. Ladizinsky, N. Ladizinsky, T. Oh, and S. Han, “Compound Josephson-junction coupler for flux qubits with minimal crosstalk,” Phys. Rev. B 80, 052506 (2009). [CrossRef]  

17. D. Marcos, M. Wubs, J. M. Taylor, R. Aguado, M. D. Lukin, and A. S. Sørensen, “Coupling nitrogen-vacancy centers in diamond to superconducting flux qubits,” Phys. Rev. Lett. 105, 210501 (2010). [CrossRef]  

18. Z.-L. Xiang, X.-Y. Lü, T.-F. Li, J. Q. You, and F. Nori, “Hybrid quantum circuit consisting of a superconducting flux qubit coupled to a spin ensemble and a transmission-line resonator,” Phys. Rev. B 87, 144516 (2013). [CrossRef]  

19. W. L. Song, W. L. Yang, Z. Q. Yin, C. Y. Chen, and M. Feng, “Controllable quantum dynamics of inhomogeneous nitrogen-vacancy center ensembles coupled to superconducting resonators,” Sci. Rep. 6, 33271 (2016). [CrossRef]   [PubMed]  

20. Y. Maleki and A. M. Zheltikov, “Generating maximally-path-entangled number states in two spin ensembles coupled to a superconducting flux qubit,” Phys. Rev. A 97, 012312 (2018). [CrossRef]  

21. X. Zhu, S. Saito, A. Kemp, K. Kakuyanagi, S. Karimoto, H. Nakano, W. J. Munro, Y. Tokura, M. S. Everitt, K. Nemoto, M. Kasu, N. Mizuochi, and K. Semba, “Coherent coupling of a superconducting flux qubit to an electron spin ensemble in diamond,” Nature 478, 221–224 (2011). [CrossRef]   [PubMed]  

22. X. Zhu, Y. Matsuzaki, R. Amsüss, K. Kakuyanagi, T. S. Oka, N. Mizuochi, K. Nemoto, K. Semba, W. J. Munro, and S. Saito, “Observation of dark states in a superconductor diamond quantum hybrid system,” Nat. Commun. 5, 3424 (2014). [CrossRef]   [PubMed]  

23. S. Saito, X. Zhu, R. Amsüss, Y. Matsuzaki, K. Kakuyanagi, T. S. Oka, N. Mizuochi, K. Nemoto, W. J. Munro, and K. Semba, “Towards realizing a quantum memory for a superconducting qubit: storage and retrieval of quantum states,” Phys. Rev. Lett. 111, 107008 (2013). [CrossRef]  

24. A. Imamoglu, H. Schmidt, G. Woods, and M. Deutsch, “Strongly interacting photons in a nonlinear cavity,” Phys. Rev. Lett. 79, 1467–1470 (1997). [CrossRef]  

25. M. J. Werner and A. Imamoglu, “Photon-photon interactions in cavity electromagnetically induced transparency,” Phys. Rev. A 61, 011801 (1999). [CrossRef]  

26. M. Fleischhauer, A. Imamoglu, and J. P. Marangos, “Electromagnetically induced transparency: optics in coherent media,” Rev. Mod. Phys. 77, 633–673 (2005). [CrossRef]  

27. S Rebić, S M Tan, A S Parkins, and D F Walls, “Large Kerr nonlinearity with a single atom,” J. Opt. B 1, 490–495 (1999). [CrossRef]  

28. Y. Maleki and A. M. Zheltikov, “Witnessing quantum entanglement in ensembles of nitrogen-vacancy centers coupled to a superconducting resonator,” Opt. Express 14, 17849–17858 (2018). [CrossRef]  

29. S. Rebic, J. Twamley, and G. J. Milburn, “Giant Kerr nonlinearities in circuit quantum electrodynamics,” Phys. Rev. Lett. 103, 150503 (2009). [CrossRef]   [PubMed]  

30. F. Jelezko, T. Gaebel, I. Popa, A. Gruber, and J. Wrachtrup, “Observation of coherent oscillations in a single electron spin,” Phys. Rev. Lett. 92, 076401 (2004). [CrossRef]   [PubMed]  

31. G. Chen, H. Zhang, Y. Yang, R. Wang, L. Xiao, and S. Jia, “Qubit-induced high-order nonlinear interaction of the polar molecules in a stripline cavity,” Phys. Rev. A 82, 013601 (2010). [CrossRef]  

32. H. Schmidt and A. Imamoglu, “Giant Kerr nonlinearities obtained by electromagnetically induced transparency,” Opt. Lett. 21, 1936–1938 (1996). [CrossRef]   [PubMed]  

33. P. D. Drummond and D. F. Walls, “Quantum theory of optical bistability. I. nonlinear polarisability model,” J. Phys. A 13, 725–741 (1980). [CrossRef]  

34. M. D. Reid and D. F. Walls, “Squeezing via optical bistability,” Phys. Rev. A 32, 396–401 (1985). [CrossRef]  

35. M. J. Collett and D. F. Walls, “Squeezing spectra for nonlinear optical systems,” Phys. Rev. A 32, 2887–2892 (1985). [CrossRef]  

36. G. Kirchmair, B. Vlastakis, Z. Leghtas, S. E. Nigg, H. Paik, E. Ginossar, M. Mirrahimi, L. Frunzio, S. M. Girvin, and R. J. Schoelkopf, “Observation of quantum state collapse and revival due to the single-photon Kerr effect,” Nature 495, 205–209 (2013). [CrossRef]   [PubMed]  

37. D. F. Walls and G. J. Milburn, Quantum optics(Springer, 1994). [CrossRef]  

38. M. Kira, S. W. Koch, R. P. Smith, A. E. Hunter, and S. T. Cundiff, “Quantum spectroscopy with Schrödinger-cat states,” Nat. Phys. 7, 799–804 (2011). [CrossRef]  

39. A. Facon, E.-K. Dietsche, D. Grosso, S. Haroche, J.-M. Raimond, M. Brune, and S. Gleyzes, “A sensitive electrometer based on a Rydberg atom in a Schrödinger-cat state,” Nature 535, 262–265 (2016). [CrossRef]   [PubMed]  

40. J. Joo, W. J. Munro, and T. P. Spiller, “Quantum metrology with entangled coherent states,” Phys. Rev. Lett. 107, 083601 (2011). [CrossRef]   [PubMed]  

41. J. S. N. Nielsen, Y. Eto, C. W. Lee, H. Jeong, and M. Sasaki, “Quantum tele-amplification with a continuous-variable superposition state,” Nat. Photon. 7, 439–443 (2013). [CrossRef]  

42. J. Huang, X. Qin, H. Zhong, Y. Ke, and C. Lee, “Quantum metrology with spin cat states under dissipation,” Sci. Rep. 5, 17894 (2015). [CrossRef]   [PubMed]  

43. T. Unden, P. Balasubramanian, D. Louzon, Y. Vinkler, M. B. Plenio, M. Markham, D. Twitchen, A. Stacey, I. Lovchinsky, A. O. Sushkov, M. D. Lukin, A. Retzker, B. Naydenov, L. P. McGuinness, and F. Jelezko, “Quantum metrology enhanced by repetitive quantum error correction,” Phys. Rev. Lett. 116, 230502 (2016). [CrossRef]   [PubMed]  

44. A. Z. Chaudhry, “Ütilizing nitrogen-vacancy centers to measure oscillating magnetic fields,” Phys. Rev. A 90, 042104 (2014). [CrossRef]  

45. T. Wolf, P. Neumann, K. Nakamura, H. Sumiya, T. Ohshima, J. Isoya, and J. Wrachtrup, “Subpicotesla diamond magnetometry,” Phys. Rev. X 5, 041001 (2015).

46. W. L. Ma, S. S. Li, G. Y. Cao, and R. B. Liu, “Atomic-scale positioning of single spins via multiple nitrogen-vacancy centers,” Phys. Rev. Applied 5, 044016 (2016). [CrossRef]  

47. H. Liu, M. B. Plenio, and J. Cai, “Scheme for detection of single-molecule radical pair reaction using spin in diamond,” Phys. Rev. Lett. 118, 200402 (2017). [CrossRef]   [PubMed]  

48. M. Bruderer, P. F. Acebal, R. Aurich, and M. B. Plenio, “Sensing of single nuclear spins in random thermal motion with proximate nitrogen-vacancy centers,” Phys. Rev. B 93, 115116 (2016). [CrossRef]  

49. D. Leibfried, D. M. Meekhof, B. E. King, C. Monroe, W. M. Itano, and D. J. Wineland, “Experimental determination of the motional quantum state of a trapped atom,” Phys. Rev. Lett. 77, 4281 (1996). [CrossRef]  

50. M. Hofheinz, H. Wang, M. Ansmann, R. C. Bialczak, E. Lucero, M. Neeley, A. D. O’Connell, D. Sank, J. Wenner, J. M. Martinis, and A. N. Cleland, “Synthesizing arbitrary quantum states in a superconducting resonator,” Nature 459, 546–549 (2009). [CrossRef]   [PubMed]  

51. B. R. Johnson, M. D. Reed, A. A. Houck, D. I. Schuster, L. S. Bishop, E. Ginossar, J. M. Gambetta, L. DiCarlo, L. Frunzio, S. M. Girvin, and R. J. Schoelkopf, “Quantum non-demolition detection of single microwave photons in a circuit,” Nat. Phys. 6, 663 (2010). [CrossRef]  

52. E. Solano, “Selective interactions in trapped ions: state reconstruction and quantum logic,” Phys. Rev. A 71, 013813 (2005). [CrossRef]  

53. S. Haroche and J. M. Raimond, Exploring the quantum: atoms, cavities, and photons(Oxford university, 2006). [CrossRef]  

54. M. Scala, M. S. Kim, G. W. Morley, P. F. Barker, and S. Bos, “Matter-wave interferometry of a levitated thermal nano-oscillator induced and probed by a spin,” Phys. Rev. Lett. 111, 180403 (2013). [CrossRef]   [PubMed]  

55. L. P. Neukirch, E. von Haartman, J. M. Rosenholm, and A. N. Vamivakas, “Multi-dimensional single-spin nano-optomechanics with a levitated nanodiamond,” Nat. Photo. 9, 653–657 (2015). [CrossRef]  

56. J. E. Mooij, T. P. Orlando, L. Levitov, L. Tian, C. H. van der Wal, and S. Lloyd, “Josephson persistent-current qubit,” Science 285, 1036–1039 (1999). [CrossRef]   [PubMed]  

57. R. Harris, M. W. Johnson, S. Han, A. J. Berkley, J. Johansson, P. Bunyk, E. Ladizinsky, S. Govorkov, M. C. Thom, S. Uchaikin, B. Bumble, A. Fung, A. Kaul, A. Kleinsasser, M. H. S. Amin, and D. V. Averin, “Probing noise in flux qubits via macroscopic resonant tunneling,” Phys. Rev. Lett. 101, 117003 (2008). [CrossRef]   [PubMed]  

58. M. J. Hartmann, F. G. S. L. Brandão, and M. B. Plenio, “Strongly interacting polaritons in coupled arrays of cavities,” Nat. Phys. 2, 849–855 (2006). [CrossRef]  

59. K. Cai, Z.-W. Pan, R.-X. Wang, D. Ruan, Z.-Q. Yin, and G.-L. Long, “Single phonon source based on a giant polariton nonlinear effect,” Opt. Lett. 43, 1163–1166 (2018). [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 (8)

Fig. 1
Fig. 1 (a) Schematic diagram of the hybrid quantum system composed of two coupled CJJ rf-SQUID qubits and an NVE. Inset: the level structure of an NV-center [30] under a weak classical field with an amplitude ξp coupling the states | ± to | 0 . (b) The N-type level structure for the coupled flux-qubits with ω = J + J 2 + Ω q 2 and ω ¯ = J 2 + Ω q 2 J . A control field with Rabi frequency Ω and frequency ω c = 2 J is in resonance with the states | 2 and | 3 . (c) Three-dimensional plot of log 10 [ g 2 ( 0 ) ] as functions of Ω and ϵp. The inset denotes the nonlinearity factor χ varying with the amplitude ϵp of the driving field along the gradient direction (red curve). Here the coupling strengths λ ¯ / κ = 400 and λ ˜ / κ = 200 . The decay paths are set as γ 41 = γ 32 = 0.1 γ and γ 31 = γ 21 = γ 42 = γ 43 = γ . Where, γ i j represent the decay rate from the states | i to | j . Other parameters are Δ = δ = κ and γ / κ = 0.4 .
Fig. 2
Fig. 2 (a) Time evolution of the fidelity F . (b) Wigner function W ( β ) at five different time points corresponding to t1, t2, t3, t4 and t5 in (a). (c) Probability distribution of the rotated quadrature operator PX at time points t3 (blue solid curve) and t4 (black dashed curve). Here we choose κ = 1 , γ = 0.4 , Δ = δ = 1 , α = 2 e i π / 4 , ϵ p = 2 , λ ˜ = 300 , Ω = 50 and decay paths are chosen same as in Fig. 1(c).
Fig. 3
Fig. 3 (a) Optimal fidelity of the cat state as a function of the cat’s size. Inset: the dots denote the oscillating period of the fidelity for | α | 2 5 and the line means a constant obtained by fitting these dots. (b) Optimal time c t τ op in variation with the cat’s size. Inset: the dots represents Δ τ op between the nearest-neighbor ladders and the curve denotes an exponential fitting. (c) Dynamical evolution of the fidelity for three sizes of the cat states around the inflection point | α | 2 = 5.05 , which is labeled as a red dot in (a). (d) Probability distribution of the rotated quadrature operator PX at time points r o t e c t κ τ op = 0.02 ( | α | 2 = 4.8 , 5 ) and 0.025 ( | α | 2 = 5.1 , 5.8 ). Here the parameters used are the same as in Fig. 2.
Fig. 4
Fig. 4 (a) and (b) Fidelities of the cat state prepared under different flux-qubit dissipation rates and NVE decay rates, where we set u 0 / 2 π = 0.1 MHz, and for convenience of presentation the parameters below are written in units of u0. We have κ = 1 in (a) and γ = 1 in (b). The solid, dashed and dotted curves in both (a) and (b) correspond to λ ˜ = 300 , 200 and 100 , respectively. Inset: time evolution of the fidelity with t λ ˜ = 300 . (c) Probability distribution of the rotated quadrature operator PX at the time point t4 of Fig. 2. The dotted, dashed and solid curves correspond to the excitation numbers n c = 1, 2 and 5. In (c), κ = 1 and γ = 0.4 . Inset: Wigner function W ( β ) with respect to n c = 1 and 5. Other parameters in (a-c) are p h a = 2 e i π / 4 , ϵ p = 2 , λ ¯ = 400 and Ω = 50 , respectively.
Fig. 5
Fig. 5 (a) Fidelities of the cat state prepared under different dephasing rates κd. (b) A detail dynamical evolution of the fidelities for three different dephasing rates at a coupling λ ˜ = 300 . (c) The Wigner function W ( β ) atthe optimal cat state position of in (b) for the case κ d = 0.5 . Here κ = 1 and γ = 0.4 and the other parameters are same as in Fig. 4.
Fig. 6
Fig. 6 Squeezing spectrum S ( ω ) under different conditions, where (a) the solid (dashed) curve denotes Ω = 30 (36) and (b) Ω = 50 (60) in unit of u0 with u 0 = 0.1 MHz. Here we choose Δ = δ = 5 and ϵ p = 2 . Other parameters are the same as in Fig. 1(c) of main text.
Fig. 7
Fig. 7 (a) Dynamical evolution of the fidelity for different sizes of the cat states. The curves from left to right correspond to the cat states with the size | α | 2 changing form 2 to 6.5. (b) and (c) are zooming-in plots of Figs. 3(a) and 3(b) of the main text, respectively.
Fig. 8
Fig. 8 (a) and (c) Optimal fidelity of the cat state as a function of the cat’s size. Inset: the dots denote the oscillating period of the fidelity for | α | 2 5 and the line means a linear fitting of these dots. (b) and (d) The time τ op for obtaining the optimal fidelity of the cat state. Inset: the dots represent the corresponding spacing Δ τ op between the nearest neighbor two ladders and the curve denotes an exponential fitting. The black, red and green data and curves in (a) and (b) correspond to γ = 0.2 , 0.4 and 1 with κ = 1 , and in (c) and (d) they are corresponding to κ = 0.5 , 1 and 2 with γ = 0.4 . The other parameters are same in Fig. 3 of main text.

Equations (11)

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

H = Ω q 2 i = 1 2 σ x i + J σ z 1 σ z 2 + i = 1 M D S z , i 2 + i = 1 M S x , i ( g 1 σ z 1 + g 2 σ z 2 ) ,
H 1 = Δ | 4 4 | + δ | 3 3 | + i = 1 M [ | b i i 0 | ( g ¯ | 1 3 | + g ˜ | 2 4 | ) + h . c . ] ,
H 2 = H 0 + a ( λ ¯ | 3 1 | + λ ˜ | 4 2 | ) + h . c . ,
ρ ˙ = i [ H , ρ ] + κ 2 D [ a , ρ ] + j < k γ j , k 2 D [ σ j k , ρ ] ,
Δ | α | 2 C , Δ τ op = a 1 e a 2 n + a 3 ,
( | 4 | 3 | 2 | 1 ) = 1 2 ( cos  θ sin  θ sin  θ cos  θ 1 0 0 1 0 1 1 0 sin  θ cos  θ cos  θ sin  θ ) ( | e e | e g | g e | g g ) .
( | e e | e g | g e | g g ) = 1 2 ( cos  θ 1 0 sin  θ sin  θ 0 1 cos  θ sin  θ 0 1 cos  θ cos  θ 1 0 sin  θ ) ( | 4 | 3 | 2 | 1 ) .
σ z 1 = cos  θ | 4 3 | + sin  θ | 4 2 | + sin  θ | 3 1 | cos  θ | 2 1 | + h . c . , σ z 2 = cos  θ | 4 3 | sin  θ | 4 2 | + sin  θ | 3 1 | + cos  θ | 2 1 | + h . c
H 3 = Δ | 4 4 | + δ | 3 3 | + i = 1 M [ e i D t | b i i 0 | + h . c . ] [ G ¯ ( cos   θ e i ω ¯ t | 4 + sin   θ e i D t | | 1 ) 3 | + G ˜ ( sin θ e i D t | 4 cos θ e i ω ¯ t | 1 ) 2 | + h . c . ] ,
H 4 = Δ | 4 4 | + δ | 3 3 | + sin  θ i = 1 M [ | b i i 0 | ( G ¯ | 1 3 | + G ˜ | 2 4 | ) + h . c . ] ,
S ( ω ) = 2 ξ 2 ( 1 3 ξ 2 ) 2 + 16 ξ 2 ξ [ ( ω 2 / κ 2 + 1 3 ξ 2 ) ( 1 3 ξ 2 ) + 16 ξ 2 ] ( 1 3 ξ 2 ) 2 + 16 ξ 2 [ ( ω 2 / κ 2 + 1 3 ξ 2 ) 2 + 12 ξ 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.