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

Optical continuous-variable quadratic phase gate via Faraday interaction

Open Access Open Access

Abstract

The continuous-variable (CV) quadratic phase gate is one of the most fundamental CV quantum gates for universal CV quantum computation, while its experimental realization still remains a challenge. Here we propose a novel and experimentally feasible scheme to realize optical CV quadratic phase gate via Faraday interaction in an atomic ensemble. The gate is performed by simply sending an optical beam three times through an atomic medium prepared in coherent spin state. The fidelity of the gate can ideally run up to one. We show that the scheme also works well as a device to generate optical polarization squeezing. Considering the noise effects due to atomic decoherence and light losses, we find that the observed fidelities of gate operation and the attainable degree of polarization squeezing are still quite high.

© 2014 Optical Society of America

1. Introduction

Quantum computation (QC), which strives to utilize the principles of quantum mechanics to realize efficient computation, is one of the most fascinating and fruitful area of research. In recent years, with the seminal discoveries of many quantum algorithms [1, 2], QC has been proved to be able to work faster than any known classic computation. Though originally based upon discrete variables, QC over CV has also recently been developed by Lloyd and Braunstein [3, 4].

CV QC generalizes the concept of QC to continuous degrees of freedom by the use of position x^=(a^+a)/2 and momentum p^=i(a^a^)/2 quadratures. A universal CV quantum computer is defined to be a device that can by local operations perform any desired unitary transformation over these quantum variables [3]. To realize such transformation, the canonical universal set of gates {k(s), (s), ĈZ} is used, where k(s) = exp(isx̂k) for k = 1, 2, 3 and for all sR [5], with 1(s) the displacement gate, 2(s) the quadratic phase (QP) gate which maps the quadratures into

D^2(s)x^D^2(s)=x^,D^2(s)p^D^2(s)=p^+sx^,
3(s) the cubic gate, (s) = exp[is(2 + 2)] the rotation gate, and CZ = exp(ix̂) the two-mode controlled-Z gate. By repeated use of these gates, together with the approximation [3]
eiA^seiB^seiA^seiB^se[A^,B^]s2+O(s3)
in the limit s → 0 for arbitrary Hamiltonians  and consisting of a general polynomials of and , one is able to simulate any Hamiltonian transformations to any precision, and thus realize the universal CV QC.

The universal set without the cubic operation 3(s) is still universal for arbitrary Gaussian transformations [5]. For quadrature amplitudes of light, operations 1(s) and (s) are the most readily available Gaussian transformations, which can be easily realized by using only coherent light resources and linear optics [5, 6]. Operation CZ can be finitely decomposed into beam splitters and single-mode online squeezers [5, 7, 8]. Operation 2(s) (together with operation (s) providing online squeezing [9]), however, is the most challenging Gaussian transformation, as its realization usually involves nonlinear optical processes. To date, many methods have been proposed to perform optical QP operation or, equivalently, online squeezer [1015]. A typical one is based on the nonlinear processes in the optical crystals or optical fibers. For these optical systems, since the nonlinear effect inside a nonlinear media is normally very weak, strong pumper power [10,11] or cavities [12] are usually required to obtain sufficient nonlinearities. Such enhancement, however, suffers several drawbacks. First, strongly optical pumping induced nonlinear process are usually not pure and introduces excess noise to the input signals; second, the use of cavity complicates the experimental realization and makes it difficult to inject the signals into the system. To overcome these drawbacks, an alternative approach using only linear optical elements, offline squeezing, and homodyne detection has also recently been proposed [1315]. This approach, however, suffers the drawback that its gate fidelities in some extent are limited by the degree of squeezing of the input offline squeezed state.

In this paper, we propose a new and experimentally feasible method to realize optical QP gate in a free-space macroscopic atomic ensemble. We show that ideal optical QP transformation can be performed by simply sending a coherent light three times through a coherent-prepared atomic ensemble (see Fig. 1). In contrast to previous measurement-based schemes [1315], our method requires neither non-classical offline squeezing nor homodyne detection, which greatly saves the resources needed and simplifies the experimental implementation. Moreover, we also checked the ability of the scheme to produce optical polarization squeezing, finding that substantial squeezing is obtainable within the presently experimentally available parameters.

 figure: Fig. 1

Fig. 1 Scheme setup for realization of optical QP gate in an atomic ensemble. The light beam first enters an atomic ensemble at an angle α with respect to z to experience Ĥ1 interaction. The outgoing beam then passes through a delay line. After the whole of the beam run through the atomic sample, it enters the ensemble again to experience Ĥ2 interaction and subsequently Ĥ3 interaction. When the beam exits, an optical QP gate is performed.

Download Full Size | PDF

The remainder of this paper is organized as follows. In section 2, we first give details of QP operation based on Faraday interaction with an atomic ensemble, and then calculate the optical squeezing generated. In section 3, we will consider the noise effects including atomic decay and light reflection. After that, the experimental feasibility of the scheme is also discussed. Finally, section 4 contains brief conclusions.

2. The protocol

2.1. Quadratic phase gate

The scheme is based on the Faraday interaction between light and atoms. Suppose that we have an atomic gas consisting of a large number of spin- 12 atoms interacting with a light pulse traveling along the z direction. The atomic sample is initially prepared in a coherent spin state (CSS), i.e., a fully polarized state along the x axis. For such a state, the transverse collective spin components Ĵy and Ĵz maintain their quantum nature, while the macroscopic x component can be treated as classical constant number, that is ĴxJx. The light pulse interacting with atoms is composed of a strong component polarized along the x axis and a weak component (quantum field) along y. For such a polarized state, we analogously have the relevant quantum variables Ŝy, Ŝz and the classical variable Sx, where Ŝ represents the Stokes operators characterizing the polarization properties of a light pulse. In the far off-resonant coupling regime, one obtains the Faraday interaction between light and atoms: ĤeffĴzŜz [1618], which leads to a rotation of the optical linear polarization due to the population difference between the atomic magnetic sublevels, and a rotation of spin around the z axis because of the different ac-Stark shift on the sublevels caused by the field intensity difference between σ+ and σ components [19]. This kind of interactions have been extensively studied to realize, e.g., polarization squeezing [19], spin squeezing [2022], and quantum memory [2326].

We here consider how to use this interaction to realize the optical QP gate. Supposing that a light pulse described above propagates through an atomic ensemble in the YZ plane at a certain angle α (0 ≤ α < 2π) with respect to direction z, (see Fig. 1). In this case, the light beam not only see the spin z component but also the spin y component. Consequently, the Hamiltonian is changed into [27]: Ĥ1 = κx̂L(A sinα + A cosα), where κ is the coupling strength with the dimension 1/time, and we have defined the new quantum variables (x^A,p^A)=(J^y,J^z)/Jx for atoms, obeying the the canonical commutator relation [A, A] = i, and (x^L,p^L)=(S^z(t),S^y(t))/Sx(t) for light, satisfying [L(t), L(t′)] = (tt′). Corresponding to this Hamiltonian, one may derive the relations between input and output state of light [19, 28]:

x^Lout1(t)=x^Lin(t),p^Lout1(t)=p^Lin(t)κ[x^A(t)sinα+p^A(t)cosα].
Here, the field variables ϑ^Lin(ϑ^Lout1)[ϑ^{x^,p^}] represents light before (after) interacting with atoms. In the Heisenberg picture, the spin variables evolve according to
ddtx^A(t)=κx^Lin(t)cosα,ddtp^A(t)=κx^Lin(t)sinα.
This set of equations can be easily solved and generates
x^A(t)=x^A(0)+κcosα0tx^Lin(τ)dτ,p^A(t)=p^A(0)κsinα0tx^Lin(τ)dτ.
We now define the dimensionless collective mode for light ϑ^Lin,out1=1T0Tϑ^Lin,out1(t)dt and the new variables for atoms ϑ^Ain=ϑ^A(0), ϑ^Aout1=ϑ^A(T), where T denotes the duration of light pulse. With these definition, inserting (5) into (3), one obtains
yout1=S(κ˜,α)yin,
where we have collected quadratures into a vector y = (L, L, A, A)T and defined the interaction matrix
S(κ˜,α)=(100001κ˜sinακ˜cosακ˜cosα010κ˜sinα001),
where the dimensionless coupling strength κ˜=κT. The off-diagonal matrix elements in (7) characterize the fact that information about the L and L quadratures is now stored in the atomic mode, and, simultaneously, information of A and A is also printed onto the light mode. While for QP operation, according to Eq. (1), the L quadrature contains no other information but the conjugate L. To pick up the information of L recorded in the atoms during the first pass, we suggest reflecting the pulse back into sample again to experience a second interaction Ĥ2 = κx̂L(A sinβ + A cosβ). Before the second interaction, the light is reflected by mirrors and propagates in a delay line (see Fig. 1), which ensures the two interactions do not overlap in the time line [29]. The second interaction transfers the atomic and light states into yout2 = S (κ̃, β)yout1 = S (α, β)yin, where the effective interaction matrix is given by
S(α,β)=(1000κ˜2sin2ϕ12κ˜sinϕ+cosϕ2κ˜cosϕ+cosϕ2κ˜cosϕ+cosϕ0102κ˜sinϕ+cosϕ001)
with ϕ± = (α ± β)/2. The off-diagonal matrix element κ̃2 sin2ϕ in the first column indicates that the position L quadrature is now mixed into L as desired. Particularly, for a special value ϕ = π/4 the output light state is exactly the same as the spin squeezed state created in [30]. As a result, the obtained light state is a polarization squeezed state. Such state, however, is not a pure state, as it still contains information about the atomic state, as can be seen from the off-diagonal elements in (8). This kind of information is unwanted, because they, on the one hand, reduce the amount of squeezing [31] and, on the other hand, prevent the realization of perfect QP gate. To eliminate them the pulse is sent back into the sample to experience a third interaction, Ĥ3 = κx̂L(A sinθ + A cosθ), which drives the system into yout3 = S(κ̃, θ)yout2 = S (α, β, θ,)yin, where the matrix is
S(α,β,θ)=(1000κ˜2fαβ,αθ,βθs1κ˜fα,β,θsκ˜fα,β,θcκ˜fα,β,θc010κ˜fα,β,θs001).
Here we have defined the new variables fα,β,θs=φ=α,β,θsinφ and fα,β,θc=φ=α,β,θcosφ. Obviously, to realize a perfect optical QP gate it is required that
fα,β,θs=fα,β,θc=0,
fαβ,αθ,βθs0.
Equation (10) leads to the limitations: |αβ| = 2π/3, |α + β − 2θ| = 2π or |αβ| = 4π/3, |α + β − 2θ| = 0. According to these limitations one may choose a special set of angle parameters, e.g., α = π/6, β = 5π/6, θ = 3π/2, with which Eq. (11) can be easily calculated to give f2π3,4π3,2π3s=32κ˜20. Finally, we obtain the input-output relations for light
x^Lout3=x^Lin,p^Lout3=p^Lin32κ˜2x^Lin,
which is the main result of this paper. Apparently, this optical state is exactly the same as the state given in Eq. (1) with s=3κ˜2/2. Therefore, for arbitrary non-zero κ̃ a perfect optical QP gate is successfully performed. Moreover, it is worth noting here that, the sign of fαβ,αθ,βθs can be flipped freely via suitable choice of α, β and θ, e.g., for α = 5π/6, β = π/6, θ = 3π/2 yielding f2π3,2π3,4π3s=32κ˜2. Such characteristic is especially valuable, since two conjugate QP operations 2(s) and D^2(s) together with the beam splitters enable the performance of perfect CV controlled-Z gate (see Appendix for details).

2.2. Optical polarization squeezing

As mentioned above, the performance of QP gate is always accompanied with the generation of optical squeezing. To find out the amount of squeezing generated, we perform a rotation of the final output state: ϑ^Lout3R^(s)ϑ^Lout3R^(s), leading to

(x^Lout3p^Lout3)=(cos(s)sin(s)sin(s)cos(s))(x^Lout3p^Lout3).
Assuming that the input light field is in the vacuum state, the variance of x^Lout3 can then be readily calculated to give
Δ2x^Lout3=12[132κ˜2sin(2s)+34κ˜4cos2(s)],
which can be minimized at smin=12arctan(4/3κ˜2), resulting in the optimized squeezing parameter ζ(smin)=1+3κ˜4[11+16/(3κ˜4)]/8limκ˜4/(3κ˜4). In contrast to the double-pass scheme [19] whose squeezing parameter is ζκ̃→∞ ≈ 3/(2κ̃2), our current scheme strongly enhances the degree of optical polarization squeezing.

3. Experimental considerations

3.1. Noise effects

So far, we have neglected the noise effects. While in reality, every photon on its way through the atomic media has a mall probability for being absorbed [32, 33], and thus cause photon losses. Such losses, however, can be modeled by a beam splitter type admixture of vacuum components [18,34], which transforms the output light quadratures as ϑ^Loutϑ^Lout=1εϑ^Lout+εϑ^Lv, where ε denotes the absorbtion coefficient and ϑ^Lv represents the vacuum noise quadrature with zero mean and 1/2 variance. Besides, absorption losses due to the propagation in delay lines can also be taken into account by replacing ε with ξ = ε + r, where r represents the the overall absorptivity of the delay line. On the other hand for atoms, due to collisional relaxation and weak excitation by light, they also undergo dissipation. When take the noise effects into account, as discussed in [17, 18, 32], the atomic state becomes ϑ^Aoutϑ^Aout=1ηϑ^Aout+ηϑ^Av where η is the atomic decay parameter and ϑ^Av is the atomic vacuum noise. With these modifications, the atomic and light states after the first pass are then transformed into:

yout1=D¯(ξ,η)S(κ˜,α)yin+D(ξ,η)yv1,
where yvi denotes the noise vector of ith pass, and D(ξ,η)=diag(ξ,ξ,η,η), D¯(ξ,η)=1D(ξ,η)2. Analogously, the final state created after three passages can then be calculated by iterating the map defined by the above equation. For the rest two interactions, however, one should take into account the fact that, the classical variables Jx and Sx also decay from pass to pass as: 〈Jx〉 → (1 − η)〈Jx,Sx〉 → (1 − ξ)〈Sx〉, which reduces the coupling strength by a factor τ=(1η)(1ξ) [32]. As a result, after three passages the atomic and light states become
yout3=D¯(ξ,η)S(τ2κ˜,θ)[D¯(ξ,η)S(τκ˜,β)yout1+D(ξ,η)yv2]+D(ξ,η)yv3.
By inserting the expression for the interaction matrix given in Eq. (7) into this equation and using the set of special angle parameters chosen above, we finally arrive at the modified input-output relations for light
x^Lout3=(1ξ)32x^Lin+ξf^x,
p^Lout3=(1ξ)32(p^Lin32κ˜2x^Lin)+η(1ξ)32κ˜[(η32)x^Ain+32p^Ain]+ηf^Av+ξf^Lv,
where we have defined κ˜=(1η)(1ξ+ξη)κ˜, f^ϑ=(1ξ)ϑ^Lv1+1ξϑ^Lv2+ϑ^Lv3, and the collective vacuum operator for atoms f^Av=(1η)(1ξ)3/2κ˜[(1/2η)x^Av1+3p^Av1/2+(1η)x^Av2) and for light f^Lv=3(1η)2(1ξ)2κ˜2x^Lv1/2+f^p. The second line of Eq. (18) shows that, now the atomic quadratures can not be completely canceled due to the decay of the atomic system. As a result, compared to the ideal case of Eq. (12), the light state obtained here is somewhat distorted. In order to quantify the distortion of this state, we use the formula of fidelity (ρout3,ρout3)=[Tr(ρout3ρout3ρout3)1/2]2 which is a measure of how well the distorted output state ρ̂out3′ of Eqs. (17) and (18) compares to the ideal output state ρout3 of Eq. (12). In principle, the input light state can be arbitrary. However, here for simplicity and concreteness (but without loss of generality), we assume the input light state is a Gaussian state, e.g. a coherent state or a squeezed state. In this case, the output state conserves the Gaussian character of the input state due to the fact that the interaction Hamiltonians involved here are bilinear in the canonical operators [4]. As a result, the fidelity becomes [35]
(ρout3,ρout3)=1det(Γout3+Γout3)+δδ×exp[12𝒜T(Γout3+Γout3)1𝒜],
where Γ stands for the covariance matrix having its elements Γij = 〈yiyj + yjyi〉 − 2〈yi yj〉, and where δ = (detΓout3 − 1)(detΓout3′ − 1), 𝒜 = mout3mout3′ with m the vector of mean values. If the ideal output state is a pure Gaussian state, we will have det(Γout3) = 1 and in turn δ = 0 [35]. Moreover, as an example we assume that the input Gaussian states center around zero, such that 𝒜 = 0. Equation (19) then reduces to a much simpler form (ρout3,ρout3)=2/det(Γout3+Γout3).

From Eqs. (12), (17), and (18), Γout3 and Γout3′ can be respectively calculated, and finally one is able to show the gate fidelity in its dependence on κ̃, η, and ξ. In the case of coherent state input, the fidelity versus κ̃ is depicted in Fig. 2(a) for different loss parameters η, ξ. As can be seen from the figure that, the achievable fidelities are strongly dependent on the coupling strength κ̃. This result is reasonable since the larger the coupling strength is, the more “nonclassical” the output state become. Such (highly-nonclassical) output states are more fragile to noise effects, resulting in lower achievable fidelities. In the weak interaction regime [corresponding to small κ̃, where the CV QC works as shown in Eq. (2)], however, high fidelities are obtainable and the scheme is robust against the noises. Putting κ̃ = 2, the insert of Fig. 2(a) shows the fidelity in its dependence on light reflections ξ for different decay rate. One can see from the insert that, in contrast to the atomic decay the light losses affect the fidelities more significantly. For squeezed state input, in Fig. 2(b) we plot the fidelity vs κ̃ for different squeezing of the input state and for η = ξ = 0.10 (which corresponds to the experimental conditions of [36]). It clearly shows that the gate fidelity is very sensitive to the squeezing degree of the input state. The fidelity is greatly reduced for different input squeezing even when κ̃ = 0 (corresponding to no atom in the sample), which reflects the fact that the squeezed states are very sensitive to photon losses. The insert of Fig. 2(b) indicates that, in the weak interaction regime the scheme works quite well within certain range of squeezing.

 figure: Fig. 2

Fig. 2 (a) Gate operation fidelity for coherent states vs coupling κ̃ in the presence of atomic decay and light reflection. Inset: the gate fidelity in its dependent on ξ for different decay rate and κ̃ = 2. (b) Gate operation fidelity for squeezed states vs coupling κ̃ including 10% atomic decay and 10% light losses. Inset: the gate fidelity in its dependent on input squeezing for decay parameters η = ξ = 0.10 and coupling κ̃ = 0.10. Choosing the light losses ξ = 0.05, the fidelity of a fixed QP operation with s=3/2 varies with the atomic decay (c) for ρ = 30 and with the optical depth (d). The inset of (d) shows how the optimal η varies with r.

Download Full Size | PDF

So far, the coupling strength κ̃ and the atomic decay η have been treated independently, while in fact they are linked to each other through κ̃2 = ρη [17], where ρ is the optical depth of the atomic sample. For a realistic system such as room-temperature vapor, we typically have ρ ≈ 30 [37]. If one performs a fixed 2(s) operation in such system with, e.g., s=3/2 (corresponding to κ̃ = 1), one is able to show in Fig. 2(c) the fidelity in its dependence on η (which can be tuned by choosing either the detuning from atomic resonance or the total photon number of the pulse [17]) for different squeezing of the input state including ξ = 0.05 absorption losses. In each case, there exist an optimal fidelity, = 0.98, = 0.93, and = 0.81, for η = 0.028, η = 0.032, and η = 0.034, respectively. The η-optimized fidelities vs ρ is also depicted in Fig. 2(d), showing that, the larger the optical depth is, the higher the fidelities one will obtain. Besides, it is worth noting in figure 2(d) that, for the region of 0 to 20, it is quite economical to tune the optical depth to improve the gate fidelity. Once the optical depth of the systems surpasses this region, it might be more efficient to use other ways (e.g., light-reflection reduction) to improve the performance of gate.

Unlike the gate performance, optical polarization squeezing usually works in the strong interaction regime. Along the lines presented in Sec. 2 above, one may directly calculate the amount of squeezing in the presence of noise effect based on Eqs. (17) and (18). Finally, in Fig. 3(a) we are able to plot the squeezing parameter ζ, optimized with respect to rotation s and direction θ, in its dependence on the coupling κ̃, which indicates the amount of achievable squeezing are bounded by the added noise. For the parameters κ̃2 = 3 and η = 0.10 of room temperature atomic vapors [37], the achievable squeezing are 4.8 dB, 6.3 dB, 8.0 dB for the light losses ξ = 0.1, 0.05, 0.01, respectively. The s, θ -optimized squeezing vs ξ is also depicted in Fig. 3(b). One can see from the figure that, the atomic decay plays a minor role in the scheme, while the light losses are the major factor, which limits the creation of polarization squeezing quantitatively and qualitatively.

 figure: Fig. 3

Fig. 3 (a) The optimized polarization squeezing ζ vs coupling κ̃ in the presence of atomic decay and light reflection. The inset shows how the optimal parameters smin and θ vary with κ̃. (b) Optimized squeezing vs light losses ξ for different atomic decay.

Download Full Size | PDF

3.2. Experimental feasibility

To successfully and efficiently perform the CV QP gate, it is required that (i) T < TDL and r ≪ 1, where TDL represents a time in which the pulse pass through the delay lines, and (ii) η, ε ≪ 1. Condition (ii) can be fulfilled in both room-temperature [36, 37] and cold [3840] atomic systems, since the optical depth in these systems usually satisfy ρ ≫ 1 (e.g., ρ ≥ 50 for cold atoms [40]). As a result, the decay parameter satisfies η = κ̃2/ρ ≪ 1 for small κ̃. If one chooses NpNa (Np the total photon number and Na the total atom number), we will have εη ≪ 1 [18]. Condition (i), however, is the main challenge of current scheme. For previous experiments of quantum interface between light and atoms [28], the typical value of the width of a pulse is about T ≈ 1 ms. Correspondingly, hundreds of kilometer-long delay lines between the passages are normally required, which poses a challenge to the experimental realization. Fortunately, the situation has recently been changed dramatically. Faraday interaction using a very short pulse—the width of which is about T = 100 ns and Laser-cooled atoms has been successfully demonstrated [38, 39]. Besides, optical delay line using cylindrical mirrors with TDL ≈ 1μs which conserves the polarization of light pulse has also been reported in [41], showing that reflective losses r less than 2% can be achieved with the aid of high reflective mirrors. Therefore, condition (i) is also feasible.

4. CONCLUSIONS

In conclusion, we have presented a novel and experimentally feasible scheme for realizing optical CV QP gate in an atomic ensemble. The process is based on off-resonant interaction between light and spin-polarized atomic ensembles. By sending a pulse three times through an coherent prepared atomic ensemble, we found that a QP operation is successfully performed, and its fidelities can ideally run up to one. We also found that the sign of the parameter s of 2(s) can be freely flipped. This is a quite good characteristic, since it enables the performance of perfect two-mode CV controlled-Z gate. Such a gate is a key resource for CV QC, especially for CV one-way QC [29, 42, 43], while it is by now still experimentally challenging. We also studied the capacity of polarization squeezing of current scheme, showing that, in contrast to double-pass scheme, the amount of squeezing is strongly enhanced. Finally, the influences of the noise effects including the atomic decay and photon reflections losses are considered. For coherent input, the scheme works quite well in weak iteration regime, while for squeezed input, the scheme is sensitive to photon losses. When it comes to polarization squeezing, the situation is analogous to the double-pass scheme given in [19], that is, the attainable degree of squeezing of the scheme are mainly limited by light losses. This confirms our motivation to search for new techniques to reduce the losses in the loop, e.g. if one can use a little more shorter pulse than [38, 39], e.g. T = 10 ns, the delay line are no longer necessary, and light losses will be diminished significantly. We expect that our scheme can be beneficial in the context of CV QC.

5. Appendix

In this Appendix we present the details on how to realize the two-mode CV Controlled-Z operation using QP gates. Suppose that we want to perform a CZ gate between two light beams i and j. An ideal controlled-Z gate CZ = exp(ix̂ij) transfers the quadratures into [29, 42, 43]: i,ji,j, i,ji,j + j,i. Such transformation can be achieved by three steps. In the first step, the two beams are mixed on a beam splitter, which transfers the vector of quadratures v = (i, i, j, j)T into vout1 = ij(θ1)vin, where the beam splitter operator can be written as [4]:

B^ij(θ1)=(cosθ10sinθ100cosθ10sinθ1sinθ10cosθ100sinθ10cosθ1).
In the second step, the two output beams are then injected into two atomic ensembles described above to experience D^2i(s) and D^2j(s) transformations, respectively. After this operation the light states become vout2=D^2i(s)D^2j(s)vout1=D^ij(s)vout1, where
D^ij(s)=(1000s100001000s1).
In the last step, the two beams are mixed on a second beam splitter to generate the final output state vout3 = ij(θ2)vout2 = ij(θ2)ij(s)ij(θ1)vin = ij(s)vin. Taking the special parameters for beam splitters θ1 = −θ2 = π/4 and setting s = 1, we finally have
ij(s)=(1000011000101001).
Consequently, a perfect CV controlled-Z gate between two light beams is successfully performed.

Acknowledgments

This work was supported by the Natural Science Foundation of Zhejiang province (Grants No. LY12A05001), the Natural Science Foundation of China (Grants No. 11074190), and the department of education of Zhejiang province ( Y201120838).

References and links

1. P. W. Shor, “Polynomial-time algorithms for prime factorization and discrete logarithms on a quantum computer,” SIAM J. Comput. 26, 1484–1509 (1997). [CrossRef]  

2. S. Lloyd and S. L. Braunstein, “Quantum computation over continuous variables,” Phys. Rev. Lett. 82, 1784–1787 (1999). [CrossRef]  

3. L. K. Grover, “Quantum mechanics helps in searching for a needle in a haystack,” Phys. Rev. Lett. 79, 325–328 (1997). [CrossRef]  

4. S. L. Braunstein and P. van Loock, “Quantum information with continuous variables,” Rev. Mod. Phys. 77, 513–577 (2005). [CrossRef]  

5. M. Gu, C. Weedbrook, N. C. Menicucci, T. C. Ralph, and P. van Loock, “Quantum computing with continuous-variable clusters,” Phys. Rev. A 79, 062318 (2009). [CrossRef]  

6. M. G. Paris, “Displacement operator by beam splitter,” Physics Letters A 217, 78–80 (1996). [CrossRef]  

7. S. L. Braunstein, “Squeezing as an irreducible resource,” Phys. Rev. A 71, 055801 (2005). [CrossRef]  

8. M. Reck, A. Zeilinger, H. J. Bernstein, and P. Bertani, “Experimental realization of any discrete unitary operator,” Phys. Rev. Lett. 73, 58–61 (1994). [CrossRef]   [PubMed]  

9. S. Sefi, “Squeezing and photon counting with the cubic phase state,” arXiv:quant-ph/1310.2429 (2013).

10. J. A. Levenson, I. Abram, T. Rivera, and P. Grangier, “Reduction of quantum noise in optical parametric amplification,” J. Opt. Soc. Am. B 10, 2233–2238 (1993). [CrossRef]  

11. J. A. Levenson, I. Abram, T. Rivera, P. Fayolle, J. C. Garreau, and P. Grangier, “Quantum optical cloning amplifier,” Phys. Rev. Lett. 70, 267–270 (1993). [CrossRef]   [PubMed]  

12. J. P. Poizat and P. Grangier, “Experimental realization of a quantum optical tap,” Phys. Rev. Lett. 70, 271–274 (1993). [CrossRef]   [PubMed]  

13. R. Filip, P. Marek, and U. L. Andersen, “Measurement-induced continuous-variable quantum interactions,” Phys. Rev. A 71, 042308 (2005). [CrossRef]  

14. J.-i. Yoshikawa, T. Hayashi, T. Akiyama, N. Takei, A. Huck, U. L. Andersen, and A. Furusawa, “Demonstration of deterministic and high fidelity squeezing of quantum information,” Phys. Rev. A 76, 060301 (2007). [CrossRef]  

15. Y. Miwa, J.-i. Yoshikawa, P. van Loock, and A. Furusawa, “Demonstration of a universal one-way quantum quadratic phase gate,” Phys. Rev. A 80, 050303 (2009). [CrossRef]  

16. A. Kuzmich, N. P. Bigelow, and L. Mandel, “Atomic quantum non-demolition measurements and squeezing,” Europhys. Lett. 42, 481–486 (1998). [CrossRef]  

17. K. Hammerer, A. S. Sørensen, and E. S. Polzik, “Quantum interface between light and atomic ensembles,” Rev. Mod. Phys. 82, 1041–1093 (2010). [CrossRef]  

18. L.-M. Duan, J. I. Cirac, P. Zoller, and E. S. Polzik, “Quantum communication between atomic ensembles using coherent light,” Phys. Rev. Lett. 85, 5643–5646 (2000). [CrossRef]  

19. J. F. Sherson and K. Mølmer, “Polarization squeezing by optical faraday rotation,” Phys. Rev. Lett. 97, 143602 (2006). [CrossRef]   [PubMed]  

20. Y. Takahashi, K. Honda, N. Tanaka, K. Toyoda, K. Ishikawa, and T. Yabuzaki, “Quantum nondemolition measurement of spin via the paramagnetic faraday rotation,” Phys. Rev. A 60, 4974–4979 (1999). [CrossRef]  

21. A. Kuzmich, L. Mandel, J. Janis, Y. E. Young, R. Ejnisman, and N. P. Bigelow, “Quantum nondemolition measurements of collective atomic spin,” Phys. Rev. A 60, 2346–2350 (1999). [CrossRef]  

22. R. Inoue, S.-I.-R. Tanaka, R. Namiki, T. Sagawa, and Y. Takahashi, “Unconditional quantum-noise suppression via measurement-based quantum feedback,” Phys. Rev. Lett. 110, 163602 (2013). [CrossRef]   [PubMed]  

23. B. Julsgaard, J. Sherson, J. I. Cirac, J. Fiurasek, and E. S. Polzik, “Experimental demonstration of quantum memory for light,” Nature 432, 482–486 (2004). [CrossRef]   [PubMed]  

24. J. Fiurášek, J. Sherson, T. Opatrný, and E. S. Polzik, “Single-passage readout of atomic quantum memory,” Phys. Rev. A 73, 022331 (2006). [CrossRef]  

25. C. A. Muschik, K. Hammerer, E. S. Polzik, and J. I. Cirac, “Efficient quantum memory and entanglement between light and an atomic ensemble using magnetic fields,” Phys. Rev. A 73, 062329 (2006). [CrossRef]  

26. T. Takano, M. Fuyama, R. Namiki, and Y. Takahashi, “Continuous-variable quantum swapping gate between light and atoms,” Phys. Rev. A 78, 010307 (2008). [CrossRef]  

27. J. Stasińska, C. Rodó, S. Paganelli, G. Birkl, and A. Sanpera, “Manipulating mesoscopic multipartite entanglement with atom-light interfaces,” Phys. Rev. A 80, 062304 (2009). [CrossRef]  

28. B. Julsgaard, “Quantum memory and teleportation using macroscopic gas samples,” Ph.D. thesis, Aarhus University (2003).

29. M.-F. Wang, N.-Q. Jiang, Q.-L. Jin, and Y.-Z. Zheng, “Continuous-variable controlled-z gate using an atomic ensemble,” Phys. Rev. A 83, 062339 (2011). [CrossRef]  

30. M. Takeuchi, S. Ichihara, T. Takano, M. Kumakura, T. Yabuzaki, and Y. Takahashi, “Spin squeezing via one-axis twisting with coherent light,” Phys. Rev. Lett. 94, 023003 (2005). [CrossRef]   [PubMed]  

31. C. M. Trail, P. S. Jessen, and I. H. Deutsch, “Strongly enhanced spin squeezing via quantum control,” Phys. Rev. Lett. 105, 193602 (2010). [CrossRef]  

32. K. Hammerer, K. Mølmer, E. S. Polzik, and J. I. Cirac, “Light-matter quantum interface,” Phys. Rev. A 70, 044304 (2004). [CrossRef]  

33. L. B. Madsen and K. Mølmer, “Spin squeezing and precision probing with light and samples of atoms in the gaussian description,” Phys. Rev. A 70, 052324 (2004). [CrossRef]  

34. K. Hammerer, E. S. Polzik, and J. I. Cirac, “Teleportation and spin squeezing utilizing multimode entanglement of light with atoms,” Phys. Rev. A 72, 052313 (2005). [CrossRef]  

35. H. Scutaru, “Fidelity for displaced squeezed thermal states and the oscillator semigroup,” J. Phys. A 31, 3659–3663 (1998). [CrossRef]  

36. J. F. Sherson, H. Krauter, R. K. Olsson, B. Julsgaard, K. Hammerer, I. Cirac, and E. S. Polzik, “Quantum teleportation between light and matter,” Nature 443, 557–560 (2006). [CrossRef]   [PubMed]  

37. J. F. Sherson, “Quantum memory and teleportation using macroscopic gas samples,” Ph.D. thesis, Aarhus University (2006).

38. M. Takeuchi, T. Takano, S. Ichihara, A. Yamaguchi, M. Kumakura, T. Yabuzaki, and Y. Takahashi, “Fast polarimetry system for the application to spin quantum non-demolition measurement,” Appl. Phys. B 83, 33–36 (2006). [CrossRef]  

39. T. Takano, M. Fuyama, R. Namiki, and Y. Takahashi, “Spin squeezing of a cold atomic ensemble with the nuclear spin of one-half,” Phys. Rev. Lett. 102, 033601 (2009). [CrossRef]   [PubMed]  

40. M. Kubasik, M. Koschorreck, M. Napolitano, S. R. de Echaniz, H. Crepaz, J. Eschner, E. S. Polzik, and M. W. Mitchell, “Polarization-based light-atom quantum interface with an all-optical trap,” Phys. Rev. A 79, 043815 (2009). [CrossRef]  

41. E. Jeffrey, M. Brenner, and P. Kwiat, “Spin squeezing of a cold atomic ensemble with the nuclear spin of one-half,” Proc. SPIE 5161, 269–279 (2004). [CrossRef]  

42. N. C. Menicucci, P. van Loock, M. Gu, C. Weedbrook, T. C. Ralph, and M. A. Nielsen, “Universal quantum computation with continuous-variable cluster states,” Phys. Rev. Lett. 97, 110501 (2006). [CrossRef]   [PubMed]  

43. N. C. Menicucci, X. Ma, and T. C. Ralph, “Arbitrarily large continuous-variable cluster states from a single quantum nondemolition gate,” Phys. Rev. Lett. 104, 250503 (2010). [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 (3)

Fig. 1
Fig. 1 Scheme setup for realization of optical QP gate in an atomic ensemble. The light beam first enters an atomic ensemble at an angle α with respect to z to experience Ĥ1 interaction. The outgoing beam then passes through a delay line. After the whole of the beam run through the atomic sample, it enters the ensemble again to experience Ĥ2 interaction and subsequently Ĥ3 interaction. When the beam exits, an optical QP gate is performed.
Fig. 2
Fig. 2 (a) Gate operation fidelity for coherent states vs coupling κ̃ in the presence of atomic decay and light reflection. Inset: the gate fidelity in its dependent on ξ for different decay rate and κ̃ = 2. (b) Gate operation fidelity for squeezed states vs coupling κ̃ including 10% atomic decay and 10% light losses. Inset: the gate fidelity in its dependent on input squeezing for decay parameters η = ξ = 0.10 and coupling κ̃ = 0.10. Choosing the light losses ξ = 0.05, the fidelity of a fixed QP operation with s = 3 / 2 varies with the atomic decay (c) for ρ = 30 and with the optical depth (d). The inset of (d) shows how the optimal η varies with r.
Fig. 3
Fig. 3 (a) The optimized polarization squeezing ζ vs coupling κ̃ in the presence of atomic decay and light reflection. The inset shows how the optimal parameters smin and θ vary with κ̃. (b) Optimized squeezing vs light losses ξ for different atomic decay.

Equations (22)

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

D ^ 2 ( s ) x ^ D ^ 2 ( s ) = x ^ , D ^ 2 ( s ) p ^ D ^ 2 ( s ) = p ^ + s x ^ ,
e i A ^ s e i B ^ s e i A ^ s e i B ^ s e [ A ^ , B ^ ] s 2 + O ( s 3 )
x ^ L out 1 ( t ) = x ^ L in ( t ) , p ^ L out 1 ( t ) = p ^ L in ( t ) κ [ x ^ A ( t ) sin α + p ^ A ( t ) cos α ] .
d d t x ^ A ( t ) = κ x ^ L in ( t ) cos α , d d t p ^ A ( t ) = κ x ^ L in ( t ) sin α .
x ^ A ( t ) = x ^ A ( 0 ) + κ cos α 0 t x ^ L in ( τ ) d τ , p ^ A ( t ) = p ^ A ( 0 ) κ sin α 0 t x ^ L in ( τ ) d τ .
y out 1 = S ( κ ˜ , α ) y in ,
S ( κ ˜ , α ) = ( 1 0 0 0 0 1 κ ˜ sin α κ ˜ cos α κ ˜ cos α 0 1 0 κ ˜ sin α 0 0 1 ) ,
S ( α , β ) = ( 1 0 0 0 κ ˜ 2 sin 2 ϕ 1 2 κ ˜ sin ϕ + cos ϕ 2 κ ˜ cos ϕ + cos ϕ 2 κ ˜ cos ϕ + cos ϕ 0 1 0 2 κ ˜ sin ϕ + cos ϕ 0 0 1 )
S ( α , β , θ ) = ( 1 0 0 0 κ ˜ 2 f α β , α θ , β θ s 1 κ ˜ f α , β , θ s κ ˜ f α , β , θ c κ ˜ f α , β , θ c 0 1 0 κ ˜ f α , β , θ s 0 0 1 ) .
f α , β , θ s = f α , β , θ c = 0 ,
f α β , α θ , β θ s 0 .
x ^ L out 3 = x ^ L in , p ^ L out 3 = p ^ L in 3 2 κ ˜ 2 x ^ L in ,
( x ^ L out 3 p ^ L out 3 ) = ( cos ( s ) sin ( s ) sin ( s ) cos ( s ) ) ( x ^ L out 3 p ^ L out 3 ) .
Δ 2 x ^ L out 3 = 1 2 [ 1 3 2 κ ˜ 2 sin ( 2 s ) + 3 4 κ ˜ 4 cos 2 ( s ) ] ,
y out 1 = D ¯ ( ξ , η ) S ( κ ˜ , α ) y in + D ( ξ , η ) y v 1 ,
y out 3 = D ¯ ( ξ , η ) S ( τ 2 κ ˜ , θ ) [ D ¯ ( ξ , η ) S ( τ κ ˜ , β ) y out 1 + D ( ξ , η ) y v 2 ] + D ( ξ , η ) y v 3 .
x ^ L out 3 = ( 1 ξ ) 3 2 x ^ L in + ξ f ^ x ,
p ^ L out 3 = ( 1 ξ ) 3 2 ( p ^ L in 3 2 κ ˜ 2 x ^ L in ) + η ( 1 ξ ) 3 2 κ ˜ [ ( η 3 2 ) x ^ A in + 3 2 p ^ A in ] + η f ^ A v + ξ f ^ L v ,
( ρ out 3 , ρ out 3 ) = 1 det ( Γ out 3 + Γ out 3 ) + δ δ × exp [ 1 2 𝒜 T ( Γ out 3 + Γ out 3 ) 1 𝒜 ] ,
B ^ i j ( θ 1 ) = ( cos θ 1 0 sin θ 1 0 0 cos θ 1 0 sin θ 1 sin θ 1 0 cos θ 1 0 0 sin θ 1 0 cos θ 1 ) .
D ^ i j ( s ) = ( 1 0 0 0 s 1 0 0 0 0 1 0 0 0 s 1 ) .
i j ( s ) = ( 1 0 0 0 0 1 1 0 0 0 1 0 1 0 0 1 ) .
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.