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

Strong photon antibunching with weak second-order nonlinearity under dissipation and coherent driving

Open Access Open Access

Abstract

By solving the master equation in the steady-state limit and calculating the zero-delay-time second-order correlation function, a strong photon antibunching is found in two coupled cavities with weak nonlinearity χ(2). An optimal antibunching condition is derived that is complementary to the result in a recent publication [Phys. Rev. A 89, 031803(R) (2014)]. Numerical simulations confirm the optimal condition. These results are extended from the case with a single drive to the case with two drives. We find that the parameter added to the system due to the second drive can lift the restriction on the detuning for the antibunching to occur. This may open a new door towards the application of the antibunching into single photon sources.

© 2016 Optical Society of America

1. Introduction

There is a crucial requirement for the generation and manipulation of the single photon in information and communication technology. A single-photon source typically relies on a system which can produce sub-Poissonian light when it is driven by a classical light field. The basic physical principle for the generation of single photon is the photon blockade (PB) effect, which requires strong optical nonlinearity in most proposals published in the literature. This topic has become a research focus in modern quantum optics for the important applications.

The PB, also known as conventional phonon blockade, is a quantum optical effect preventing the injection of more than one photon into a nonlinear cavity mode [1, 2]. The PB origins from the strong photon-photon interaction which requires the nonlinear strength much greater than the decay rate of the system. It is generally considered that the actualization of PB demands a strong third-order (Kerr) nonlinearity for a photonic mode, which was first observed in an optical cavity coupled to a single trapped atom [3]. Subsequently, a sequence of experimental groups observed the strong antibunching behaviors in different systems [4–6]. Afterward it was proposed that a single-photon blockade can be achieved in second-order (χ(2)) nonlinear system [7, 8]. The potential applications of PB cover the realization of single-photon transistor [9], interferometers [10] and quantum optical diode with semiconductor microcavities [8].

Recently, a new mechanism for the generation of sub-Poissonian light was put forward by Liew and Savona, where a strong photon antibunching can be obtaind with nonlinearity smaller than the decay rate of the cavity mode [11]. This mechanism is termed unconventional photon blockade (UPB), which only requires a much smaller optical nonlinearity than its conventional counterpart [12, 13]. Different from the conventional mechanism, the blockade is enforced by quantum interference among multiple excitation pathways. The typical system for UPB consists of two coupled resonators, a very weak third-order nonlinearity characterizes the two resonators [14–18]. Up to present, many other systems are proposed to realize the UPB, such as coupled optomechanical systems [19], bimodal optical cavity with a quantum dot [20, 21], and second-order nonlinearity systems [22–25].

In [23], the UPB can be realized in microcavities with second-order nonlinearity, where the optimal condition that leads to strong antibunching is obtained. In this paper, we explore the strong antibunching under the condition complementary to that in [23]. We first derive a Hamiltonian to describe two coupled cavities with second-order nonlinearity, which has the same form as that in [23]. Then we investigate the optimal condition for strong antibunching. Although the model is same, the method and conclusion in ours are different from that in [23]. The strong antibunching conditions in [23] are obtained by transforming the second-order nonlinearity term in the three-mode Hamiltonian into a two-mode Kerr-type nonlinear Hamiltonian under the condition Δc ≫ 2Δb, where Δi represents the detuning between the driving laser and the corresponding cavity-mode i. Whereas in our scheme, we shall investigate the photon antibunching condition directly by using the second-order nonlinear Hamiltonian for the three-mode system. To get the strong antibunching condition with our method, the detuning Δi = 0 is needed. So we analyze the case of Δi = 0 with a single drive that is different from that in [23]. In addition, when two drives are applied to the system, we find that the strong photon antibunching can also be obtained when Δi = 0. Different from the case of single drive, the UPB with two drives has the following features: (i) The optimal conditions for strong antibunching exist and can be found when Δi ≠ 0. (ii) The UPB can be obtained even if both the χ(2) nonlinearity and the coupling strength are small. This study is quite different from our previous work of the UPB based on a two-mode (by contrast, three modes are considered in this paper) system with second-order nonlinearity [22], where the coupling J between the modes was not taken into account, and the antibunching cannot happen with a single drive since the quantum interference pathways cannot be formed. Besides, the effect of cavity decay on the UPB is different for the two-mode and three-mode system.

2. Model

Let us start with the nonlinear response of dielectric material to an electric field described by

Di(r,t)=ε0εijEj(r,t)+ε0[χijk(2)(r)Ej(r,t)Ek(r,t)]+
where the indices run over x, y and z. Assuming that the dielectric is inhomogeneous but isotropic, i.e., εij(r) → ε(r) [7], and restricting ourself to consider only the second-order nonlinearity χ(2)(r), i.e., χijkm(n)(r)=0, n ≥ 3, with the electromagnetic field for the three-mode system,
E^(r,t)=j=a,b,ch¯ωj20[j^ϕj(r)ε(r)eiωjt+H.C.],
where ĵ (j = a, b, c) is the destruction operator of mode j, and (r) = −∇ × Ê(r)/ω0, we can derive a Hamiltonian for the system.

Now we go to the details. To derive the Hamiltonian with second-order nonlinearity, we set ωc = 2ωa = 2ωb. The three dimensional cavity field for the three modes ϕi(r) can be normalized by ∫ |ϕ⃗i(r)|2d3r = 1 (i = a, b, c). The classical expression of the energy density is given by Hem=12[E^(r,t)D^(r,t)+H^(r,t)B^(r,t)]d3r, where Ĥ = B̂/μ0. Then the Hamiltonian for this system can be calculated as

H^em=h¯ωaa^a+h¯ωbb^b^+h¯ωcc^c^+h¯J(a^b^+b^a^)+h¯g[(b^)2c^+c^b^2]+h¯v[(a^)2c^+c^a^2)]+h¯u[a^b^c^+c^b^a^],
where
h¯J=h¯2ωaωbεij(r)ε(r)ϕa(r)ϕb(r)d3r,h¯g=h¯ωb22ωc20χ(2)(r)[ε(r)]3/2ϕb(r)ϕb(r)ϕc(r)d3r,h¯v=h¯ωa22ωc20χ(2)(r)[ε(r)]3/2ϕa(r)ϕa(r)ϕc(r)d3r,h¯u=2ε0[h¯2ε0]3/2ωaωbωcχ(2)(r)[ε(r)]3/2ϕa(r)ϕb(r)ϕc(r)d3r.
We consider the situation that mode a is in the first cavity, mode b and the second-harmonic mode c are in the second cavity, as shown in Fig. 1. Since the modes in different cavities are coupled, we only keep the leading order in Hamiltonian (3) and drop the nonlinear interaction coefficients v and u by assuming that the overlapped of these wave functions is very small. With these considerations, the Hamiltonian Eq. (3) reduces to
H^0=h¯ωaa^a^+h¯ωbb^b^+h¯ωcc^c^+h¯J(a^b^+b^a^)+h¯g[(b^)2c^+c^b^2].
For the scheme to work, external driving for a mode is essential. The driving frequency and driving strength are denoted by ωL and F, respectively. The Hamiltonian for the system with such a drive reads:
H^=H^0+h¯F(a^eiωLt+a^eiωLt).
For convenience, we study the system in a rotating frame defined by Û(t) =exp[Lt(ââ + + 2ĉĉ)], which leads to a rotated Hamiltonian by Ĥd =ÛĤÛih̄ÛdU/dt,
H^d=h¯Δaa^a^+h¯Δbb^b^+h¯Δcc^c^+h¯J(a^b^+b^a^)+h¯g[(b^)2c^+c^b^2]+h¯F(a^+a^),
where Δi = ωiωL (i = a, b) and Δc = ωc − 2ωL denote the detunings of the a, b and c modes from the driving laser, respectively. The derivation of the Hamiltonian Eq. (3) strictly relies on the relation ωc = 2ωa = 2ωb = 2ω, or equivalently, Δc = 2Δa = 2Δb. Hereafter, we only analyze the resonant case (Δc = 2Δa = 2Δb = 2Δ = 0) because our method cannot obtain a condition for antibunching with non-zero Δ (for details, see the APPENDIX).

 figure: Fig. 1

Fig. 1 The system of two coupled cavities. Modes a and b in the cavities are tunnel-coupled, mode b and mode c are coupled via the second-order nonlinear χ(2).

Download Full Size | PDF

Taken the cavity decay into account, the evolution of the system can be described by a master equation in the Lindblad form

ρ^t=ih¯[H^d,ρ^]+κa2(2a^ρ^a^+a^a^ρ^ρa^a^)+κb2(2b^ρ^b^b^b^ρ^ρb^b^)+κc2(2cρ^c^c^c^ρ^ρc^c^),
with κa, κb and κc being the damping rates of a mode, b mode and c mode, respectively. For the case of photonic crystal cavities, the higher order mode has always a lower quality factor Q (and thus a higher κ). The most optimistic assumptions in literature take Qa = Qb = 2Qc [7], which resulting in κc = 4κa = 4κb = 4κ. We will use these parameters in this paper. To discuss the UPB, we need to calculate the second-order correlation function defined by
g(2)(τ)=a^(0)a^(τ)a^(τ)a^(0)a^(0)a^(0)2,
where τ is the time delay between different detectors. In the following, we consider the zero-time delay correlation function in steady state for mode a
g(2)(0)=Tr[a^a^a^a^ρs]Tr[a^a^ρs]2,
where ρs is the steady-state density matrix calculated from Eq. (8). The second-order correlation function g(2)(0) < 1 corresponds to sub-poissonian statistics, which indicates photon antibunching.

3. The comparison of the optimal condition with numerical simulations for single drive

In the weak driving limit, the zero-time delay correlation function g(2)(0) can be given approximately by g(2)(0) ≃ 2|C200|2/|C100|4. This suggests that the conditions for g(2)(0) → 0 are equivalent to C200 = 0. So the optimal condition of photon antibunching can be obtained by solving the steady-state solution of the system and setting C200 = 0.

We can expand the wave function of the system on a Fock-state basis |nanbnc〉, where na, nb and nc denote the photon number in mode a, b and c, respectively. The system is truncated up to at most two photons, i.e., na + nb + 2nc ≤ 2. In this circumstances, the state of the system can be expanded as

|ψ=C000|000+C010|010+C100|100+C020|020+C200|200+C110|110+C001|001.
We take the non-Hermitian Hamiltonian H˜=H^diκ2(a^a^+b^b^)2iκc^c^ to describe the system evolution, where κc = 4κa = 4κb = 4κ. Substituting the state Eq. (11) and Hamiltonian into Schrödinger’s equation ih̄∂t|ψ〉 = |ψ〉, we get a set of equations for the coefficients
ih¯tC0000,ih¯tC010=iκ2C010+JC100+FC110,ih¯tC100=FC000+JC010iκ2C100+2FC200,ih¯tC020=iκC020+2JC110+2gC001,ih¯tC200=2FC100iκC200+2JC110,ih¯tC110=FC010+2JC020+2JC200iκC110,ih¯tC001=2gC0202iκC001.
Under the weak driving condition Fκ, we have |C000| ≫ {|C100|, |C010|} ≫ {|C200|, |C110|, |C020|, |C001|}. Thus the occupation number of the vacuum state is approximately equal to one, i.e., C000 = 1. For one-photon states, the steady-state coefficients are determined by
iκ2C010+JC100=0,F+JC010iκ2C100=0,
and for the coefficients of two-photon states, we have
iκC020+2JC110+2gC001=0,2FC100iκC200+2JC110=0,FC010+2JC020+2JC200iκC110=0,2gC0202iκC001=0.
The steady-state solution can be found by solving Eqs. (13) and Eqs. (14) as follows,
C010=4FJ4J2+κ2,C100=2iFκ4J2+κ2,C020=82F2J2κ2(4J2+κ2)(4J2κ2+κ4+g2(2J2+κ2)),C200=22F2(κ4+g2(2J2+κ2))(4J2+κ2)(4J2κ2+κ4+g2(2J2+κ2)),C110=8iF2Jκ(g2+κ2)(4J2+κ2)(4J2κ2+κ4+g2(2J2+κ2)),C001=8iF2gJ2κ(4J2+κ2)(4J2κ2+κ4+g2(2J2+κ2)).
To derive the condition of UPB in mode a, we set C200 = 0, and then we get the condition for optimal photon antibunching as
κ42g2J2+g2κ2=0.
When the above optimal condition for strong antibunching is satisfied, the UPB would occur.

In order to describe the UPB effect, we plot the second-order correlation function g(2)(0) as a function of the parameters in the system, this can be done by substituting Hamiltonian (Eq. (7)) into the master equation and solving Eq. (8) numerically. The Hilbert spaces of the system are truncated to five dimensions for cavity modes a, b and two dimensions for mode c. We rescale all parameters by the dissipation rate κ of the cavity modes in this paper.

In the derivation of the optimal photon antibunching condition, we have assumed that the detunings satisfy Δc = 2Δa = 2Δb = 2Δ = 0. In Fig. 2, we plot the numerical result for the zero-time second-order correlation functions g(2)(0) versus g/κ. We find that g(2)(0) ≪ 1 appears at g/κ ≃ ±0.1, which agrees with the analytic solution calculated by Eq. (16). And the position of g/κ where the correlation function arrives at its minimum would not be affected obviously with the change of Δ/κ. However, the value of g(2)(0) gets higher with the increase of Δ/κ. There is no obvious UPB phenomenon with Δa = 0.1. Although the antibunching effect is the strongest in the resonance case, the strict resonance condition is not necessary. The optimal analytical condition of Eq. (16) can be written in the form of

g/κ=±1(2(J/κ)21).
Clearly, the strong antibunching depends on the coupling strength J/κ. We can also learn from optimal condition Eq. (17) that real g and κ require J/κ(,1/2)(1/2,). Thus there will be no photon antibunching in the range J/κ[1/2,1/2] even if the optimal condition is satisfied.

 figure: Fig. 2

Fig. 2 The zero-delay-time second-order correlation functions g(2)(0) versus the nonlinear strength g/κ with driving strengths F/κ = 0.5, J/κ = 7 and the detuning Δ/κ = 0, Δ/κ = 0.05, Δ/κ = 0.1 respectively.

Download Full Size | PDF

In [23], the dependence of the photon antibunching is checked against the quality factor Qc, where the second-harmonic mode loss rate ranges from κcκ up to κc ≃ 102κ (κa = κb = κ), see Fig. 3(a). It is shown that the photon antibunching can be observed for both Qc > Q and Qc < Q (Qa = Qb = Q). The reason is as follows: in their scheme, the Hamiltonian in Eq. (7) is transformed into a Kerr-type nonlinear Hamiltonian as Ĥd = Δââ + Δ + h̄J(â + â) + h̄geff[ + b̂b̂] + h̄F(â + â), and the antibunching conditions are obtained based on this Hamiltonian. The mode c is decoupled from mode a and mode b, thus the Hamiltonian is equivalent to a two mode system Hamiltonian, which makes the influence of κc on g(2)(0) smaller and g(2)(0) ≪ 1 can be obtained even if κc changes on a large scale. In our scheme, g(2)(0) as a function of the quality factor Qc is also plotted, which is shown in Fig. 3(b). Different from the result in Fig. 3(a), the optimal value of photon antibunching appears at a fixed point QC/Q = 1/2. In fact, if we do not consider the condition κc = 4κ, a more general antibunching conditions can be obtained as 8g2J2 − 4g2κ2κ3κc = 0, which is different from the antibunching conditions in [23], because κc arise in the condition. The antibunching condition of Eq. (16) is satisfied only when κc = 4κ, so the optimum value of photon antibunching appeares at a fixed point QC/Q = 1/2.

 figure: Fig. 3

Fig. 3 Dependence of antibunching on the second-harmonic quality factor Qc. (a) In [23], the parameters F/κ = 1, g = 0.1 and J = 19.45. (b) In our scheme, F/κ = 1, g/κ = 0.1 and J/κ = J = 7.1.

Download Full Size | PDF

It is worth evaluating the actual experimental possibilities of the strong photon antibunching effect with weak nonlinearity. Here we choose the parameters in [23]. The present scheme could be realized with state-of-the-art technology employing the main III–V materials, such as GaAs [26, 27], GaN, and AIN [28], where bulk nonlinear susceptibility can be of the order of 10–200 pm/V in optoelectronics. A realistic order of magnitude estimates for the second-order nonlinearity is h̄g ≈ 1μ eV with engineered photonic crystal cavities, which is the same with that in [23]. Working in the typical telecommunication band, h̄ω ∼ 0.8eV, the loss rate is about h̄κ = 10μ eV corresponding to a fundamental mode Q factor Q ∼ 80000 [23], the Qc for the second harmonic mode is Qc ∼ 40000. These values could be routinely achieved in photonic crystal cavities made of III–V semiconductor materials [29]. The coupling strength is h̄J ≈ 71μ eV, which is smaller than that in [23], and it could be further reduced for larger g/κ according to Eq. (17). The time interval of antibunching occurs for UPB is limited by π/J [23].

4. The case with two drives

In this section, we investigate the photon antibunching with two drives. The Hamiltonian for such a system is Ĥd1 = Δaââ + Δb + Δcĉĉ + h̄J(â + â) + h̄g[()2ĉ + ĉ2)] + h̄F1(â + â) + h̄F2(ĉ + ĉ), where Δc = 2Δa = 2Δb = 2Δ, the driving strengths for mode a and mode c are denoted by F1 and F2, respectively. In the following discussion, we will set F1 = F, F2/F1 = n, where n is called strength ratio. The sign of n (minus or plus) can be tuned by adjusting the corresponding phases of the driving fields [14, 31].

Different from the case with single drive, the photon antibunching can be obtained in the mode a even if the detuning exists. The optimal strong photon antibunching conditions can be calculated as:

Δ[2gJ2n+F(4g2+7κ220Δ2)]=0,4Fg2J22Fg2κ22Fκ4+4gJ4n+gJ2κ2n+8Fg2Δ2+36Fκ2Δ24gJ2nΔ216FΔ4=0,
where we have considered the restrictions on the detunings, Δc = 2Δa = 2Δb = 2Δ. The photon antibunching would take place at the location where the requirements in Eqs. (18) meet.

For the sake of convenience, we rewrite Eqs. (18) as

{g1=20J2κ23κ4+112J2Δ224κ2Δ248Δ443J,n1=F[3(κ2+4Δ2)2+J2(10κ2+56Δ2)]23J34J2(5κ228Δ2)3(κ2+4Δ2)2,
{g2=20J2κ23κ4+112J2Δ224κ2Δ248Δ443J,n2=F[3(κ2+4Δ2)2+J2(10κ2+56Δ2)]23J34J2(5κ228Δ2)3(κ2+4Δ2)2.
The two group of solutions (g1, n1) and (g2, n2) are equivalent with Eqs. (18). The zero-time second-order correlation functions g(2)(0) are plotted in Fig. 4. In Fig. 4(a), the dips in the plot of g(2)(0) ≪ 1 correspond to two rings. This numerical condition agrees well with the analytical result given by
g/κ=±16[528(Δ/κ)2]3[1+4(Δ/κ)2]283.
We can also learn from optimal condition Eq. (21) that to have real g and κ, we must require Δ/κ in (−2.9382, −0.447545) ∪ (0.447545, 2.9382). Then the nonlinear interaction strength g/κ can only take values in [−2.16, 0) ∪ (0, 2.16]. In Fig. 4(b), the two symmetrical-quadratic-function-like analytical solution can be calculated as
n=±0.0036{3[1+4(Δ/κ)2]2+4[10+56(Δ2/κ)]}16[528(Δ/κ)2]3[1+4(Δ/κ)2]2,
this restricts Δ/κ to take values in (−2.9382, −0.447545) ∪ (0.447545, 2.9382).

 figure: Fig. 4

Fig. 4 (a) Logarithmic plot of the zero-time second-order correlation functions g(2)(0) as functions of detunings Δ/κ and the nonlinear interaction strength g/κ with ratio n given by n1 and n2. (b) Logarithmic plot of the zero-time second-order correlation functions g(2)(0) as functions of Δ/κ and n for the nonlinear interaction strength g/κ given by g1 and g2. In both (a) and (b), the tunnel coupling J/κ = 2 and F/κ = 0.1 are taken, and the dashed lines are analytical results.

Download Full Size | PDF

So far, we have investigated the photon antibunching driven by two fields with non-zero detunings Δ/κ, and we find the numerical simulation is in excellent agreement with the analytical solution. Next we will consider the case of resonance, i.e., the detunings are zero. In this case, the first equation of Eqs. (18) vanishes and Eqs. (18) reduce to

4Fg2J22Fg2κ22Fκ4+4gJ4n+gJ2κ2n=0.
To have the UPB with weak nonlinearity, we usually need a strong J/κ. For the single drive, there will be no photon antibunching in the range J/κ[1/2,1/2]. But for two drives, we will show that the UPB can be obtained even with weak nonlinearity and weak coupling strength, which is shown in Fig. 5. In addition, although the antibunching conditions are complicated for two drives compared with the single drive, both the driving strengths and the strength ratio n are included in the antibunching conditions and can be tuned precisely in experiments [14, 31], which makes our scheme available for tunable single photon sources.

 figure: Fig. 5

Fig. 5 The Logarithmic plot of g(2)(0) as functions of detunings J/κ for different n. The parameters F/κ = 0.5, g/κ = 0.1.

Download Full Size | PDF

The optimal analytical condition for strong photon antibunching depends on the strength ratio n

g/κ=(J/κ)2n4(J/κ)4n±8(F/κ)[2(F/κ)+4(F/κ)(J/κ)2]+[(J/κ)2n+4(J/κ)4n]24[F/κ+2(F/κ)(J/κ)2].
So the antibunching properties of the photons can be controlled by modulating the strength ratio. We find that the shapes of the valleys in the Logarithmic plot of the zero-time second-order correlation functions can be changed by modulating the parameter J/κ, which is shown in Fig. 6. In Fig. 6(a), by setting J/κ = 2, two valleys corresponding to g(2)(0) ≪ 1 appear in the Logarithmic plot of the zero-time second-order correlation functions. It is shown that the nonlinear interaction strength g/κ changes continuously with the ratio n in the photon antibunching region. The value of the nonlinear interaction strength is g/κ ≃ ±0.378 when the ratio is n = 0. When J/κ = 0.2, the shapes of two valleys change which is shown in Fig. 6(b). g/κ changes discontinuously with the ratio n for the photon antibunching region. And n can only take n ⊂ (−∞, −0.082] ∪ [0.082, ∞). We also compare the current scheme with that based on Kerr nonlinearity which is also with two drives [14]. For Kerr nonlinearity system, in order to make sure that the strong antibunching occurs in mode a in the weak nonlinear regime, the strength ratio n should be larger than one, but second-order nonlinearity system is not restricted by this condition. As the strength ratio n is not too large, e.g., n < 10, the antibunching effect is strong. With the increase of n, the strength of antibunching becomes weaker. But in second-order nonlinearity system, the influence of n to the value of g(2)(0) seems very small.

 figure: Fig. 6

Fig. 6 Logarithmic plot of the zero-time second-order correlation functions g(2)(0) as functions of the detunings g/κ and the strength ratio n. (a) The tunnel coupling J/κ = 2 and F/κ = 0.1. (b)The tunnel coupling J/κ = 0.2 and F/κ = 0.001.

Download Full Size | PDF

The system with two coupled cavities (with coupling strength J) for the case that both the two cavity modes are driven coherently have been demonstrated in different systems [14, 18, 21, 30], such as a quantum-dot-bimodal-cavity system [21], Kerr nonlinear system [14, 18, 30] and second-order nonlinearity system [25]. Here we also adopt photonic crystal cavities, and the second-order nonlinearity might take h̄g ≈ 1μ eV, the loss rate is about h̄κ = 10μ eV, h̄F = 5μeV, n = 0.329447 and h̄J = 3μ eV. The strength ratio n can be tuned precisely in experiments [14,31]. The coupling strength between the cavity modes depends exponentially upon the space of the cavities, and only a small J is needed to obtain antibunching. There is a range for J/κ to obtain antibunching when the others parameters are fixed. The range is called tolerance. When g(2)(0) < 0.5, antibunching can be obtained, the range of J/κ is about 0.1J/κ (the tolerance is about 0.1J/κ), which is shown in Fig. 5. The range depends on the effective nonlinearity g, which is the same with that in [23]. The tolerance is small due to a small J/κ, we can get a big tolerance when we take a big J/κ. To pump the two resonators with two distinct lasers, we might drive the mode a from the left side of the first cavity with frequency ω and drive the mode c from the right side of the second cavity with frequency 2ω, this scheme was used in [18]. In [25], second-order nonlinear system for two drives can be realized in a double quantum well embedded in a semiconductor microcavity, which might be a candidate to realize our scheme.

The physical principle of the photon antibunching is the quantum interference among multiple excitation pathways, which is shown in Fig. 7, where we show the energy-level diagram and the transition paths. There are three paths for two-photon excitation : (i) Directly input two photons in mode a, i.e., |000F1|100F1|200. (ii) Input one photon in mode a and swap it to mode b, input the second photon in mode a and the photon in mode b coming back to mode a, i.e., |000F1|100J|010F1|110J|200. (iii) Input one photon in mode c, transfer the photon to mode b and mode a by the second-order nonlinearity and the linear process, i.e. |000F2|001g|020J|110J|200. The photons coming from the three pathways take destructive interference when the antibunching condition Eqs. (18) are satisfied. The interference of the three excitation pathways leads to zero population of photon in the state |200〉, i.e., the photons can not occupy the state |200〉. This mechanism is different from the photon antibunching based on χ(3), in which the level structure has been changed.

 figure: Fig. 7

Fig. 7 Energy-level diagram showing the zero-, one-, and two-photon states (horizontal black short lines) and the transition paths leading to the quantum interference responsible for the strong antibunching (black lines with arrows). |nanbnc〉 represents the Fock state, where na, nb and nc denote respectively the photon number in mode a, b and c.

Download Full Size | PDF

5. Conclusion

In conclusion, we have studied the photon blockade caused by weak χ(2) nonlinearity in two coupled cavities. By solving the master equation in the steady-state limit and computing zero-delay-time second-order correlation function, strong photon antibunching is predicted in the first cavity. The exact numerical simulation for optimal antibunching agrees well with the analytical results. When both two cavities are driven, the ratio of the first driving strength to the second one becomes an additional parameter to modulate the strong photon antibunching. Thus we can always obtain the optimal antibunching independent of the detuning. We believe that our results may find new applications in generating single-photon sources due to the range of validity of the scheme.

APPENDIX: Non-zero Δ results in no photon antibunching conditions

In this section, we will prove that there are no antibunching conditions when the detuning is not zero. In the presence of Δ, the optimal strong photon antibunching conditions can be derived as

4g2+7κ220Δ2=0,2g2J2g2κ2κ4+4g2Δ2+18κ2Δ28Δ4=0,
where we set Δc = 2Δa = 2Δb = 2Δ. This condition is necessary, as mentioned in the maintext. Now we solve the equations and find 4 solutions for g and Δ,
{Δ1=5J23κ2+J2(25J2+72κ2)23,g1=25J236κ2+5J2(25J2+72κ2)23,
{Δ2=5J23κ2+J2(25J2+72κ2)23,g2=25J236κ2+5J2(25J2+72κ2)23,
{Δ3=5J23κ2+J2(25J2+72κ2)23,g3=25J236κ2+5J2(25J2+72κ2)23,
{Δ4=5J23κ2+J2(25J2+72κ2)23,g4=25J236κ2+5J2(25J2+72κ2)23.
To have a real gi, we must have
25J236κ2+5J2(25J2+72κ2)0.
By solving this inequality, we find that the condition of gi have a real solution is κ ≤ 0. Thus in the case with single drive, we only consider the resonant case, i.e., Δa = Δb = Δc = 0. Real g is necessary, otherwise the Hamiltonian is not hermitian(see the maintext).

Funding

This work is supported by the National Natural Science Foundation of China (NSFC) under Grants Nos. 11534002, 61475033 and 11204028.

References and links

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

2. A. Miranowicz, M. Paprzycka, Y. X. Liu, J. Bajer, and F. Nori, “Two-photon and three-photon blockades in driven nonlinear systems,” Phys. Rev. A 87, 023809 (2013). [CrossRef]  

3. K. M. Birnbaum, A. Boca, R. Miller, A. D. Boozer, T. E. Northup, and H. J. Kimble, “Photon blockade in an optical cavity with one trapped atom,” Nature 436, 87–90 (2005). [CrossRef]   [PubMed]  

4. A. Faraon, I. Fushman, D. Englund, N. Stoltz, P. Petroff, and J. Vučković, “Coherent generation of non-classical light on a chip via photon-induced tunnelling and blockade,” Nature Phys. 4, 859–863 (2008). [CrossRef]  

5. A. J. Hoffman, S. J. Srinivasan, S. Schmidt, L. Spietz, J. Aumentado, H. E. Tureci, and A. A. Houck, “Dispersive photon blockade in a superconducting circuit,” Phys. Rev. Lett. 107, 053602 (2011). [CrossRef]   [PubMed]  

6. Y. X. Liu, X. W. Xu, A. Miranowicz, and F. Nori, “From blockade to transparency: controllable photon transmission through a circuit-QED system,” Phys. Rev. A 89, 043818 (2014). [CrossRef]  

7. A. Majumdar and D. Gerace, “Single-photon blockade in doubly resonant nanocavities with second-order nonlinearity,” Phys. Rev. B 87, 235319 (2013). [CrossRef]  

8. H. Z. Shen, Y. H. Zhou, and X. X. Yi, “Quantum optical diode with semiconductor microcavities,” Phys. Rev. A 90, 023849 (2014). [CrossRef]  

9. D. E. Chang, A. S. Sorensen, E. A. Demler, and M. D. Lukin, “A single-photon transistor using nanoscale surface plasmons,” Nat. Phys. 3, 807–812 (2007). [CrossRef]  

10. D. Gerace, H. E. Tureci, A. Imamoglu, V. Giovannetti, and R. Fazio, “The quantum-optical josephson interferometer,” Nat. Phys. 5, 281–284 (2009). [CrossRef]  

11. T. C. H. Liew and V. Savona, “Single photons from coupled quantum modes,” Phys. Rev. Lett. 104, 183601 (2010). [CrossRef]   [PubMed]  

12. H. J. Carmichael, “Photon antibunching and squeezing for a single atom in a resonant cavity,” Phys. Rev. Lett. 55, 2790 (1985). [CrossRef]   [PubMed]  

13. M. Bamba, A. Imamoglu, I. Carusotto, and C. Ciuti, “Origin of strong photon antibunching in weakly nonlinear photonic molecules,” Phys. Rev. A 83, 021802 (2011). [CrossRef]  

14. X. W. Xu and Y. Li, “Tunable photon statistics in weakly nonlinear photonic molecules,” Phys. Rev. A 90, 043822 (2014). [CrossRef]  

15. X. W. Xu and Y. Li, “Strong photon antibunching of symmetric and antisymmetric modes in weakly nonlinear photonic molecules,” Phys. Rev. A 90, 033809 (2014). [CrossRef]  

16. S. Ferretti, V. Savona, and D. Gerace, “Optimal antibunching in passive photonic devices based on coupled nonlinear resonators,” New. J. Phys. 15, 025012 (2013) [CrossRef]  

17. H. Z. Shen, Y. H. Zhou, and X. X. Yi, “Tunable photon blockade in coupled semiconductor cavities,” Phys. Rev. A 91, 063808 (2015). [CrossRef]  

18. S. Ferretti, L. C. Andreani, H. E. Tureci, and D. Gerace, “Photon correlations in a two-site nonlinear cavity system under coherent drive and dissipation,” Phys. Rev. A 82, 013841 (2010) [CrossRef]  

19. X. W. Xu and Y. J. Li, “Antibunching photons in a cavity coupled to an optomechanical system,” J. Opt. B: At. Mol. Opt. Phys. 46, 035502 (2013).

20. A. Majumdar, M. Bajcsy, A. Rundquist, and J. VuČković, “Loss-enabled sub-poissonian light generation in a bimodal nanocavity,” Phys. Rev. Lett. 108, 183601 (2012). [CrossRef]   [PubMed]  

21. W. Zhang, Z. Y. Yu, Y. M. Liu, and Y. W. Peng, “Optimal photon antibunching in a quantum-dot-bimodal-cavity system,” Phys. Rev. A 89, 043832 (2014). [CrossRef]  

22. Y. H. Zhou, H. Z. Shen, and X. X. Yi, “Unconventional photon blockade with second-order nonlinearity,” Phys. Rev. A 92, 023838 (2015). [CrossRef]  

23. D. Gerace and V. Savona, “Unconventional photon blockade in doubly resonant microcavities with second-order nonlinearity,” Phys. Rev. A 89, 031803 (2014). [CrossRef]  

24. O. Kyriienko and T. C. H. Liew, “Triggered single-photon emitters based on stimulated parametric scattering in weakly nonlinear systems,” Phys. Rev. A 90, 063805 (2014). [CrossRef]  

25. O. Kyriienko, I. A. Shelykh, and T. C. H. Liew, “Tunable single-photon emission from dipolaritons,” Phys. Rev. A 90, 033807 (2014). [CrossRef]  

26. R. W. Boyd, Nonlinear Optics (Academic, 2008).

27. S. Bergfeld and W. Daum, “Second-harmonic generation in GaAs: experiment versus theoretical predictions of χxyh(2),” Phys. Rev. Lett. 90, 036801 (2003). [CrossRef]  

28. J. Chen, Z. H. Levine, and J. W. Wilkins, “Calculated second-harmonic susceptibilities of BN, AlN, and GaN,” Appl. Phys. Lett. 66, 1129 (1995). [CrossRef]  

29. S. Combrié, A. De Rossi, N.-Q.-V. Tran, and H. Benisty, “GaAs photonic crystal cavity with ultrahigh Q: microwatt nonlinearity at 1.55μ m,” Opt. Lett. 33, 1908 (2008). [CrossRef]  

30. H. Flayac and V. Savona, “Input-output theory of the unconventional photon blockade,” Phys. Rev. A 88, 033836 (2013). [CrossRef]  

31. H.-A. Bachor and T. C. Ralph, A Guide to Experiments in Quantum Optics (Wiley VCH, 2004).

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 (7)

Fig. 1
Fig. 1 The system of two coupled cavities. Modes a and b in the cavities are tunnel-coupled, mode b and mode c are coupled via the second-order nonlinear χ(2).
Fig. 2
Fig. 2 The zero-delay-time second-order correlation functions g(2)(0) versus the nonlinear strength g/κ with driving strengths F/κ = 0.5, J/κ = 7 and the detuning Δ/κ = 0, Δ/κ = 0.05, Δ/κ = 0.1 respectively.
Fig. 3
Fig. 3 Dependence of antibunching on the second-harmonic quality factor Qc. (a) In [23], the parameters F/κ = 1, g = 0.1 and J = 19.45. (b) In our scheme, F/κ = 1, g/κ = 0.1 and J/κ = J = 7.1.
Fig. 4
Fig. 4 (a) Logarithmic plot of the zero-time second-order correlation functions g(2)(0) as functions of detunings Δ/κ and the nonlinear interaction strength g/κ with ratio n given by n1 and n2. (b) Logarithmic plot of the zero-time second-order correlation functions g(2)(0) as functions of Δ/κ and n for the nonlinear interaction strength g/κ given by g1 and g2. In both (a) and (b), the tunnel coupling J/κ = 2 and F/κ = 0.1 are taken, and the dashed lines are analytical results.
Fig. 5
Fig. 5 The Logarithmic plot of g(2)(0) as functions of detunings J/κ for different n. The parameters F/κ = 0.5, g/κ = 0.1.
Fig. 6
Fig. 6 Logarithmic plot of the zero-time second-order correlation functions g(2)(0) as functions of the detunings g/κ and the strength ratio n. (a) The tunnel coupling J/κ = 2 and F/κ = 0.1. (b)The tunnel coupling J/κ = 0.2 and F/κ = 0.001.
Fig. 7
Fig. 7 Energy-level diagram showing the zero-, one-, and two-photon states (horizontal black short lines) and the transition paths leading to the quantum interference responsible for the strong antibunching (black lines with arrows). |nanbnc〉 represents the Fock state, where na, nb and nc denote respectively the photon number in mode a, b and c.

Equations (30)

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

D i ( r , t ) = ε 0 ε i j E j ( r , t ) + ε 0 [ χ i j k ( 2 ) ( r ) E j ( r , t ) E k ( r , t ) ] +
E ^ ( r , t ) = j = a , b , c h ¯ ω j 2 0 [ j ^ ϕ j ( r ) ε ( r ) e i ω j t + H . C . ] ,
H ^ em = h ¯ ω a a ^ a + h ¯ ω b b ^ b ^ + h ¯ ω c c ^ c ^ + h ¯ J ( a ^ b ^ + b ^ a ^ ) + h ¯ g [ ( b ^ ) 2 c ^ + c ^ b ^ 2 ] + h ¯ v [ ( a ^ ) 2 c ^ + c ^ a ^ 2 ) ] + h ¯ u [ a ^ b ^ c ^ + c ^ b ^ a ^ ] ,
h ¯ J = h ¯ 2 ω a ω b ε i j ( r ) ε ( r ) ϕ a ( r ) ϕ b ( r ) d 3 r , h ¯ g = h ¯ ω b 2 2 ω c 2 0 χ ( 2 ) ( r ) [ ε ( r ) ] 3 / 2 ϕ b ( r ) ϕ b ( r ) ϕ c ( r ) d 3 r , h ¯ v = h ¯ ω a 2 2 ω c 2 0 χ ( 2 ) ( r ) [ ε ( r ) ] 3 / 2 ϕ a ( r ) ϕ a ( r ) ϕ c ( r ) d 3 r , h ¯ u = 2 ε 0 [ h ¯ 2 ε 0 ] 3 / 2 ω a ω b ω c χ ( 2 ) ( r ) [ ε ( r ) ] 3 / 2 ϕ a ( r ) ϕ b ( r ) ϕ c ( r ) d 3 r .
H ^ 0 = h ¯ ω a a ^ a ^ + h ¯ ω b b ^ b ^ + h ¯ ω c c ^ c ^ + h ¯ J ( a ^ b ^ + b ^ a ^ ) + h ¯ g [ ( b ^ ) 2 c ^ + c ^ b ^ 2 ] .
H ^ = H ^ 0 + h ¯ F ( a ^ e i ω L t + a ^ e i ω L t ) .
H ^ d = h ¯ Δ a a ^ a ^ + h ¯ Δ b b ^ b ^ + h ¯ Δ c c ^ c ^ + h ¯ J ( a ^ b ^ + b ^ a ^ ) + h ¯ g [ ( b ^ ) 2 c ^ + c ^ b ^ 2 ] + h ¯ F ( a ^ + a ^ ) ,
ρ ^ t = i h ¯ [ H ^ d , ρ ^ ] + κ a 2 ( 2 a ^ ρ ^ a ^ + a ^ a ^ ρ ^ ρ a ^ a ^ ) + κ b 2 ( 2 b ^ ρ ^ b ^ b ^ b ^ ρ ^ ρ b ^ b ^ ) + κ c 2 ( 2 c ρ ^ c ^ c ^ c ^ ρ ^ ρ c ^ c ^ ) ,
g ( 2 ) ( τ ) = a ^ ( 0 ) a ^ ( τ ) a ^ ( τ ) a ^ ( 0 ) a ^ ( 0 ) a ^ ( 0 ) 2 ,
g ( 2 ) ( 0 ) = Tr [ a ^ a ^ a ^ a ^ ρ s ] Tr [ a ^ a ^ ρ s ] 2 ,
| ψ = C 000 | 000 + C 010 | 010 + C 100 | 100 + C 020 | 020 + C 200 | 200 + C 110 | 110 + C 001 | 001 .
i h ¯ t C 000 0 , i h ¯ t C 010 = i κ 2 C 010 + J C 100 + F C 110 , i h ¯ t C 100 = F C 000 + J C 010 i κ 2 C 100 + 2 F C 200 , i h ¯ t C 020 = i κ C 020 + 2 J C 110 + 2 g C 001 , i h ¯ t C 200 = 2 F C 100 i κ C 200 + 2 J C 110 , i h ¯ t C 110 = F C 010 + 2 J C 020 + 2 J C 200 i κ C 110 , i h ¯ t C 001 = 2 g C 020 2 i κ C 001 .
i κ 2 C 010 + J C 100 = 0 , F + J C 010 i κ 2 C 100 = 0 ,
i κ C 020 + 2 J C 110 + 2 g C 001 = 0 , 2 F C 100 i κ C 200 + 2 J C 110 = 0 , F C 010 + 2 J C 020 + 2 J C 200 i κ C 110 = 0 , 2 g C 020 2 i κ C 001 = 0 .
C 010 = 4 F J 4 J 2 + κ 2 , C 100 = 2 i F κ 4 J 2 + κ 2 , C 020 = 8 2 F 2 J 2 κ 2 ( 4 J 2 + κ 2 ) ( 4 J 2 κ 2 + κ 4 + g 2 ( 2 J 2 + κ 2 ) ) , C 200 = 2 2 F 2 ( κ 4 + g 2 ( 2 J 2 + κ 2 ) ) ( 4 J 2 + κ 2 ) ( 4 J 2 κ 2 + κ 4 + g 2 ( 2 J 2 + κ 2 ) ) , C 110 = 8 i F 2 J κ ( g 2 + κ 2 ) ( 4 J 2 + κ 2 ) ( 4 J 2 κ 2 + κ 4 + g 2 ( 2 J 2 + κ 2 ) ) , C 001 = 8 i F 2 g J 2 κ ( 4 J 2 + κ 2 ) ( 4 J 2 κ 2 + κ 4 + g 2 ( 2 J 2 + κ 2 ) ) .
κ 4 2 g 2 J 2 + g 2 κ 2 = 0 .
g / κ = ± 1 ( 2 ( J / κ ) 2 1 ) .
Δ [ 2 g J 2 n + F ( 4 g 2 + 7 κ 2 20 Δ 2 ) ] = 0 , 4 F g 2 J 2 2 F g 2 κ 2 2 F κ 4 + 4 g J 4 n + g J 2 κ 2 n + 8 F g 2 Δ 2 + 36 F κ 2 Δ 2 4 g J 2 n Δ 2 16 F Δ 4 = 0 ,
{ g 1 = 20 J 2 κ 2 3 κ 4 + 112 J 2 Δ 2 24 κ 2 Δ 2 48 Δ 4 4 3 J , n 1 = F [ 3 ( κ 2 + 4 Δ 2 ) 2 + J 2 ( 10 κ 2 + 56 Δ 2 ) ] 2 3 J 3 4 J 2 ( 5 κ 2 28 Δ 2 ) 3 ( κ 2 + 4 Δ 2 ) 2 ,
{ g 2 = 20 J 2 κ 2 3 κ 4 + 112 J 2 Δ 2 24 κ 2 Δ 2 48 Δ 4 4 3 J , n 2 = F [ 3 ( κ 2 + 4 Δ 2 ) 2 + J 2 ( 10 κ 2 + 56 Δ 2 ) ] 2 3 J 3 4 J 2 ( 5 κ 2 28 Δ 2 ) 3 ( κ 2 + 4 Δ 2 ) 2 .
g / κ = ± 16 [ 5 28 ( Δ / κ ) 2 ] 3 [ 1 + 4 ( Δ / κ ) 2 ] 2 8 3 .
n = ± 0.0036 { 3 [ 1 + 4 ( Δ / κ ) 2 ] 2 + 4 [ 10 + 56 ( Δ 2 / κ ) ] } 16 [ 5 28 ( Δ / κ ) 2 ] 3 [ 1 + 4 ( Δ / κ ) 2 ] 2 ,
4 F g 2 J 2 2 F g 2 κ 2 2 F κ 4 + 4 g J 4 n + g J 2 κ 2 n = 0 .
g / κ = ( J / κ ) 2 n 4 ( J / κ ) 4 n ± 8 ( F / κ ) [ 2 ( F / κ ) + 4 ( F / κ ) ( J / κ ) 2 ] + [ ( J / κ ) 2 n + 4 ( J / κ ) 4 n ] 2 4 [ F / κ + 2 ( F / κ ) ( J / κ ) 2 ] .
4 g 2 + 7 κ 2 20 Δ 2 = 0 , 2 g 2 J 2 g 2 κ 2 κ 4 + 4 g 2 Δ 2 + 18 κ 2 Δ 2 8 Δ 4 = 0 ,
{ Δ 1 = 5 J 2 3 κ 2 + J 2 ( 25 J 2 + 72 κ 2 ) 2 3 , g 1 = 25 J 2 36 κ 2 + 5 J 2 ( 25 J 2 + 72 κ 2 ) 2 3 ,
{ Δ 2 = 5 J 2 3 κ 2 + J 2 ( 25 J 2 + 72 κ 2 ) 2 3 , g 2 = 25 J 2 36 κ 2 + 5 J 2 ( 25 J 2 + 72 κ 2 ) 2 3 ,
{ Δ 3 = 5 J 2 3 κ 2 + J 2 ( 25 J 2 + 72 κ 2 ) 2 3 , g 3 = 25 J 2 36 κ 2 + 5 J 2 ( 25 J 2 + 72 κ 2 ) 2 3 ,
{ Δ 4 = 5 J 2 3 κ 2 + J 2 ( 25 J 2 + 72 κ 2 ) 2 3 , g 4 = 25 J 2 36 κ 2 + 5 J 2 ( 25 J 2 + 72 κ 2 ) 2 3 .
25 J 2 36 κ 2 + 5 J 2 ( 25 J 2 + 72 κ 2 ) 0 .
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.