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

Squeezed state generation in photonic crystal microcavities

Open Access Open Access

Abstract

The feasibility of using a parametric down-conversion process to generate squeezed electromagnetic states in three dimensional photonic crystal microcavity structures is investigated for the first time. The spectrum of the squeezed light is theoretically calculated by using an open cavity quantum mechanical formalism. The cavity communicates with two main channels, which model vertical radiation losses and coupling into a single-mode waveguide respectively. The amount of squeezing is determined by the correlation functions relating the field quadratures of light coupled into the waveguide. All of the relevant model parameters are realistically estimated for structures made in Al0.3Ga0.7As, using finite-difference time-domain simulations. Squeezing up to ~30% below the shot noise level is predicted for 10 mW average power, 80 MHz repetition, 500 ps excitation pulses using in a [111] oriented wafer.

©2008 Optical Society of America

1. Introduction

Deterministic and efficient nonclassical light sources play a central role in implementing future quantum information protocols [1, 2]. Two examples are single photon and squeezed state light sources. Optimum implementations of quantum cryptography [3] and quantum computation with linear optics [4] rely on single photon sources whereas quantum information processing with continuous variables mainly depends on squeezed state sources [5, 6, 7], which produce two strongly correlated (entangled) light beams. Semiconductor-based implementations of these nonclassical light sources are particularly attractive due to their small size, low power consumption and the possibility of integration with solid-state qubits [8, 9].

In the first protocol proposed for continuous variable teleportation [6], the entangled EPR (Einstein, Podolsky, and Rosen) beams were generated by combining two independent squeezed beams produced by a parametric down-conversion process in a bulk nonlinear crystal. The work described below investigates the feasibility of miniaturizing the squeezed state generation process by monolithically integrating a three dimensionally (3D)-confined photonic crystal microcavity with a single-mode ridge waveguide in a sub-micron thick slab of AlGaAs. This was motivated by quantitative measurements of second-order sum-frequency generation in an InP membrane-based microcavity [10], which suggested that the inverse process of parametric down-conversion in the microcavity could possibly be used as a practical, ultra-miniature source of squeezed light if it can be efficiently coupled to a single mode waveguide. Already there have been reported attempts to use one-dimensional (1D) photonic crystal slabs and 2D photonic crystal waveguides to generate squeezing [11] and entangled photon pairs [12, 13, 14], but as far as we know this is the first attempt to quantify the squeezing process in cubicwavelength scale, 3D photonic crystal microcavity structures.

Due to the nonclassical characteristics of squeezed light, a quantum mechanical model is needed to describe the down-conversion process in the cavity, and how it couples to various leakage channels. The overall aim of this work was to obtain a quantitative estimate for the (nonclassical) spectrum of radiation that propagates away from a 3D, wavelength-scale microcavity that supports a distinct localized electromagnetic mode with frequency at ω 1, and which is pumped/excited by a classical source at 2ω 1. We adapt a system-reservoir quantization formalism from Ref.[15], that allows us to rigorously relate the output channels’ quantum field variables to those of the input channels, and the leakage from the excited mode in the cavity (the so-called input-output relations [16, 17, 18, 19]).

In the following, an interaction Hamiltonian due to the second-order nonlinear coupling between photons of the fundamental cavity mode (degenerate down-conversion) is added to the system-reservoir Hamiltonian and all the appropriate approximations are explained explicitly. The crucial component of this model, that must be solved numerically, is the overlap integral describing the nonlinear interaction of the classical excitation field at 2ω 1, with the quantized cavity mode. This is evaluated using a finite-difference time-domain (FDTD) solver [20] to find the relevant amplitudes for all the fields involved in the parametric down-conversion process. All of these calculations are done using a photonic crystal cavity coupled to a single-mode ridge waveguide following a design that was recently fabricated and shown to exhibit coupling efficiencies in excess of 55% [21].

2. Field quantization in an open optical cavity

The system of interest consists of a defect-state optical microcavity in a thin semiconductor membrane perforated by a regular 2D array of through-holes, Fig. 1(a). The light is confined in the cavity by Bragg reflection in the plane (due to the photonic crystal bandgap) and conventional total internal reflection in the vertical direction. For appropriately designed defect states and large enough surrounding photonic crystal material, the intrinsic loss of the cavity is due entirely to out-of-plane scattering [22, 23]. When a 1D single-mode waveguide is introduced in proximity to the microcavity, as in Fig. 1(b), it is possible to have the coupling of the cavity mode(s) to this waveguide channel dominate the total leakage. It was shown that preferential coupling to a single-mode output channel can account for up to 90% of the total leakage [24].

For modeling purposes, this system is conceptualized as consisting of a cavity field that can be driven by polarization induced in the microcavity by classical and/or quantum field sources incident via the continua channels (either from the top or bottom half space radiation modes, or other in-plane waveguide channel modes). Each continuum channel carries away radiation excited by the incident fields, possibly mediated by resonant scattering in the cavity. Using the system-reservoir formalism [15], the cavity is considered as a system with a discrete set of quantized modes (associated with normal modes of an isolated cavity), and the reservoir consists of a set of continuum field operators which model all the distinct channels that the cavity can leak into.

 figure: Fig. 1.

Fig. 1. 2D photonic crystal microcavities, a) an isolated cavity and b) adding a 1D waveguide channel to the cavity structure.

Download Full Size | PDF

Here we introduce the key elements of the system-reservoir Hamiltonian formalism, which is derived using the Feshbach projection technique [25, 26]. The exact eigenmodes of the vector potential for the multi-mode photonic crystal microcavity and several continuum channels that couple to the cavity, f m(r,ω), are expressed in terms of the discrete eigenmodes of the cavity, U λ (r), and the continuum of modes, V n(r,ω′), in the channels [15, 27]:

fm(r,ω)=λαλm(ω)Uλ(r)+ndωβnm(ω,ω)Vn(r,ω).

The discrete indexm labels an asymptotic solution in which an incoming wave through channel m is scattered by the structure into all other channels. Here U λ (r) is a localized cavity mode with a discrete spectrum, and V n(r,ω′) is a continuum eigenmode of channel n. These cavity and channel modes are orthogonal and they have nonzero values only in the appropriate regions (i.e. inside or outside of the [arbitrary] 3D boundary that delineates the cavity). The coefficients αm λ and βm n (ω,ω′) are the coupling coefficients between the cavity/channel modes and the exact eigenmode associated with excitation of the system through channel m.

The quantization of the fields may be achieved by expanding the vector potential in a complete set of mode functions and imposing canonical commutation relations for the expansion coefficients [15, 28]. By using the eigenmodes of the structure, Eq.(1), the Hamiltonian of the electromagnetic field in a linear medium becomes:

Ĥ=λh̅ωλâλâλ+mdωh̅ωr̂m(ω)r̂m(ω)
+h̅λmdω[Wλm(ω)âλr̂m(ω)+Wλm*(ω)âλr̂m(ω)
+Tλm(ω)âλr̂m(ω)+Tλm*(ω)âλr̂m(ω)],

where â λ and â λ are the creation and annihilation operators of the discrete cavity modes, whereas r̂m and r̂ m are the corresponding operators of the continuum modes of channel m. In addition, Wλm and Tλm are the coupling coefficients between the cavity and channel operators at the interface that defines the cavity region [15]. Equation (2) exactly describes the electromagnetic degrees of freedom in a textured dielectric medium. Its utility, for structures designed to locally support discrete modes coupled preferentially to a set of waveguide continua channels, is that it represents a rigorous formulation in terms of heuristic localized and continua mode amplitudes, regardless of whether or not the last term can be approximated as a weak damping contribution to the localized mode amplitude(s). Although the following derivation will only utilize this weak coupling limit, Eq. (2) provides a much more general starting point for analysis of practical systems that may well not satisfy the associated approximations.

Equation (2) contains resonant (â r̂, r̂ â) and nonresonant (â r̂, â r̂) terms. In the weak coupling limit, when the photonic crystal microcavity modes have high quality factors, a rotating-wave approximation can be used and the nonresonant terms in the Hamiltonian can be ignored.

3. Parametric down-conversion in an open cavity

To address the parametric down-conversion process inside a single-mode cavity with a non-zero second-order susceptibility, χ (2), the Hamiltonian in Eq. (2) has to include an additional term:

ĤI=2ε03d3rÊ(r,t).χ(2)(r):Ê(r,t)Ê(r,t),

where Ê(r,t), the total field in the cavity region, is the superposition of the fields associated with the cavity mode and the pump field. The quantized field due to the cavity mode can be written as

Êc(r,t)=ih̅ω12ε0[â1(t)U1(r)â1(t)U1*(r)].

The pump field, which is taken to be a classical coherent state at 2ω 0 (assuming ω 0ω 1, where ω 1 is the cavity mode frequency) is written as:

Ep(r,t)=iAp(t)[Up(r)e2iω0tUp*(r)e+2iω0t],

where Ap and U p(r) are the envelope function and the normalized spatial distribution of the pump field in the cavity region respectively. By substituting the total field Ê(r,t)=Êc(r,t)+E p(r,t) in the Hamiltonian of Eq.(3) and only considering the resonant terms we have

ĤI=ih̅[gâ1(t)â1(t)e2iω0tg*â1(t)â1(t)e2iω0t],

where

g=ω13Ap(t)d3r[U1.χ(2):U1Up+U1.χ(2):UpU1+Up.χ(2):U1U1].

Here we suppose the spatial distribution of the fields in Eq.(7) are real valued functions and therefore g is a real number. By adding the Hamiltonian of the nonlinear interaction, Eq.(6), to the original system-reservoir Hamiltonian (of a single-mode cavity), Eq.(2), and Fourier transforming the Heisenberg equation of motion for the cavity mode we get:

a˜1(Ω)=2g2γ1r˜1in(Ω)+2gm=22γmr˜min(Ω)[Γi(Δ+Ω)][Γ+i(ΔΩ)]4g2
+[Γ+i(ΔΩ)][2γ1r˜1in(Ω)+m=22γmr˜min(Ω)][Γi(Δ+Ω)][Γ+i(ΔΩ)]4g2,

where we used a rotating frame in which, a1(t)=a˜1(t)eiω0t, and Δ=ω 0-ω 1 is the offset of the cavity mode with respect to the half of the pump frequency. Here γm=π|W 1m|2 (m=1, …) represents damping of the cavity mode due to the coupling into channel m and Γ=Σm γm accounts for the total loss of the cavity mode.

We made two assumptions to derive the above equation of motion, Eq.(8). First we supposed that due to the weak coupling of the cavity mode to the reservoir, its eigenfrequency is not affected by the reservoir. The second assumption is the Markov approximation (reservoir correlation times are assumed negligibly short compared to the characteristic time scale associated with the cavity mode dynamics [17]), in which we assume that the coupling coefficients γm(ω) are independent of frequency.

In writing Eq. (8), we separated the channels that a microcavity mode can communicate with into two parts. Channel 1 is a 1D single-mode waveguide and r̃m (m=2, …) are the rest of the reservoir operators which the cavity modes can couple with, Fig. (2).

 figure: Fig. 2.

Fig. 2. The model cavity that communicates with several output channels.

Download Full Size | PDF

To measure the degree of squeezing in the output channel 1, we have to evaluate the following correlation functions [17, 28]:

SX(Ω)=<Xout(Ω),Xout(Ω)>1,
SY(Ω)=<Yout(Ω),Yout(Ω)>1,

where SX (Ω) and SY (Ω) are called the spectrum of squeezing for the X and Y field quadratures respectively, with

Xout(Ω)=r˜1out(Ω)+r˜1out(Ω),
Yout(Ω)=i[r˜1out(Ω)r˜1out(Ω)].

The usual derivation of SX and SY [16, 17] for a single continuum channel combines Eq.(8) with the input-output relation [15, 16]

r˜1out(Ω)r˜1in(Ω)=2γ1a˜1(Ω),

and uses the fact that for a reservoir in the vacuum state <rin 1(Ω)rin 1(Ω)>=1, and <rin 1 (Ω)rin 1 (Ω)>=<rin 1(Ω)rin 1(Ω)>=<rin 1(Ω)rin 1(Ω)>=0. The result is

SX(Ω)=8gγ1(γ12g)2+Ω2,

and

SY(Ω)=8gγ1(γ1+2g)2+Ω2.

By using Eq.(8) with the generalized input-output relation,

r˜mout(Ω)r˜min(Ω)=2γ1a˜1(Ω),

and the generalized expectation values for reservoir vacuum fields (<rinn(Ω)rin m(Ω)>=δnm and <rinn(Ω)rinm(Ω)>=<rin n(Ω)rin m(Ω)>=<rin n(Ω)rinm(Ω)>=0), one gets

SX(Ω)=8gγ1[(γ1+γx)2g]2+Ω2,

and

SY(Ω)=8gγ1[(γ1+γx)+2g]2+Ω2,

where

γx=m=2γm,

Before we show the numerical calculation of the spectrum of squeezing in a parametric down conversion process generated in a photonic crystal microcavity, it is useful to compare the squeezing spectra for the single and multiple channel cases in the limit of zero detuning. Figure 3(a) shows the spectrum of squeezing for the Y quadrature for two different cases, when there is only one output channel, the waveguide channel (solid line), and when loss through channels 1 and all other channels are equal, for the threshold situation when 2g=γ 1+γx. When the nonlinear cavity is coupled only to a single-mode channel, S Y=-1 at threshold when Ω=0. In this situation there is a perfect correlation between the Y field quadratures of the degenerate photons at the waveguide output [31, 32, 33]. By plotting Eq. (17) versus g for Ω=0, Fig. 3(b), in the single channel case (solid), and the case where there is equal coupling to the waveguide and all other channels (dashed), we see that the degree of squeezing is always less in the latter case, and at threshold the maximum amount of squeezing is

SY(Ω=0)=γ1γ1+γx,

rather than unity. In another words the squeezing degree is equal to the coupling efficiency into the single-mode waveguide channel.

 figure: Fig. 3.

Fig. 3. a) Spectrum at threshold of squeezing for the Y quadrature in a degenerate down-conversion process. b) Squeezing versus g factor at Ω=0. The solid lines are for the case when the cavity couples to a single channel (γx=0) and the dashed lines are when γx=γ 1.

Download Full Size | PDF

4. Numerical estimation of the squeezing spectrum

Crucial steps in quantitatively estimating the degree of squeezing expected for a realistic cavity/waveguide system, excited by a realistic pump source involve obtaining numerical estimates of g (the nonlinear coupling coefficient in Eq.(7)), and γ 1 and γx (cavity mode losses through channel 1 and all other channels). To estimate g for a particular microcavity and host material, one needs to determine the tensor χ ⃡(2), and the 3D field profiles in the cavity region corresponding to the localized microcavity mode and the pump field.

Here we consider a “3-missing hole” cavity [34] made in an Al 0.3Ga0.7As membrane (the refractive index, n=3.23 around 1.5 µm). The concentration of Al in Al 0.3Ga0.7As is chosen such that its bandgap [35] is at 1.798 ev (or 689 nm), so that the pump beam (at ~720 nm, twice the cavity mode frequency) does not generate any free carriers. Free carrier generation in the cavity region changes the refraction index of the medium and induces undesired losses due to free carrier absorption (Any loss in the waveguide-cavity structure introduces uncorrelated fluctuations, which degrade the squeezing degree [16]. The overall loss can be modeled as an extra channel that the light can couple into). Free carriers can still be generated by two photon absorption of the pump, but that is a weaker process and can’t be easily avoided using conventional III–V semiconductors.

The value of the g coefficient, Eq.(7), depends on the χ ˡ(2) symmetry of the host photonic crystal material. We used a coordinate system in which the microcavity and incident pump field polarization are in the x-y plane. Al0.3Ga0.7As belongs to the 4̄3m symmetry group and if the wafer used to fabricate the microcavity is oriented in the [001] direction, then χ ˡ(2) has the form

χ[001](2)(ω1;2ω0,ω1)=(χxxx(2)χxxy(2)χxxz(2)χxyx(2)χxyy(2)χxyz(2)χxzx(2)χxzy(2)χxzz(2)χyxx(2)χyxy(2)χyxz(2)χyyx(2)χyyy(2)χyyz(2)χyzx(2)χyzy(2)χyzz(2)χzxx(2)χzxy(2)χzxz(2)χzyx(2)χzyy(2)χzyz(2)χzzx(2)χzzy(2)χzzz(2))
=β(000001010001000100010100000),

where β=216 pm/V for ω 1~1.5 µm [36]. The sparse nature of this tensor restricts the possible orientations of the photonic crystal with respect to the electronic axes of the host crystal. If a transverse pump field is incident normal to a [001] oriented crystal, only a z-oriented second order polarization can be generated, which does not couple to the microcavity modes that derive from TE-polarized slab modes, and are therefore polarized in-plane. To reduce these restrictions, we consider wafers grown along [111] direction. The χ ˡ(2) tensor for wafers oriented along [111] is of the form:

χ[111](2)(ω1;2ω0,ω1)=β(330223013022301223013013023023022301223023012230323)

The nonlinear coupling coefficient also depends on the spatial distribution of the cavity mode and the pump laser, which can be determined using an FDTD simulation [20], as described in the next section.

5. Degenerate parametric down-conversion

We consider a 3-missing hole cavity in an hexagonal 2D photonic crystal structure to estimate the squeezed spectrum in the case of degenerate parametric down-conversion involving a single cavity mode pumped at twice its natural frequency. The structure is shown schematically in Fig. 4(a), it has a lattice constant of a=420 nm, an air hole radius r=0.29a, and the semiconductor membrane thickness is h=200 nm. Three holes next to the cavity are shifted to increase the cavity Q factor [37]. The shift at the A, B, and C locations are 0.2a, 0.025a, and 0.2a respectively. The cavity is tilted with respect to the waveguide axis, Fig. 4(b), to increase its coupling efficiency to the waveguide [24].

The cavity shows a fundamental mode at f 1=207.98 THz (1.44 µm). The spatial distribution of the cavity mode and the pump beam at 2f 1=415.96 THz (~721 nm), with the same polarization as the cavity mode and focussed to a spot size of 2.5 µm, are shown in Fig. (5). The quality factor of both an isolated cavity and one connected to the single-mode ridge waveguide as in Fig. 4 (b), are found to be Qi=79420, and QT=20940. By using

1QT=1Qi+1Qwg,

we can extract the quality factor of the cavity mode due to its coupling with the waveguide, which is Qwg=28440. The values of the cavity losses for the model calculation are thus γ1

γ1=ω1Qwg=46GHz,γx=ω1Qi=16GHz.

The design and properties of the ridge-waveguide and its interface with the PhC waveguide is described in Ref. [21]. If a CW excitation beam with an average power of 10 mW is assumed, Eq. (7) yields

g[111]=0.839GHz.

By using 500 ps pulses (the linewidth of the cavity mode is ~16 larger than the pump laser linewidth in this case.) with 80 MHz repetition rate and 10mWaverage power, the g coefficients for the [111] direction increases to:

g[111]=4.199GHz.

Figure (6) shows the calculated spectrum of squeezing of the Y quadrature for the [111] growth direction for the 500 ps pump laser case. The squeezing is ~30% below the shot noise level. The threshold of the degenerate down-conversion process for this structure is at 2g=γ 1+γx=62 GHz, therefore, with the above chosen pump beam parameters we are well below the threshold. The optimum squeezing degree for the structure at threshold is ~70%.

 figure: Fig. 4.

Fig. 4. a) Shift of the holes next to the cavity in order to increase its Q and also its coupling efficiency to the waveguide, b) the cavity is tilted with respect to the waveguide to boost their coupling efficiencies.

Download Full Size | PDF

 figure: Fig. 5.

Fig. 5. Intensity profile of a) X component and b) Y component of the electric field associated with the cavity in Fig.(4), and c) total intensity of the pump beam in the vicinity of the cavity. Center of the cavity is located at (x=0,y=0).

Download Full Size | PDF

 figure: Fig. 6.

Fig. 6. The spectrum of squeezing for the Y quadrature of the sample in Fig. (4) for a crystal oriented along the [111] direction pumped with 500 ps pulses of 80 MHz repetition rate and 10 mW average power.

Download Full Size | PDF

In summary, a 30% squeezing of the light coupled into a single-mode ridge waveguide at λ=1.44µm is predicted, when that waveguide is terminated in an L3 planar photonic crystal microcavity that is excited from the top half-space by a classical source at λ=0.72µm of 10 mW average power with 500 ps pulse duration and an 80 MHz repetition rate. The calculation is based on a degenerate down-conversion formalism for a cavity coupled to multiple continua, with the linear and nonlinear coupling parameters obtained using FDTD simulations of a realistic Al0.3Ga0.7As membrane structure. In order to maximize the coupling efficiency between the cavity and the single-mode channel, the cavity was tilted 60 degrees with respect to the waveguide axis on a [111] oriented AlGaAs crystal. The Q of the cavity without the waveguide was 80000. This amount of squeezing is within a factor of 2 of the maximum possible at the threshold of the process, given the coupling efficiency to the single channel output of 70%. The calculation was based on using a bulk AlGaAs alloy membrane with no attempt made to resonantly enhance χ (2) using quantum wells or quantum dot resonances.

Acknowledgments

This work was supported by the Natural Sciences and Engineering Research Council of Canada, the Canadian Institute for Advanced Research, the Canadian Foundation for Innovation, and the British Columbia Knowledge Development Fund.

References and links

1. A. J. Shields, “Semiconductor quantum light sources,” Nature Photon . 1, 215–223 (2007). [CrossRef]  

2. M. Hijlkema, B. Weber, H. P. Specht, S. C. Webster, A. Kuhn, and G. Rempe, “A single-photon server with just one atom,” Nature Phys. 3, 253–255 (2007). [CrossRef]  

3. M. Dusek, N. Lutkenhaus, and M. Hendrych, “Quantum Cryptography,” Progress in Optics, E. Wolf, ed., (Elsevier, 2006), Vol. 49.

4. E. Knill, L. Laflamme, and G. J. Milburn, “Efficient linear optics quantum computation,” Nature 409, 46–52 (2001). [CrossRef]   [PubMed]  

5. S. L. Braunstein and A. K. Pati, Quantum information with continuous variables (Kluwer Academic Publishers, 2003).

6. A. Furusawa, J. L. Sorensen, S. L. Braunstein, C. A. Fuchs, H. J. Kimble, and E. S. Polzik, “Unconditional quantum teleportation,” Science 282, 706–709 (1998). [CrossRef]   [PubMed]  

7. S. L. Braunstein and P. van Loock, “Quantum information with continuous variables,” Rev. Mod. Phys. 77513–577 (2005). [CrossRef]  

8. J. A. Gaj, et al., “Semiconductor heterostructures for spintronics and quantum information,” C. R. Physique 8, 243–252 (2007). [CrossRef]  

9. G. Burkard, “Spin qubits: Connect the dots,” Nature Phys. 2, 807–808 (2006). [CrossRef]  

10. M. W. McCutcheon, G. W. Rieger, I. W. Cheung, J. F. Young, D. Dalacu, S. Frederick, P. J. Poole, G. C. Aers, and R. L. Williams, “Resonant scattering and second-harmonic spectroscopy of planar,” photonic crystal microcavities Appl. Phys. Lett. 87, 221110 (2005).

11. J. P. Karr, A. Bass, R. Houdre, and E. Giacobino, “Squeezing in semiconductor microcavities in the strong-coupling regime,” Phys. Rev. A 69, 031802(R) (2004).

12. A. N. Vamivakas, B. E. A. Saleh, A. V. Sergienko, and M. C. Teich, “Theory of spontaneous parametric down-conversion from photonic crystals,” Phys. Rev. A 70, 043810 (2004). [CrossRef]  

13. G. Weihs, “Parametric down-conversion in photonic crystal waveguides,” Int. J. Mod. Phys. B 20, 1543–1550 (2006). [CrossRef]  

14. L. Xiao, Y. Wang, W. Zhang, Y. Huang, and J. Peng, “A 2-D photonic crystal based source of polarization entangled photon pairs with high nonlinear conversion efficiency and without walk-off compensation,” Opt. Commun. 272, 525–528 (2007). [CrossRef]  

15. C. Viviescas and G. Hackenbroich, “Field quantization for open optical cavities,” Phys. Rev. A 67, 013805 (2003). [CrossRef]  

16. H. J. Kimble, Fundamental Systems in Quantum Optics (Elsevier Science Publishing, 1992) Chap. 10.

17. H. Carmichael, An open systems approach to quantum optics (Springer-Verlag, 1993).

18. S. Reynaud, C. Fabre, and E. Giacobino, “Quantum fluctuations in a 2-mode parametric oscillator,” J. Opt. Soc. Am. B 4, 152–1524 (1987). [CrossRef]  

19. C. Fabre, E. Giacobino, A. Heidmann, L. Lugiato, S. Reynaud, M. Vadacchino, and Wang Kaige, “Squeezing in detuned degenerate optical parametric oscillators,” Quantum Opt. 2, 159–187 (1990). [CrossRef]  

20. A. Taflove and S. C. Hagness, Computational Electrodynamics: The Finite-Difference Time-Domain Method (Artech House Publishers, 2005).

21. M. G. Banaee, A. G. Pattantyus-Abraham, M. W. McCutcheon, G. W. Rieger, and Jeff F. Young, “Efficient coupling of photonic crystal microcavity modes to a ridge waveguide,” Appl. Phys. Lett. 90, 193106 (2007). [CrossRef]  

22. K. Srinivasan and O. Painter, “Momentum space design of high-Q photonic crystal optical cavities,” Opt. Express 10, 670–684 (2002). [PubMed]  

23. Y. Akahane, T. Asano, B. S. Song, and S. Noda, “High-Q photonic nanocavity in a two-dimensional photonic crystal,” Nature 425, 944–947 (2003). [CrossRef]   [PubMed]  

24. A. Faraon, E. Waks, D. Englund, I. Fushman, and J. Vuckovic, “Efficient photonic crystal cavity-waveguide couplers,” Appl. Phys. Lett. 90, 073102 (2007). [CrossRef]  

25. H. Feshbach, “A unified theory of nuclear reactions. II,” Ann. Phys. 19, 287–313 (1962). [CrossRef]  

26. S. Nordholm and S. A. Rice, “A quantum ergodic theory approach to unimolecular fragmentation,” J. Chem. Phys. 62, 157–168 (1975). [CrossRef]  

27. M. Shapiro and P. Brumer, “Quantum control of bound and continuum state dynamics,” Phys. Rep. 425, 195–264 (2006). [CrossRef]  

28. L. Mandel and E. Wolf, Optical coherence and quantum optics (Cambridge University Press, 1995).

29. M. Hillery and L. Mlodinow, “Quantized fields in a nonlinear dielectric medium: a microscopic approach,” Phys. Rev. A 55, 678–689 (1997). [CrossRef]  

30. M. Hillery, Quantum Squeezing (Springer-Verlag, 2004), edited by P. D. Drummond and Z. Ficek, chap. 2.

31. Z. Y. Ou, S. F. Pereira, and H. J. Kimble, “Realization of the Einstein-Podolsky-Rosen paradox for continuous variables in nondegenerate parametric amplification,” Appl. Phys. B 55, 265–278 (1992). [CrossRef]  

32. M. D. Reid and P.D. Drummond, “Quantum correlations of phase in nondegenerate parametric oscillation,” Phys. Rev. Lett , 60, 2731–2733 (1988). [CrossRef]   [PubMed]  

33. M. D. Reid and P.D. Drummond, “Correlations in nondegenerate parametric oscillation II, below threshold results,” Phys. Rev. A 41, 3930–3949 (1990). [CrossRef]   [PubMed]  

34. Y. Akahane, M. Mochizuki, T. Asano, Y. Tanaka, and S. Noda, “Design of a channel drop filter by using a donortype cavity with high-quality factor in a two-dimensional photonic crystal slab,” Appl. Phys. Lett. 82, 1341–1343 (2003). [CrossRef]  

35. S. Adachi, GaAs and Related Materials: Bulk Semiconducting and Superlattice Properties (World Scientific Publishing Company, 1994). [CrossRef]  

36. Y. Dumeige, I. Sagnes, P. Monnier, P. Vidakovic, I. Abram, C. Meriadec, and A. Levenson, “Phase-matched frequency doubling at photonic band edges: efficiency scaling as the fifth power of the length,” Phys. Rev. Lett. 89, 043901 (2002). [PubMed]  

37. T. Asano, Bong-Shik Song, Y. Akahane, and S. Noda, “Ultrahigh-Q Nanocavities in Two-Dimensional Photonic Crystal Slabs,” IEEE J. Sel. Top. Quantum Electron. 12, 1123–1134 (2006). [CrossRef]  

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 (6)

Fig. 1.
Fig. 1. 2D photonic crystal microcavities, a) an isolated cavity and b) adding a 1D waveguide channel to the cavity structure.
Fig. 2.
Fig. 2. The model cavity that communicates with several output channels.
Fig. 3.
Fig. 3. a) Spectrum at threshold of squeezing for the Y quadrature in a degenerate down-conversion process. b) Squeezing versus g factor at Ω=0. The solid lines are for the case when the cavity couples to a single channel (γx =0) and the dashed lines are when γx =γ 1.
Fig. 4.
Fig. 4. a) Shift of the holes next to the cavity in order to increase its Q and also its coupling efficiency to the waveguide, b) the cavity is tilted with respect to the waveguide to boost their coupling efficiencies.
Fig. 5.
Fig. 5. Intensity profile of a) X component and b) Y component of the electric field associated with the cavity in Fig.(4), and c) total intensity of the pump beam in the vicinity of the cavity. Center of the cavity is located at (x=0,y=0).
Fig. 6.
Fig. 6. The spectrum of squeezing for the Y quadrature of the sample in Fig. (4) for a crystal oriented along the [111] direction pumped with 500 ps pulses of 80 MHz repetition rate and 10 mW average power.

Equations (30)

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

f m ( r , ω ) = λ α λ m ( ω ) U λ ( r ) + n d ω β n m ( ω , ω ) V n ( r , ω ) .
H ̂ = λ h ̅ ω λ a ̂ λ a ̂ λ + m d ω h ̅ ω r ̂ m ( ω ) r ̂ m ( ω )
+ h ̅ λ m d ω [ W λ m ( ω ) a ̂ λ r ̂ m ( ω ) + W λ m * ( ω ) a ̂ λ r ̂ m ( ω )
+ T λ m ( ω ) a ̂ λ r ̂ m ( ω ) + T λ m * ( ω ) a ̂ λ r ̂ m ( ω ) ] ,
H ̂ I = 2 ε 0 3 d 3 r E ̂ ( r , t ) . χ ( 2 ) ( r ) : E ̂ ( r , t ) E ̂ ( r , t ) ,
E ̂ c ( r , t ) = i h ̅ ω 1 2 ε 0 [ a ̂ 1 ( t ) U 1 ( r ) a ̂ 1 ( t ) U 1 * ( r ) ] .
E p ( r , t ) = i A p ( t ) [ U p ( r ) e 2 i ω 0 t U p * ( r ) e + 2 i ω 0 t ] ,
H ̂ I = i h ̅ [ g a ̂ 1 ( t ) a ̂ 1 ( t ) e 2 i ω 0 t g * a ̂ 1 ( t ) a ̂ 1 ( t ) e 2 i ω 0 t ] ,
g = ω 1 3 A p ( t ) d 3 r [ U 1 . χ ( 2 ) : U 1 U p + U 1 . χ ( 2 ) : U p U 1 + U p . χ ( 2 ) : U 1 U 1 ] .
a ˜ 1 ( Ω ) = 2 g 2 γ 1 r ˜ 1 in ( Ω ) + 2 g m = 2 2 γ m r ˜ m in ( Ω ) [ Γ i ( Δ + Ω ) ] [ Γ + i ( Δ Ω ) ] 4 g 2
+ [ Γ + i ( Δ Ω ) ] [ 2 γ 1 r ˜ 1 in ( Ω ) + m = 2 2 γ m r ˜ m in ( Ω ) ] [ Γ i ( Δ + Ω ) ] [ Γ + i ( Δ Ω ) ] 4 g 2 ,
S X ( Ω ) = < X out ( Ω ) , X out ( Ω ) > 1 ,
S Y ( Ω ) = < Y out ( Ω ) , Y out ( Ω ) > 1 ,
X out ( Ω ) = r ˜ 1 out ( Ω ) + r ˜ 1 out ( Ω ) ,
Y out ( Ω ) = i [ r ˜ 1 out ( Ω ) r ˜ 1 out ( Ω ) ] .
r ˜ 1 out ( Ω ) r ˜ 1 in ( Ω ) = 2 γ 1 a ˜ 1 ( Ω ) ,
S X ( Ω ) = 8 g γ 1 ( γ 1 2 g ) 2 + Ω 2 ,
S Y ( Ω ) = 8 g γ 1 ( γ 1 + 2 g ) 2 + Ω 2 .
r ˜ m out ( Ω ) r ˜ m in ( Ω ) = 2 γ 1 a ˜ 1 ( Ω ) ,
S X ( Ω ) = 8 g γ 1 [ ( γ 1 + γ x ) 2 g ] 2 + Ω 2 ,
S Y ( Ω ) = 8 g γ 1 [ ( γ 1 + γ x ) + 2 g ] 2 + Ω 2 ,
γ x = m = 2 γ m ,
S Y ( Ω = 0 ) = γ 1 γ 1 + γ x ,
χ [ 001 ] ( 2 ) ( ω 1 ; 2 ω 0 , ω 1 ) = ( χ xxx ( 2 ) χ xxy ( 2 ) χ xxz ( 2 ) χ xyx ( 2 ) χ xyy ( 2 ) χ xyz ( 2 ) χ xzx ( 2 ) χ xzy ( 2 ) χ xzz ( 2 ) χ yxx ( 2 ) χ yxy ( 2 ) χ yxz ( 2 ) χ yyx ( 2 ) χ yyy ( 2 ) χ yyz ( 2 ) χ yzx ( 2 ) χ yzy ( 2 ) χ yzz ( 2 ) χ zxx ( 2 ) χ zxy ( 2 ) χ zxz ( 2 ) χ zyx ( 2 ) χ zyy ( 2 ) χ zyz ( 2 ) χ zzx ( 2 ) χ zzy ( 2 ) χ zzz ( 2 ) )
= β ( 0 0 0 0 0 1 0 1 0 0 0 1 0 0 0 1 0 0 0 1 0 1 0 0 0 0 0 ) ,
χ [ 111 ] ( 2 ) ( ω 1 ; 2 ω 0 , ω 1 ) = β ( 3 3 0 2 2 3 0 1 3 0 2 2 3 0 1 2 2 3 0 1 3 0 1 3 0 2 3 0 2 3 0 2 2 3 0 1 2 2 3 0 2 3 0 1 2 2 3 0 3 2 3 )
1 Q T = 1 Q i + 1 Q wg ,
γ 1 = ω 1 Q wg = 46 GHz , γ x = ω 1 Q i = 16 GHz .
g [ 111 ] = 0.839 GHz .
g [ 111 ] = 4.199 GHz .
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.