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

Vacuum induced transparency in metamaterials

Open Access Open Access

Abstract

For the cavity-based electromagnetically induced transparent (EIT), as the coherent driving field is enhanced by the optical cavity, the weak probe field can propagate through the atomic ensemble without absorption even if the driving field is weak. The extreme case of vacuum in the cavity is called “vacuum-induced transparency” (VIT) to distinguish it from the cavity EIT. Here we construct a new kind of cavity made of Metamaterials, i.e. ε-negative (EN) and μ-negative (MN) slabs, and study the VIT phenomena of the atomic ensemble doped within it. When the impedances of the MN and EN slabs are matched to each other and the dissipation of the material is small, it behaves as a surface plasmon cavity with a huge Q factor. And the VIT phenomenon in this cavity appears. By adjusting the position of atoms, the coupling strength between the atom and the structure could be changed. Two kinds of extremes of VIT, the coherent population trapping (CPT) and the Autler-Townes splitting (ATS), can be achieved in this system easily. Our proposal could be used in the realization of ultra-strong coupling and integrated devices on quantum memory or optical switch.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Electromagnetically induced transparency (EIT) has been widespread concerned because of its wonderful feature and promising applications [1], such as laser sideband cooling, light velocity controlling, quantum memory, and so on. Besides in the three-level atom systems, EITs have also been found in other systems such as four-wave-mixing schemes [2, 3], optomechanics [4–6], metamaterials [7, 8], plasma [9], and even vacuum [10]. It is well known that vacuum possesses numerous virtual photons due to vacuum fluctuation, which is the reason of Casimir force [11] and Lamb shift [12, 13]. More important, the vacuum fluctuation can be modified by artificial electromagnetic environments, and the vacuum-related quantum phenomena can also be controlled by these man-made structures.

High-finesse optical cavity has been well developed and widely used in QED systems [14–17]. It is also used in EIT experiment to strengthen the coupling between atoms and the driving field [18, 19]. In 2011, an extreme case of EIT was realized in the experiment, which is called as “Vacuum-Induced Transparency” (VIT) because the driving field is replaced by the vacuum cavity mode, there was no real photon initially in the cavity [20–22]. Although VIT has promising applications on photon switch or transistor [23], it is hard to be achieved in the conventional cavity and is difficult to be widely used up to other optical or quantum devices, because of the mediocre coupling coefficient between atoms and cavity.

In recent decades, the development of metamaterials provides new tools to control light or photons, even vacuum. Such materials can possess negative effective permittivity (ε-negative, EN), negative effective permeability (μ-negative, MN), or negative refractive index (left-hand materials, LHM) at interesting frequencies after subtle design. Until now metamaterials had been achieved from microwave [24] to visible light [25–29]. Due to their unique properties, metamaterial have promising applications on quantum optics, such as quantum interference enhancement [30–32], long-lived entanglement [33], perfect lens [34], superradiance [35], and Casimir force [36–38]. In 2003, a structure which pairing an EN slab and an MN slab was proposed by Alu and Engheta [39]. It was found that highly local of the fields appear near the interface, because such structure can support both the tunnel modes and surface plasmonic modes simultaneously. If a dipole is placed near the interface, the high local field provides a strong coupling between dipole and field. Therefore various harmonic systems, i.e. classical oscillator [40–42] and two-level atom [43], are considered to be sandwiched between MN and EN slabs, and the spectra of Rabi splitting were observed.

Taking advantage of the high locality of the surface mode near the interface between MN and EN slabs, in this work, we consider a thin layer of three-lever atoms doped very close to such interface. The combo structure made by the matched MN and EN slabs acts as a high Q-cavity which resonantly couples to one of the atomic dipole moments, and greatly alter the response of another atomic dipole to the weak probe field. Compared to the standing wave mode in the ordinary cavity, the surface modes in our scheme lead to the huge mode density and provide an extremely strong coupling to the atomic dipole. We focus on the case that the field in the combo structure is in the vacuum state, and discuss the phenomenon of VIT. As the atom-field coupling in our combo structure is determined by the location of atoms, various VIT phenomena can be observed if the atomic layer locates in different positions. For comparison, it is hard to observe such significant VIT phenomena in conventional micro-cavity because the atom-filed coupling coefficient in conventional micro-cavity is a few orders less than that in our scheme.

Our paper is organized as follows. In Sec. II, we introduce the cavity-based EIT model and extend it to VIT naturally. In Sec. III, We discussed the calculation of the cooperation parameter, which is the key parameter related to VIT, by Green tensor concerning the matching MN and EM slabs. In Sec. IV, the transmission spectra are calculated, and reveal the VIT phenomena within such metamaterials structure. The relative discussions and an application of single-photon switch are presented in Sec V. We draw the conclusion in Sec. VI.

2. Model and theory

We start from the usual EIT system, which is a Λ-type three-level atom interacting with a weak probe field and the coherent driving field simultaneously. The atomic scheme and relative parameters are shown in the inset of Fig. 1. By ignoring the dissipation, the Hamiltonian of the system is given

H=ωab|aa|+ωcb|cc|(Ωpeiωpt|ab|+Ωceiωct|ac|+H.c.),
where ωp, ωcand ωij are frequencies of the probe field, the coherent driving field, and the atomic transition |i|j, respectively. Ωp(Ωc) is Rabi frequency that represents the coupling between the probe (coherent driving) field and the atomic transition |a|b (|a|c).

 figure: Fig. 1

Fig. 1 A single layer of Λ-type three-level atoms located near the interface between a μ-negative slab and a ε-negative slab. The level scheme of an atom is shown in the dashed circle. The directions of polarization are indicated by double-headed arrows.

Download Full Size | PDF

The response of atom to probe field is embodied by the density matrix ρab. In the case of the weak probe field, it is reasonable to assume a vanishing atomic excitation. And then, in the steady state, one is able to deduce the analytical results of density matrix according to the master equation [1, 44]. If we extend the single atom to a bulk atomic gas with total number N and volume V, the linear susceptibility of the atomic gas can be expressed as

χ=|μab|2Nε0V4δ(|Ωc|24δΔ)4Δγbc2+i8δ2γab+i2γbc(|Ωc|2+γbcγab)||Ωc|2+γabγbc4Δδ+i2δγab+i2Δγbc|2.

Where μab is the electric dipole moment of atomic transition |a|b andN/V is the atom number density. HereΔ=ωabωp is the detuning between the transition |a|band the probe fields, Δc=ωacωc is the detuning between the transition |a|cand the coherent driving fields, while δ=ΔΔc=ωcb(ωpωc) is the two-photon detuning. The decoherence rates are defined as γab=Γab+Γac+γa and γbc=γc in which Γab (Γac) is the spontaneous emission rates from state |a to states |b (|c). γaand γc are the dephasing rates of states |a and |c, respectively.

The decoherence rateγbcis a key quantity of EIT. If γbc0, the susceptibility should have a nonzero imaginary value at resonanceΔ=Δc=δ=0, i.e. χiγbc/|Ωc|2when Ωc>>γbc,γab. As γbc cannot be eliminated exactly in experimentally, the EIT can only be improved by enhancing the Rabi frequency of the coherent driving field Ωc.

The Rabi frequency Ωc relates to both the coupling coefficient g (2g is the vacuum Rabi frequency) and the intensity of the driving field. For the weak driving filed which can be described by photon number nc, the Rabi frequency Ωccan be rewritten as

Ωc=2gnc+1

For a high g, Ωc could be significant even if the coherent driving field is absent, i.e.nc=0, and has the value of Ωc=2g. When 2g is larger than the decoherent rate γbc, the EIT phenomena are still apparent in absent of driving field. Such a phenomenon is named as vacuum induced transparent (VIT).

The effective way to enhance the coupling coefficient g is placing the atom into a cavity. Therefore several cavity-based VIT or vacuum Rabi splitting were analyzed, such as cavity-atoms system [10, 20], semiconductor quantum dots in photonic crystal [45], and semiconductor circuit [46].

For the usual cavity-based VIT scheme, the transition |a|c is resonant with the cavity mode in order to enhance g. Meanwhile, the transition |a|b, which corresponds to the probe field, is far detuning from the cavity mode. To describe the cavity-based VIT, a quantity named as the single-atom cooperation parameter is defined as

η=4g2/γabγbc

As mentioned before γaband γbc include several channels. However, they can be simplified in the QED system. Forγbc, its prominent element is the cavity losses κ at the cavity frequency or equivalentωac, i.e. γbcκ. Meanwhile, for γab, its dominant contribution comes from the decay rate from |ato|b, i.e.γabΓab. Thus the single-atom cooperativity parameter is approximated as [20, 23]

η=4g2/κΓab

Correspondingly, the susceptibility χ in Eq. (2) can be rewritten as the function of η, which is (for derivation, see Appendix) [13]

χ=|μab|2Nε0V[2δ˜(ηδ˜Δ˜)2Δ˜+i2(δ˜2+η+1)(η+1Δ˜δ˜)2+(Δ˜+δ˜)]1Γab

Here Δ˜=2Δ/Γab and δ˜=2δ/κare the normalized probe-atom detuning and the two-photon detuning, respectively [20].

In the conventional cavity-based VIT system, the single-atom cooperativity parameter can reach 7.2±0.5 theoretically, and achieve only about 3.4 in experiment [20], which is limited by the dipole momentum and standing-wave cavity mode.

Here we propose a new kind of cavity possessing larger coupling coefficient g but smaller cavity loss κ, so that it is helpful to improve the VIT phenomena. Such special cavity is a double-layer structure, which is combined with a μ-negative(MN) slab and a ε-negative(EN) slab, shown in Fig. 1. The interface between two single-negative slabs is defined as the x-y plane at z = 0.

The permeabilities and permittivities of the EN and the MN slabs can be reasonably described by Drude models as

ε1=2, μ1=1ωm2ω2+iγmω
ε2=1ωe2ω2+iγeω,μ2=2

Where ωe and ωm are the bulk plasma frequencies. γe and γm refer to the dissipations of the materials. In this paper, these parameters are set as ωe=ωm=1.732ω0 (ω0=3.0×1015Hz), γe=γm=1.67×108ω0=5×107s1. In addition, the thicknesses of slabs are set to be d1=d2=λ0=2πc/ω0.

Such two-layer structure exhibits the unique properties at the frequency ω0=3.0×1015Hz. According to ref [42], at such frequency ε1(ω0)=μ2(ω0)=2and μ1(ω0)=ε2(ω0)=2, the structure supports a tunneling mode and various surface plasmonic modes. For both kinds of modes, the electric fields are concentrated at the interface between EN and MN slabs, which acts as a high-Q cavity. This Q factor does not relate to the Q factor of each material, but from this special combination of EN and MN slabs. The tunneling mode responds to propagating wave out of the structure, and contributes to the cavity loss κ. Meanwhile, the surface plasmon modes can coherently interact with the atom, and provide a huge g when the atom is located at the interface.

To show the tunneling mode, we consider the propagating wave incident from left to the combo of EN and MN slabs without three-level atom, and plot the transmittance spectrum in Fig. 2.

 figure: Fig. 2

Fig. 2 The transmittance spectrum of the combo of EN and MN slabs for the normal incident. Relative parameters are presented in the context.

Download Full Size | PDF

From Fig. 2, the transmission spectrum shows a serial of transparent windows. Especially at ωω0=3×1015s1, the transparent peak is ultra narrow which refers to the tunneling mode. It is a perfect cavity mode with low loss and high Q-factor. In addition, around ω=8.2×1015s1 (the probe window we marked), as both the permittivity and the permeability of the slabs become positive and act as dielectrics, there is a wide transparent band.

To show the property of the surface plasmon modes at ω0, we choose three surface modes with propagating constant K||=K0,10K0,100K0 and plot their normalized electric field distributions in Fig. 3.

 figure: Fig. 3

Fig. 3 The normalized electric field intensity distribution of the surface mode in the combo of EN and MN slabs at ω=ω0=3×1015s1. Three modes with propagating constants K||=K0,10K0,100K0 are shown.

Download Full Size | PDF

From Fig. 3, the normalized electric field of three surface modes concentrates at the interface z = 0 and decays exponentially away from the interface. The FWHM (full width at half maximum) of the mode distribution decreases with the increasing of propagating constant K. As the surface modes (K>K0=ω0/c) can exist for arbitrary propagating constant K [43], the combo structure provides a huge coupling coefficient g when an atom with transition frequency ω0 is located at the interface. Furthermore, the coupling coefficient g would decrease exponentially when the atom is away from the interface. As a result, the combo of EN and MN provide a perfect platform to realize the controllable VIT.

3. Cooperativity parameter

Based on the unique property of the combo structure mentioned above, the atomic level scheme is chosen so that the transition |a|c is resonant with the surface mode and the tunneling mode, i.e. ωac=ω0=3.0×1015s1and its dipole momentum μac=6×1029Cm is along the z-axis direction. Therefore it achieves to high g and lower κ; Meanwhile, the transition |a|bpossesses the frequency at the center of the wider transparent window, i.e. ωab=8.2×1015s1and the dipole momentum μab=6×1030Cm in the x-axis. so the probe field can transmit through the combo and interact with the dipole μab. We consider a thin layer of three-level atoms doped at z0.

According to Eq. (5), the single-atom cooperativity parameter can be determined by the value of 4g2,Γab, and κ. As mentioned before, 4g2 is the coupling intensity between the vacuum mode and atomic transition |a|c. It can be calculated by the classical electromagnetic Green tensor G(z0,z0,ω) [32, 47–50], and has the expression as

4g2=2ωac2μac2c2ε0Reμ14π0K3K12K1z[1+2R12TMr10TMe2iK1zd1+R12TMe2iK1zz0+r10TMe2iK1z(d1z0)1R12TMr10TMe2iK1zd1]dK|ω=ωac

The integral over K in Eq. (9) is from 0 to infinite but mainly originate from the surface modes (the part meet that K>ω0/c). So all quantities are the function of frequency ω=ωac=ω0=3.0×1015s1.

Γabis the spontaneous decay rate of transition |a|b, and is mainly related to the electromagnetic modes at ωab=8.2×1015s1. Similar to Eq. (9), using Green tensor in a different polarization, it has the detail expressions as

Γab=2ωab2μab2c2ε0Reμ18π0KK1z[1+2R12TEr10TEe2iK1zd1+R12TEe2iK1zz0+r10TEe2iK1z(d1z0)1R12TMr10TMe2iK1zd1]+K1z2K12[1+2R12TMr10TMe2iK1zd1+R12TMe2iK1zz0+r10TMe2iK1z(d1z0)1R12TMr10TMe2iK1zd1]dK|ω=ωab

Where K1=ε1μ1ω/c is wave number in MN slab, K is the xy plane component of the wave vector, and K1z satisfyingK1z2=K12K2 is the z-component of the wavevector. rijTE and rijTM are the reflection coefficients at the interface between the ith and jth slab when the TE and TM polarized field incident from ith slab to jth slab, respectively (ij;i,j{0,1,2} is marked in Fig. 1). R12q (q{TE,TM}) are the reflection of the MN slab when the TE or TM polarized field incident from EN slab to MN slab, which is defined

R12q=r12q+r20qe2ik2zd21+r12qr20qe2ik2zd2

The cavity loss κ can be measured by the FWHM of the tunnel mode centered at ωac=ω0=3.0×1015s1. From Fig. 2, we get the FWHM of tunnel mode is 5×107s1. So in this paper, the cavity loss is κ=5×107s1and it is independent of the atom.

As 4g2 related to the surface mode directly, it is sensitive to the position of atom film. To achieve a proper cooperativity parameter, we plot 4g2, Γab and η as a function of the position of atom film in Fig. 4.

 figure: Fig. 4

Fig. 4 Cooperativity parameter η (black solid line), the coupling of control field 2g (blue dash line), and the decay rate of atomic transition|a|b Γab (red dot line) change with the position of atoms.

Download Full Size | PDF

For the decay rate of atomic transition|a|b, as the combo are dielectric at ωab=8.2×1015s1, Γab is nearly the constant for different positions. However, 2g increase exponentially when atoms tend to the interface z=0. For example, η is 10−10 for z0=λ0, η reaches 1 for z0 = 0.02λ0, and η is about 107 for z0 = 0.0001λ0. As a result, the VIT phenomena would be easily achieved in such structure theoretically.

4. Result

To discuss the input-output problem, a thin atomic layer is set near the surface at z0. The atomic areal density is N/σ, and then the atom layer possesses the resonant optical depth of the ensemble =|μab|2ωabN/2cε0Γabσ with a length L along the probe beam [20]. Thus the transmission amplitude of probe field is t=exp(iKabLχ/2), where the influence of the structure on the transmission is ignored because the frequency of probe field is set at the center of the transparent window. The transmission amplitude is only affected by the atomic susceptibility χ. χ is related directly to Eq. (6) and should be modified correspondingly for the atomic layer as follows (for derivation, see Appendix),

χ=4KabLΔ˜(ηΔ˜δ˜)δ˜i(η+1+δ˜2)(η+1+δ˜2)2+(Δ˜+δ˜)2

When both the probe field and the cavity mode are resonant with or very close to their respective atomic transitions, i.e. Δ=δ=0, as well as the atoms are close to the interface between EN slab and MN slabs enough, the phenomenon of VIT can be found for weak probe light. Recently, several experiments discussed the laser-cooled atoms optically trapped inside a high-finesse optical cavity, for example=0.4 or =0.5for a double-pass [20], and =0.9 in a later experiment [23]. Therefore, in our scheme, we assume a low areal density of the atomic layer as 2×1012m2 and the optical depth=0.25.

When the atom layer is in free space, the transmission spectrum shows a Lorentz dip near resonance and T(Δ=0)=0.37, shown in the red dashed curve in Fig. 6. However, when the atom layer is doped in the combo structure, the result changes. For example, when the position of atoms is z0=0.01λ0, the cooperative parameter reachesη=8.4, which is higher than the normal cavity [20, 23]. The susceptibility at resonance is χ(Δ=0)=0.43i which become nearly zero, shown in Fig. 5 and then the transmission probability (T=|t|2) at resonance changes to T(Δ=0)=0.9, shown in the blue solid curve in Fig. 6. The VIT phenomena happen. Because the dissipation of metamaterials brings a large κ=5×107s1, which also limits the coherence, the transparent window doesn’t get 1. The FWHM of the VIT peak is about 3.3×108s1, which is narrower than the vacuum Rabi frequency Ωc|nc=0=2g=1.4×109s1, and larger than the cavity loss κ.

 figure: Fig. 5

Fig. 5 The real part (blue solid line) and imaginary part (red dash line) of susceptibility of the atom layer in the combo structure. The atom layer is located at z0=0.01λ0 which leads to 2g=1.4×109s1 and η=8.4. Other parameters are Δc=0, Γab=4.7×109s1, κ=2π×8×106Hz, =0.25.

Download Full Size | PDF

 figure: Fig. 6

Fig. 6 The transmission spectrum of the atomic layer in the combo structure (Blue solid curve). The atom layer is located at z0=0.01λ0 which leads to 2g=1.4×109s1 and η=8.4. Other parameters are Δc=0, Γab=4.7×109s1, κ=2π×8×106Hz, =0.25. For comparison, the transmission spectrum of the atom layer in free space is also plotted in the red dashed curve.

Download Full Size | PDF

In a real device, if the atom areal density is improved to 5×1012m2, which means there is one atom in each circle with a radius of a half wavelength, the transmittance of atom layer in free space is less than 0.1, but the transmittance of atom layer in our combo structure is still around 0.8.

5. Discussion

One of the advantages of our structure is the coupling between atomic transition and the cavity mode could be controlled by the position of atoms. With the controllable coupling, different phenomena could be found.

For the usual EIT system, when the coherent driving field is very weak, a sharp transmission window possesses a tiny linewidth. In this case, the sharp transmission peak is based on quantum interference primarily. If we further decrease the power of the driving field, the FWHM of this narrow peak will decrease and stop at a value which equals to the decoherence rate γbcκ. This quantum interference phenomenon is called coherent population trapping (CPT) [51] and be used to measure the dephasing time experimentally [52]. In our VIT case, if the atoms are located at z0=0.033λ0 the cooperative parameter η is about 0.31, which is smaller than 1, the interaction is in the weak-coupling region. The transmission spectrum is plotted in Fig. 7. A small peak can be found in the bottom of the transmission dip, and the FWHM of the peak is equal to the cavity loss κ. So Fig. 7 is a CPT dominating case in our VIT system [53].

 figure: Fig. 7

Fig. 7 The transmission spectrum of the atom layer in the combo structure. The atom layer is located at z0=0.033λ0, which leads to2g=2.3×108s1and η=0.24. Other parameters are the same with those in Fig. 5. It shows the phenomena of Coherent population trapping. The FWHM of the small peak is 2π×8×106Hz, which is almost equal to the cavity loss.

Download Full Size | PDF

Different with CPT, in which the peak width is controlled by nature of the quantum interference [54], there is another phenomenon called Autler-Townes splitting (ATS) [55–58], which shows two distinguished resonances in the spectrum. And in some situations, there is a crossover from EIT to ATS [59, 60]. As a special case of EIT, VIT also combines two different physical processes, CPT which only modifies the atom states, and ATS which is dominated by an external field [61].

To reach ATS region, the coupling 2g should much larger than the decoherence γbcκ. We set the atom extremely close to the interface of two slabs, z0=0.001λ0, and the cooperative parameter η=8.3×103, which is never possible in a normal cavity. The transmission spectrum splits into two independent Lorenz dips and separated by Ωc|nc=0=2g=2π×7.07×109Hz (vacuum Rabi frequency), see Fig. 8. This is the signatures of Autler-Townes splitting. In this case, the cavity loss κ is much less than the coherent coupling 2g, so the dressed states dominate the ATS phenomenon [62]. As the density of atoms is low, the lowest value of the transmission dips is far from 0. If we further increase the density of atoms, the transmission dips should close to 0.

 figure: Fig. 8

Fig. 8 The transmission spectrum of the atom layer in the combo structure. The atom layer is located at z0=0.001λ0, which leads toΩc=2g=2π×7.0×109Hzand η=8.3×103. Other parameters are the same as those in Fig. 5. Two discrete dips demonstrate the Autler-Townes splitting, and the splitting value is equal to Ωc=2g.

Download Full Size | PDF

In addition, the ATS phenomenon in our structure can be applied to the single photon switcher. As the combo structure with matching MN and EN slabs is transparency not only for the probe field near ωab but also for the tunnel mode at ωac, the coupling Ωc can be improved by the photons through tunnel mode. Therefore one or a few photons in the tunneling mode can block the transmission of the probe field [63–66].

According to Ωc=2gnc+1, by inputting one photon at ωac through tunnel mode (nc=1), the splitting of two transmission dips can change from 2g to 22g when 2g is much larger than κ. We increase the areal density of atom into 8×1012m2, so that =1, and keep other parameters unchanged. The transmission spectrum is plotted in Fig. 9. The blue solid curve refers to the case of nc=1, while the red dashed curve refers to the case of nc=0. It is clear that the splitting of two dips for nc=1 is enhanced to 2π×9.9GHz, which is larger than 2π×7.0GHz for nc=0. If we set the frequency of the probe field at Δ=±2π×3.5GHz, the transmission will change from 0.02 to 0.96 for the case of nc=1. It behaves as a single photon switch. When the number of photon in the tunnel mode increases (nc>1), the splitting of two transmission dips will also increase with the Rabi frequency Ωc, and then the signal to noise ratio of the switch will be improved.

 figure: Fig. 9

Fig. 9 The transmission spectrum of the atom layer in the combo structure. The blue solid curve refers to the case of nc=1, while the red dashed curve refers to the case of nc=0. The areal density of atom is N/σ=8×1012m2, so that=1. The atomic layer is located at z0=0.001λ0.

Download Full Size | PDF

The loss of metamaterials is the main problem with such kinds of structure, because most of metamaterials use metal structure to achieve negative permeability and permittivity. But in recent years, the all-dielectric metamaterials base on Mie resonances give a wonderful solution for low loss [67, 68]. At the wavelength of the infrared quantum dot (1530nm), perfect reflectors with a 99.7% reflectance have been made from a hexagonal lattice of silicon cylinder resonators as the single-negative metamaterials [69]. And a simulation work shows almost zero absorption reflector at visible wavelengths by a single-negative metasurface made from an array of TiO2 nanoparticle arranged in square lattice embedded in air [70]. We notice that the losses in TiO2 structures could be as low as 10-6 level (the imaginary part of the refractive index) [71], which is close to what we use here. In addition, the result of VIT with different losses when the atom is located at z0=0.0001λ0 is shown in Fig. 10. The splitting is still distinguishable (green curve) when the dissipative isγe,m=1.67×105ω0 which is as large as real experiments [72]. The single-atom cooperativity parameters can reach 830 (two orders better than the ordinary cavity) at the dissipative γe,m=1.67×106ω0.

 figure: Fig. 10

Fig. 10 VIT transmission spectrum under different losses (z0=0.0001λ0). For γe,m=1.67×104.5ω0, cavity loss κ is larger than the decay rate of the atom Γ, so what we can see is only a transmission dip around Δ=0. The transmission at Δ=0 increases when γe,m=1.67×105ω0. As the dissipations decrease to 1.67×105.5ω0 or less, the splitting becomes significant and VIT phenomena can be distinguished clearly.

Download Full Size | PDF

For the location of the three-level system, it depends on the kinds of system. For quantum dot [16, 73] or quantum well, their position could be controlled by the time of chemical vapor deposition. In meta-atoms and radio frequency metamaterials systems [74], it is much easier since the wavelength is on the scale of a few millimeters. So the meta-atoms made by antennas could be moved by translation stage or by hand directly. In the visible region, sub-wavelength structures in transmission mode have been fabricated by GaN-based integrated-resonant unit elements [75]. In the future, as microfabrication technology advances, the single-negative metamaterial structure with the novel properties and such low loss may be produced at the infrared and visible range.

6. Conclusion

The EIT properties of a three-lever atom located near the interface of pairing single-negative metamaterial slabs are studied. Due to the particular distribution of electromagnetic field mode in the structure, it could be looked as a perfect single mode cavity, and it is much easier to achieve strong coupling than the traditional cavity. Because of the expression of Rabi frequency, Ωc=2gnc+1, the transparency of a three-lever atomic layer can be induced by the vacuum mode when the coupling is strong enough. Moreover, by adjusting the position of the atomic layer, we can control the coupling between atoms and the vacuum mode, and that’s why CPT and ATS could be observed. The coupling can be improved by move the atomic layer closer to the interface, and it has little impact on the non-resonant channels. Therefore, we give a demonstration of the optical switch controlled by a single photon. Furthermore, ultrastrong coupling [76, 77] could be achieved in this system easily and it is a possible way to reach deep strong coupling [78, 79].

With the development of microfabrication, the precise location of atom or meta-atom in artificial structures is possible in the experiment, and various kinds of low loss metamaterials have been come up or made [67, 80–83]. Besides, single-negative metamaterials in the visible frequency have been designed in recent years [25, 28, 29, 70]. In addition, this model is also applicable to the microwave frequencies, which is easy to implement experimentally [74]. Based on these technological advances and application needs, this VIT structure has promising applications on the integrated devices in quantum memory and optical switch.

Funding

National Natural Science Foundation of China (11574229, 1474221, 11504272, 11474003, 11747080); Science and Technology Commission of Shanghai Municipality (18JC1410900, 15XD1503700); the Shanghai Education Commission Foundation; National Key Basic Research Special Foundation of China (2013CB632701).

Acknowledgments

We thank Prof. Xian-Shan Huang for valuable discussion.

References and links

1. M. Fleischhauer, A. Imamoglu, and J. P. Marangos, “Electromagnetically induced transparency: Optics in coherent media,” Rev. Mod. Phys. 77(2), 633–673 (2005). [CrossRef]  

2. J. C. Petch, C. H. Keitel, P. L. Knight, and J. P. Marangos, “Role of electromagnetically induced transparency in resonant four-wave-mixing schemes,” Phys. Rev. A 53(1), 543–561 (1996). [CrossRef]   [PubMed]  

3. Y. Q. Li and M. Xiao, “Enhancement of nondegenerate four-wave mixing based on electromagnetically induced transparency in rubidium atoms,” Opt. Lett. 21(14), 1064–1066 (1996). [CrossRef]   [PubMed]  

4. A. H. Safavi-Naeini, T. P. Mayer Alegre, J. Chan, M. Eichenfield, M. Winger, Q. Lin, J. T. Hill, D. E. Chang, and O. Painter, “Electromagnetically induced transparency and slow light with optomechanics,” Nature 472(7341), 69–73 (2011). [CrossRef]   [PubMed]  

5. H. Wang, X. Gu, Y. X. Liu, A. Miranowicz, and F. Nori, “Optomechanical analog of two-color electromagnetically induced transparency: Photon transmission through an optomechanical device with a two-level system,” Phys. Rev. A 90(2), 023817 (2014). [CrossRef]  

6. Y.-P. Gao, C. Cao, T.-J. Wang, Y. Zhang, and C. Wang, “Cavity-mediated coupling of phonons and magnons,” Phys. Rev. A 96(2), 023826 (2017). [CrossRef]  

7. T. Nakanishi and M. Kitano, “Implementation of electromagnetically induced transparency in a metamaterial controlled with auxiliary waves,” Phys. Rev. Appl. 4(2), 024013 (2015). [CrossRef]  

8. K. L. Tsakmakidis, M. S. Wartak, J. J. H. Cook, J. M. Hamm, and O. Hess, “Negative-permeability electromagnetically induced transparent and magnetically active metamaterials,” Phys. Rev. B 81(19), 195128 (2010). [CrossRef]  

9. S. E. Harris, “Electromagnetically induced transparency in an ideal plasma,” Phys. Rev. Lett. 77(27), 5357–5360 (1996). [CrossRef]   [PubMed]  

10. J. E. Field, “Vacuum-Rabi-splitting-induced transparency,” Phys. Rev. A 47(6), 5064–5067 (1993). [CrossRef]   [PubMed]  

11. H. B. G. Casimir and D. Polder, “The influence of retardation on the London-van der Waals forces,” Phys. Rev. 73(4), 360–372 (1948). [CrossRef]  

12. M. O. Scully, “Collective lamb shift in single photon Dicke superradiance,” Phys. Rev. Lett. 102(14), 143601 (2009). [CrossRef]   [PubMed]  

13. W. E. Lamb and R. C. Retherford, “Fine structure of the hydrogen atom by a microwave method,” Phys. Rev. 72(3), 241–243 (1947). [CrossRef]  

14. J. Ye, D. W. Vernooy, and H. J. Kimble, “Trapping of single atoms in cavity QED,” Phys. Rev. Lett. 83(24), 4987–4990 (1999). [CrossRef]  

15. S. B. Zheng and G. C. Guo, “Efficient scheme for two-atom entanglement and quantum information processing in cavity QED,” Phys. Rev. Lett. 85(11), 2392–2395 (2000). [CrossRef]   [PubMed]  

16. A. Imamoglu, D. D. Awschalom, G. Burkard, D. P. DiVincenzo, D. Loss, M. Sherwin, and A. Small, “Quantum information processing using quantum dot spins and cavity QED,” Phys. Rev. Lett. 83(20), 4204–4207 (1999). [CrossRef]  

17. Q.-T. Cao, H. Wang, C.-H. Dong, H. Jing, R.-S. Liu, X. Chen, L. Ge, Q. Gong, and Y.-F. Xiao, “Experimental demonstration of spontaneous chirality in a nonlinear microresonator,” Phys. Rev. Lett. 118(3), 033901 (2017). [CrossRef]   [PubMed]  

18. J. Sheng, Y. Chao, S. Kumar, H. Fan, J. Sedlacek, and J. P. Shaffer, “Intracavity Rydberg-atom electromagnetically induced transparency using a high-finesse optical cavity,” Phys. Rev. A 96(3), 033813 (2017). [CrossRef]  

19. M. Mücke, E. Figueroa, J. Bochmann, C. Hahn, K. Murr, S. Ritter, C. J. Villas-Boas, and G. Rempe, “Electromagnetically induced transparency with single atoms in a cavity,” Nature 465(7299), 755–758 (2010). [CrossRef]   [PubMed]  

20. H. Tanji-Suzuki, W. Chen, R. Landig, J. Simon, and V. Vuletić, “Vacuum-induced transparency,” Science 333(6047), 1266–1269 (2011). [CrossRef]   [PubMed]  

21. M. Fleischhauer, “Physics. Switching light by vacuum,” Science 333(6047), 1228–1229 (2011). [CrossRef]   [PubMed]  

22. Y. H. Lien, G. Barontini, M. Scheucher, M. Mergenthaler, J. Goldwin, and E. A. Hinds, “Observing coherence effects in an overdamped quantum system,” Nat. Commun. 7, 13933 (2016). [CrossRef]   [PubMed]  

23. W. Chen, K. M. Beck, R. Bücker, M. Gullans, M. D. Lukin, H. Tanji-Suzuki, and V. Vuletić, “All-optical switch and transistor gated by one stored photon,” Science 341(6147), 768–770 (2013). [CrossRef]   [PubMed]  

24. A. R. Katko, S. Gu, J. P. Barrett, B. I. Popa, G. Shvets, and S. A. Cummer, “Phase conjugation and negative refraction using nonlinear active metamaterials,” Phys. Rev. Lett. 105(12), 123905 (2010). [CrossRef]   [PubMed]  

25. A. Ishikawa, T. Tanaka, and S. Kawata, “Negative magnetic permeability in the visible light region,” Phys. Rev. Lett. 95(23), 237401 (2005). [CrossRef]   [PubMed]  

26. A. D. Falco, M. Ploschner, and T. F. Krauss, “Flexible metamaterials at visible wavelengths,” New J. Phys. 12(11), 113006 (2010). [CrossRef]  

27. M. I. Aslam and D. Ö. Güney, “Surface plasmon driven scalable low-loss negative-index metamaterial in the visible spectrum,” Phys. Rev. B 84(19), 195465 (2011). [CrossRef]  

28. A. Sgrò, D. De Carlo, G. Angiulli, F. C. Morabito, and M. Versaci, “Accurate computation of Drude-Lorentz model coefficients of single negative magnetic metamaterials using a micro-genetic algorithm approach,” in Multidisciplinary Approaches to Neural Computing (Springer, 2018), pp. 47–55.

29. A. R. Azeez, T. A. Elwi, and Z. A. A. AL-Hussain, “Design and analysis of a novel concentric rings based crossed lines single negative metamaterial structure,” Eng. Sci. Technol. 20, 1140–1146 (2017).

30. Y. Yang, J. Xu, H. Chen, and S. Zhu, “Quantum interference enhancement with left-handed materials,” Phys. Rev. Lett. 100(4), 043601 (2008). [CrossRef]   [PubMed]  

31. G. Song, J. P. Xu, and Y. P. Yang, “Quantum interference between Zeeman levels near structures made of left-handed materials and matched zero-index metamaterials,” Phys. Rev. A 89(5), 053830 (2014). [CrossRef]  

32. J. Xu and Y. Yang, “Quantum interference of V-type three-level atom in structures made of left-handed materials and mirrors,” Phys. Rev. A 81(1), 013816 (2010). [CrossRef]  

33. Y. Yang, J. Xu, H. Chen, and S.-Y. Zhu, “Long-lived entanglement between two distant atoms via left-handed materials,” Phys. Rev. A 82(3), 030304 (2010). [CrossRef]  

34. J. B. Pendry, “Negative refraction makes a perfect lens,” Phys. Rev. Lett. 85(18), 3966–3969 (2000). [CrossRef]   [PubMed]  

35. J. Kästel and M. Fleischhauer, “Suppression of spontaneous emission and superradiance over macroscopic distances in media with negative refraction,” Phys. Rev. A 71(1), 011804 (2005). [CrossRef]  

36. Y. Yang, R. Zeng, J. Xu, and S. Liu, “Casimir force between left-handed-material slabs,” Phys. Rev. A 77(1), 015803 (2008). [CrossRef]  

37. Y. Yang, R. Zeng, H. Chen, S. Zhu, and M. S. Zubairy, “Controlling the Casimir force via the electromagnetic properties of materials,” Phys. Rev. A 81(2), 022114 (2010). [CrossRef]  

38. R. Zeng and Y. Yang, “Tunable polarity of the Casimir force based on saturated ferrites,” Phys. Rev. A 83(1), 012517 (2011). [CrossRef]  

39. A. Alu and N. Engheta, “Pairing an epsilon-negative slab with a mu-negative slab: Resonance, tunneling and transparency,” IEEE Trans. Antenn. Propag. 51(10), 2558–2571 (2003). [CrossRef]  

40. L. Zhang, Y. Zhang, Y. Yang, H. Li, H. Chen, and S. Zhu, “Experimental observation of Rabi splitting in effective near-zero-index media in the microwave regime,” Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 78(3), 035601 (2008). [CrossRef]   [PubMed]  

41. L. Zhang, Y. Zhang, Y. Yang, and H. Chen, “Experimental study of Rabi-type oscillation induced by tunneling modes in effective near-zero-index metamaterials,” Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 83(4), 046604 (2011). [CrossRef]   [PubMed]  

42. H. T. Jiang, H. Chen, and S. Y. Zhu, “Rabi splitting with excitons in effective (near) zero-index media,” Opt. Lett. 32(14), 1980–1982 (2007). [CrossRef]   [PubMed]  

43. X. Zeng, G. Li, Y. Yang, and S. Zhu, “Enhancement of the vacuum Rabi oscillation via surface plasma modes in single-negative metamaterials,” Phys. Rev. A 86(3), 033819 (2012). [CrossRef]  

44. C. Cohen-Tannoudji, J. Dupont-Roc, and G. Grynberg, Atom-Photon Interactions: Basic Processes and Applications (Wiley, 1998).

45. G. Khitrova, H. M. Gibbs, M. Kira, S. W. Koch, and A. Scherer, “Vacuum Rabi splitting in semiconductors,” Nat. Phys. 2(2), 81–90 (2006). [CrossRef]  

46. H. Toida, T. Nakajima, and S. Komiyama, “Vacuum Rabi splitting in a semiconductor circuit QED system,” Phys. Rev. Lett. 110(6), 066802 (2013). [CrossRef]   [PubMed]  

47. S. Scheel, L. Knöll, and D. G. Welsch, “Spontaneous decay of an excited atom in an absorbing dielectric,” Phys. Rev. A 60(5), 4094–4104 (1999). [CrossRef]  

48. H. T. Dung, S. Y. Buhmann, L. Knöll, D.-G. Welsch, S. Scheel, and J. Kästel, “Electromagnetic-field quantization and spontaneous decay in left-handed media,” Phys. Rev. A 68(4), 043816 (2003). [CrossRef]  

49. J.-P. Xu, Y.-P. Yang, Q. Lin, and S.-Y. Zhu, “Spontaneous decay of a two-level atom near the left-handed slab,” Phys. Rev. A 79(4), 043812 (2009). [CrossRef]  

50. M. S. Tomaš, “Green function for multilayers: Light scattering in planar cavities,” Phys. Rev. A 51(3), 2545–2559 (1995). [CrossRef]   [PubMed]  

51. R. Wynands and A. Nagel, “Precision spectroscopy with coherent dark states,” Appl. Phys. B 68(1), 1–25 (1999). [CrossRef]  

52. L. J. Rogers, K. D. Jahnke, M. H. Metsch, A. Sipahigil, J. M. Binder, T. Teraji, H. Sumiya, J. Isoya, M. D. Lukin, P. Hemmer, and F. Jelezko, “All-optical initialization, readout, and coherent preparation of single silicon-vacancy spins in diamond,” Phys. Rev. Lett. 113(26), 263602 (2014). [CrossRef]   [PubMed]  

53. H. Y. Ling, Y.-Q. Li, and M. Xiao, “Coherent population trapping and electromagnetically induced transparency in multi-Zeeman-sublevel atoms,” Phys. Rev. A 53(2), 1014–1026 (1996). [CrossRef]   [PubMed]  

54. G. S. Agarwal, “Nature of the quantum interference in electromagnetic-field-induced control of absorption,” Phys. Rev. A 55(3), 2467–2470 (1997). [CrossRef]  

55. S. H. Autler and C. H. Townes, “Stark effect in rapidly varying fields,” Phys. Rev. 100(2), 703–722 (1955). [CrossRef]  

56. L. Zhang, X. Feng, G. Fu, X. Li, L. Han, N. B. Manson, and C. Wei, “Coherent transient in dressed-state and transient spectra of Autler-Townes doublet,” Phys. Rev. A 70(6), 063404 (2004). [CrossRef]  

57. P. M. Anisimov, J. P. Dowling, and B. C. Sanders, “Objectively discerning Autler-Townes splitting from electromagnetically induced transparency,” Phys. Rev. Lett. 107(16), 163604 (2011). [CrossRef]   [PubMed]  

58. L. Giner, L. Veissier, B. Sparkes, A. S. Sheremet, A. Nicolas, O. S. Mishina, M. Scherman, S. Burks, I. Shomroni, D. V. Kupriyanov, P. K. Lam, E. Giacobino, and J. Laurat, “Experimental investigation of the transition between Autler-Townes splitting and electromagnetically-induced-transparency models,” Phys. Rev. A 87(1), 013823 (2013). [CrossRef]  

59. C. Zhu, C. Tan, and G. Huang, “Crossover from electromagnetically induced transparency to Autler-Townes splitting in open V-type molecular systems,” Phys. Rev. A 87(4), 043813 (2013). [CrossRef]  

60. L. Y. He, T. J. Wang, Y. P. Gao, C. Cao, and C. Wang, “Discerning electromagnetically induced transparency from Autler-Townes splitting in plasmonic waveguide and coupled resonators system,” Opt. Express 23(18), 23817–23826 (2015). [CrossRef]   [PubMed]  

61. B. Peng, S. K. Özdemir, W. Chen, F. Nori, and L. Yang, “What is and what is not electromagnetically induced transparency in whispering-gallery microcavities,” Nat. Commun. 5(1), 5082 (2014). [CrossRef]   [PubMed]  

62. S.-Y. Zhu and M. O. Scully, “Quantum interference effects in the Autler-Townes spontaneous spectrum,” Phys. Lett. A 201(1), 85–90 (1995). [CrossRef]  

63. 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(7047), 87–90 (2005). [CrossRef]   [PubMed]  

64. A. Faraon, I. Fushman, D. Englund, N. Stoltz, P. Petroff, and J. Vuckovic, “Coherent generation of non-classical light on a chip via photon-induced tunnelling and blockade,” Nat. Phys. 4(11), 859–863 (2008). [CrossRef]  

65. A. Ridolfo, M. Leib, S. Savasta, and M. J. Hartmann, “Photon blockade in the ultrastrong coupling regime,” Phys. Rev. Lett. 109(19), 193602 (2012). [CrossRef]   [PubMed]  

66. C. J. Zhu, Y. P. Yang, and G. S. Agarwal, “Collective multiphoton blockade in cavity quantum electrodynamics,” Phys. Rev. A 95(6), 063842 (2017). [CrossRef]  

67. S. Jahani and Z. Jacob, “All-dielectric metamaterials,” Nat. Nanotechnol. 11(1), 23–36 (2016). [CrossRef]   [PubMed]  

68. V. Kuzmiak, P. Markos, T. Szoplik, A. Krasnok, S. Makarov, M. Petrov, R. Savelev, P. Belov, and Y. Kivshar, “Towards all-dielectric metamaterials and nanophotonics,” Proc. SPIE 9502, 950203 (2015). [CrossRef]  

69. P. Moitra, B. A. Slovick, W. li, I. I. Kravchencko, D. P. Briggs, S. Krishnamurthy, and J. Valentine, “Large-scale all-dielectric metamaterial perfect reflectors,” ACS Photonics 2(6), 692–698 (2015). [CrossRef]  

70. Y. Huang, H. Xu, Y. Lu, and Y. Chen, “All-dielectric metasurface for achieving perfect reflection at visible wavelengths,” J. Phys. Chem. C 122(5), 2990–2996 (2018). [CrossRef]  

71. J.-E. Broquin, G. Nunzi Conti, C. Wächter, R. Rizzo, F. Michelotti, P. Munzert, and N. Danz, “Leaky waveguides for low ҡ-measurement: From structure design to loss evaluation,” Proc. SPIE 9750, 975019 (2016). [CrossRef]  

72. J. Dimmock, “Losses in left-handed materials,” Opt. Express 11(19), 2397–2402 (2003). [CrossRef]   [PubMed]  

73. A. Kiraz, M. Atatüre, and A. Imamoğlu, “Quantum-dot single-photon sources: Prospects for applications in linear optics quantum-information processing,” Phys. Rev. A 69(3), 032305 (2004). [CrossRef]  

74. Y. Sun, Y. Yang, H. Chen, and S. Zhu, “Dephasing-induced control of interference nature in three-level electromagnetically induced tansparency systems,” Sci. Rep. 5(1), 16370 (2015). [CrossRef]   [PubMed]  

75. S. Wang, P. C. Wu, V. C. Su, Y. C. Lai, M. K. Chen, H. Y. Kuo, B. H. Chen, Y. H. Chen, T. T. Huang, J. H. Wang, R. M. Lin, C. H. Kuan, T. Li, Z. Wang, S. Zhu, and D. P. Tsai, “A broadband achromatic metalens in the visible,” Nat. Nanotechnol. 13(3), 227–232 (2018). [CrossRef]   [PubMed]  

76. F. Beaudoin, J. M. Gambetta, and A. Blais, “Dissipation and ultrastrong coupling in circuit QED,” Phys. Rev. A 84(4), 043832 (2011). [CrossRef]  

77. G. Scalari, C. Maissen, D. Turcinková, D. Hagenmüller, S. De Liberato, C. Ciuti, C. Reichl, D. Schuh, W. Wegscheider, M. Beck, and J. Faist, “Ultrastrong coupling of the cyclotron transition of a 2D electron gas to a THz metamaterial,” Science 335(6074), 1323–1326 (2012). [CrossRef]   [PubMed]  

78. J. Casanova, G. Romero, I. Lizuain, J. J. García-Ripoll, and E. Solano, “Deep strong coupling regime of the Jaynes-Cummings model,” 6Phys. Rev. Lett. 105(26), 263603 (2010). [CrossRef]   [PubMed]  

79. S. De Liberato, “Light-matter decoupling in the deep strong coupling regime: the breakdown of the Purcell effect,” Phys. Rev. Lett. 112(1), 016401 (2014). [CrossRef]   [PubMed]  

80. N. Kaina, F. Lemoult, M. Fink, and G. Lerosey, “Negative refractive index and acoustic superlens from multiple scattering in single negative metamaterials,” Nature 525(7567), 77–81 (2015). [CrossRef]   [PubMed]  

81. C. Kurter, T. Lan, L. Sarytchev, and S. M. Anlage, “Tunable negative permeability in a three-dimensional superconducting metamaterial,” Phys. Rev. Appl. 3(5), 054010 (2015). [CrossRef]  

82. M. Sadatgol, Ş. K. Özdemir, L. Yang, and D. Ö. Güney, “Plasmon injection to compensate and control losses in negative index metamaterials,” Phys. Rev. Lett. 115(3), 035502 (2015). [CrossRef]   [PubMed]  

83. I. Al-Naib and W. Withayachumnankul, “Editorial introduction to the special issue: Terahertz metamaterials and photonic crystals,” J. Infrared Millim. Terahertz Waves 38(9), 1031–1033 (2017). [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 (10)

Fig. 1
Fig. 1 A single layer of Λ -type three-level atoms located near the interface between a μ -negative slab and a ε -negative slab. The level scheme of an atom is shown in the dashed circle. The directions of polarization are indicated by double-headed arrows.
Fig. 2
Fig. 2 The transmittance spectrum of the combo of EN and MN slabs for the normal incident. Relative parameters are presented in the context.
Fig. 3
Fig. 3 The normalized electric field intensity distribution of the surface mode in the combo of EN and MN slabs at ω = ω 0 = 3 × 10 15 s 1 . Three modes with propagating constants K | | = K 0 , 10 K 0 , 100 K 0 are shown.
Fig. 4
Fig. 4 Cooperativity parameter η (black solid line), the coupling of control field 2g (blue dash line), and the decay rate of atomic transition | a | b Γ a b (red dot line) change with the position of atoms.
Fig. 5
Fig. 5 The real part (blue solid line) and imaginary part (red dash line) of susceptibility of the atom layer in the combo structure. The atom layer is located at z 0 = 0.01 λ 0 which leads to 2 g = 1.4 × 10 9 s 1 and η = 8.4 . Other parameters are Δ c = 0 , Γ a b = 4.7 × 10 9 s 1 , κ = 2 π × 8 × 10 6 H z , = 0.25 .
Fig. 6
Fig. 6 The transmission spectrum of the atomic layer in the combo structure (Blue solid curve). The atom layer is located at z 0 = 0.01 λ 0 which leads to 2 g = 1.4 × 10 9 s 1 and η = 8.4 . Other parameters are Δ c = 0 , Γ a b = 4.7 × 10 9 s 1 , κ = 2 π × 8 × 10 6 H z , = 0.25 . For comparison, the transmission spectrum of the atom layer in free space is also plotted in the red dashed curve.
Fig. 7
Fig. 7 The transmission spectrum of the atom layer in the combo structure. The atom layer is located at z 0 = 0.033 λ 0 , which leads to 2 g = 2.3 × 10 8 s 1 and η = 0.24 . Other parameters are the same with those in Fig. 5. It shows the phenomena of Coherent population trapping. The FWHM of the small peak is 2 π × 8 × 10 6 H z , which is almost equal to the cavity loss.
Fig. 8
Fig. 8 The transmission spectrum of the atom layer in the combo structure. The atom layer is located at z 0 = 0.001 λ 0 , which leads to Ω c = 2 g = 2 π × 7.0 × 10 9 H z and η = 8.3 × 10 3 . Other parameters are the same as those in Fig. 5. Two discrete dips demonstrate the Autler-Townes splitting, and the splitting value is equal to Ω c = 2 g .
Fig. 9
Fig. 9 The transmission spectrum of the atom layer in the combo structure. The blue solid curve refers to the case of n c = 1 , while the red dashed curve refers to the case of n c = 0 . The areal density of atom is N / σ =8 × 10 12 m 2 , so that = 1 . The atomic layer is located at z 0 = 0.001 λ 0 .
Fig. 10
Fig. 10 VIT transmission spectrum under different losses ( z 0 = 0.0001 λ 0 ). For γ e , m = 1.67 × 10 4.5 ω 0 , cavity loss κ is larger than the decay rate of the atom Γ , so what we can see is only a transmission dip around Δ =0 . The transmission at Δ =0 increases when γ e , m = 1.67 × 10 5 ω 0 . As the dissipations decrease to 1.67 × 10 5.5 ω 0 or less, the splitting becomes significant and VIT phenomena can be distinguished clearly.

Equations (13)

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

H = ω a b | a a | + ω c b | c c | ( Ω p e i ω p t | a b | + Ω c e i ω c t | a c | + H . c . ) ,
χ = | μ a b | 2 N ε 0 V 4 δ ( | Ω c | 2 4 δ Δ ) 4 Δ γ b c 2 + i 8 δ 2 γ a b + i 2 γ b c ( | Ω c | 2 + γ b c γ a b ) | | Ω c | 2 + γ a b γ b c 4 Δ δ + i 2 δ γ a b + i 2 Δ γ b c | 2 .
Ω c = 2 g n c + 1
η = 4 g 2 / γ a b γ b c
η = 4 g 2 / κ Γ a b
χ = | μ a b | 2 N ε 0 V [ 2 δ ˜ ( η δ ˜ Δ ˜ ) 2 Δ ˜ + i 2 ( δ ˜ 2 + η + 1 ) ( η + 1 Δ ˜ δ ˜ ) 2 + ( Δ ˜ + δ ˜ ) ] 1 Γ a b
ε 1 = 2
μ 1 = 1 ω m 2 ω 2 + i γ m ω
ε 2 = 1 ω e 2 ω 2 + i γ e ω , μ 2 = 2
4 g 2 = 2 ω a c 2 μ a c 2 c 2 ε 0 Re μ 1 4 π 0 K 3 K 1 2 K 1 z [ 1 + 2 R 12 T M r 10 T M e 2 i K 1 z d 1 + R 12 T M e 2 i K 1 z z 0 + r 10 T M e 2 i K 1 z ( d 1 z 0 ) 1 R 12 T M r 10 T M e 2 i K 1 z d 1 ] d K | ω = ω a c
Γ ab = 2 ω a b 2 μ a b 2 c 2 ε 0 Re μ 1 8 π 0 K K 1 z [ 1 + 2 R 12 T E r 10 T E e 2 i K 1 z d 1 + R 12 T E e 2 i K 1 z z 0 + r 10 T E e 2 i K 1 z ( d 1 z 0 ) 1 R 12 T M r 10 T M e 2 i K 1 z d 1 ] + K 1 z 2 K 1 2 [ 1 + 2 R 12 T M r 10 T M e 2 i K 1 z d 1 + R 12 T M e 2 i K 1 z z 0 + r 10 T M e 2 i K 1 z ( d 1 z 0 ) 1 R 12 T M r 10 T M e 2 i K 1 z d 1 ] d K | ω = ω a b
R 12 q = r 12 q + r 20 q e 2 i k 2 z d 2 1 + r 12 q r 20 q e 2 i k 2 z d 2
χ = 4 K a b L Δ ˜ ( η Δ ˜ δ ˜ ) δ ˜ i ( η + 1 + δ ˜ 2 ) ( η + 1 + δ ˜ 2 ) 2 + ( Δ ˜ + δ ˜ ) 2
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.