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

Spontaneous feedback for the simultaneous narrowing and elevation of fluorescence spectral lines

Open Access Open Access

Abstract

Narrow linewidth and high intensity of the fluorescence spectra are two important elements to improve the accuracy and efficiency of related practical measurements, but so far they have not been achievable simultaneously. We propose a new approach to sharpen the linewidth and to heighten the intensity simultaneously. Rather than double coherent mechanisms, it uses the spontaneous emission of inverted atoms in a cavity below the threshold as a resource for feedback control. The spontaneous feedback is simpler in principle and easier to realize technologically, and represents a kind of new way of controlling spontaneous emission.

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

1. Introduction

Control and use of spontaneous emission are one of the most important subjects in quantum optics and nonlinear optics [1–7] and have a great deal of timely renewed interest (e.g., see [8–42]). Among others, linewidth narrowing and intensity elevation of fluorescence spectra turn out to be related important aspects to enhance the accuracy and efficiency in spectra-based high-precision measurements. Intuitively, slowing down the spontaneous decay of an atom gives the spectral line narrowing, and the total fluorescence power squashes in the narrow line. So far, existing schemes generally make the spectral lines narrow at a cost of decreasing the power [29–31]. Those schemes are mainly based on doubly coherent mechanisms, one of which splits the spectrum into Mollow triplet [8,9], and the other sharpens the spectral line but also suppresses the intensity. The fall in the intensity is due to the extra spectrum splitting [29] or the coherent population trapping [30,31].

In this article we propose a fundamentally different approach, with no use of coherence. It uses spontaneous emission as a resource for its own control. This spontaneous feedback is simpler but more efficient for the double purpose of sharpening the linewidth and boosting the intensity. Its advantages are listed as follows.

First, the spectral line becomes narrow and high continuously as the atom-cavity system approaches the threshold but remains below threshold. The ultimate limits are due to quantum fluctuations. The spontaneous photons from an ensemble of inverted atoms emit into a cavity mode, which remains amplified below lasing threshold and is coupled back to the atoms. The amplification below threshold makes extremely narrow the otherwise wide spectrum. Neither of incoherent pumping for the inversion and the cavity mode below lasing threshold causes the level splitting. The coupling back to the atoms acts as spontaneous feedback mechanism, which erects an ultrasharp peak over the natural spectrum. At least a half of the total power squashes into the sharp spectral line and thus boosts the spectral intensity.

Second, the spontaneous feedback holds also for the sidebands of resonant fluorescence. When the population inversion is established between dressed states and a cavity is tuned to resonant with the inverted dressed transition, the spontaneous emission from the inverted sideband can be used as a resource for feedback control. Similarly, the cavity mode remains amplified below threshold. The coupling of the cavity mode back to the atoms simultaneously sharpens and raises the fluorescence spectrum at the sidebands.

Thirdly, spontaneous emission as a resource for its own control is much easier to realize in principle and technologically than the doubly coherent schemes. Since the cavity mode is only amplified below lasing threshold, a bad cavity is efficient for the feedback control. In the absence of the coherent population trapping, one no longer needs the nearly parallel dipole moments, which are not existent in nature. Even for the substitutions by the dressed states, the quantum interference will be washed out by the decoherence of the dressed transitions [43,44]. Any added incoherent pump in the doubly coherent mechanisms increases the linewidth [31]. In addition to the cavity, there is no need to use specific reservoirs like the photonic band gap materials [24–28].

Although Haken [45] gave a brief calculation of the cavity linewidth below the threshold in the early days of laser theory, the linewidth holds only in a limited range. First, close to threshold unlimitedly, the linewidth given approaches zero unboundedly. Obviously, this is unphysical. The physical limitation will be given below. Second, far below threshold, the spectrum of the atom-cavity coupling system should be the total contributions of two Lorentz terms and their interference term, as shown in Eq. (7), rather than one single Lorentz spectrum. Only when the narrowing term dominates over other terms, is the linewidth given in [45] appropriate. So far, these two aspects have not yet been included in the existing references [1–7, 45–50]. For the above reason, a more complete and elaborate description is needed in the whole region below threshold. With this, it is convenient for us to apply this linewidth to related schemes, such as the resonance fluorescence of the atoms or molecules in a cavity. Related to this work, we note that great progresses have been made on collective emission and cavity QED [34–42]. For example, using subradiance and selective radiance in atomic arrays makes exponential improvement in photon storage fidelities [34]; using cavity-assisted Raman transitions manipulate spin dynamics and establishes squeezing in a spinor gas [35]; exploring single-photon time-dependent spectra in coupled cavity arrays [36]; detailed calculation and description of Fano-Agarwal couplings and non-rotating wave approximation in single-photon timed Dicke subradiance [37].

The remaining part of this article is organized as follows. In Sec. 2, we first give an analytical expression of the spontaneous emission spectrum of bare atom, and then discuss the limitation of the linewidth and the height of the spectral line. In Sec. 3, we show the applications of the spontaneous feedback mechanism in resonance fluorescence. The three subsections present the narrowing and heightening of the spectral lines of two-level dressed atom, three-level dressed atom and molecule system, respectively. Finally, conclusion is given in Sec. 4.

2. Spontaneous feedback by cavity for a bare atom

The analytical expression of the spontaneous emission is given in Eq. (7). The limitation of the linewidth due to the fluctuation of the population is also given explicitly in Eq. (9). Plotted in Fig. 1(a) is the spontaneous emission spectrum of inverted two-level atoms in a cavity below threshold. The spectrum displays an ultrasharp and ultrahigh peak regardless of large natural linewidths of the atoms and the cavity. The higher the spectral peak the narrower the linewidth. The ultrasharp linewidth is determined by the smallest one of decay rates of the atom-cavity coupled system. As the system approaches its threshold the smallest rate falls close to zero, as shown in Fig. 1(b). The linewidth and height are ultimately limited by the quantum fluctuations. It is clearly shown that the spontaneous feedback is a much easier and technologically less challenging way for obtaining narrow and high spectral lines than previous doubly coherent scheme [29–31]. This mechanism underlying the above effects is presented as follows.

 figure: Fig. 1

Fig. 1 (a) Spontaneous spectrum S(ω) of inverted two-level atoms in a cavity for κ = γ, Λ = 1.2γ, γp = 0.5γ, and C = 2.7225. Note that the vertical axis is in logarithmic coordinate. (b) The smallest decay rate λ1 versus the cooperative parameter C for κ = 0.2γ (solid), γ (dashed), and 5γ (dotted).

Download Full Size | PDF

2.1. Narrow and high spontaneous emission spectrum

Here we consider an ensemble of independent atoms as in [1–6, 49–54]. This should be distinguished from the Dicke model [55], in which identical atoms form a single quantum system, as studied by Tavis and Cummings [56, 57]. The Hamiltonian for the resonant interaction of N independent atoms with the cavity field is written in the dipole approximation and in an interaction picture as

H=g(aσ21+σ12a),
where ħ is the Planck constant and g is the atom-field coupling strength. a and a are the annihilation and creation operators for the cavity field, respectively. σkl=μ=1Nσklμ (σklμ=|kμlμ|; l, k = 1, 2) are the collective spin-flip (kl) and projection (k = l) operators. The independent atoms have independent phases for σ12μ (σ21μ). The master equation for the density operator ρ of the atom-field system is derived as [3] ρ˙=i[H,ρ]+ρ, with all relaxations included in ρ=κ2aρ+μ=1N(γ2σ12μρ+Λ2σ21μρ+γp4σpμρ), where we have used the common form oρ=2oρoooρρoo for o = a, σ12μ, σ21μ, σpμ(=σ22μσ11μ). κ and γ are the cavity and atomic decay rates, respectively, γp is the additional atomic phase damping rate, and Λ is the incoherent pumping rate from the ground state to an additional excited state from which the atoms decay rapidly to the excited state. It is the very way of creating population inversion for lasing action [1–5].

Following the standard technique [2–5] we can study the quantum correlation spectra from the master equation. By arranging operators in the normal order (a, σ21, σ22, σ12, a) and by defining the corresponding c-numbers (α*, v*, z2, v, α), we derive the set of Heisenberg-Langevin equations

α˙=(κ/2)αigv+Fα(t),
v˙=(Γ/2)v+ig(z2z1)α+Fv(t),
z˙2=γz2+Λz1ig(av*va*)+Fz2(t),
together with those for (α*, v*) and the closure relation z1 + z2 = N (z1 is the c-number correspondence of σ11). The two g terms in Eqs. (2) and (3) describe clearly the feedback by the cavity to the collective atoms. Here we have defined Γ = γ + Λ + 2γp. F’s are the noise terms and have vanishing means and noise correlations 〈Fo(t)Fo′ (t′)〉 = Doo′ δ(tt′), where the nonzero diffusion coefficients read as Dv*v = Dvv* = Γ〈z2〉 and Dz2z2 = 2γz2〉.

Here we focus on the case below threshold, as shown later. As in the usual case for independent [1–6,49–54], individual atoms have weak couplings with the field gγ, and thus the atom-field correlations have a negligible effect on the steady state values. For the entire ensemble, it has a total coupling constant gN with the cavity field. While the coupling constant for a single atom is small, the cooperativity parameter C=g2NκΓ can be large if N is large. It is for this reason, many textbooks in quantum optics and laser physics [1–6, 49] separate the collective atomic operators from the field operators and then solve for the steady state. We follow the standard techniques. Our case is completely different from the one-atom case, where the individual atom has strong coupling with the cavity [29,58–60]. We can obtain the steady state solutions and the time evolutions by neglecting the noises temporarily. The steady state solutions are 〈α〉 = 〈v〉 = 0 and 〈z1,2〉 = NP1,2, where P1=γγ+Λ and P2=Λγ+Λ are the mean populations of each atom. It should be noted that the population fluctuations δzl = zl − 〈zl〉 (l = 1, 2) will not be coupled to the quantities α and v because of 〈α〉 = 〈v〉 = 0. Then we can substitute 〈z1,2〉 for zl in Eq. (3). It is easy to find from Eqs. (2,3) that α and v evolve with time according to eλ1,2t, where λ1,2 are the new decay rates of the coupled system

λ1,2=κ+Γ4[114κΓ(14CP)(κ+Γ)2],
with the population difference P = P2P1. The stability requires that Reλ1,2 > 0. We see from Eq. (5) that λ1 > 0 is satisfied for 4CP < 1 and λ2 is always positive. However, once λ1 becomes negative, α and v become amplified. It is the case for 4CP > 1. Therefore there exists a threshold λ1 = 0, i.e., 4CP = 1. Our focus is on the case below and near the threshold, λ1 → 0+, i.e., 4CP → 1, which requires population inversion P2 > P1. Then linearizing the atomic populations around the steady state, we obtain
(d/dt)δz2=(γ+Λ)δz2+Fz2(t),
where the decay rate λ3 = γ + Λ is always positive and so the δz2 decay eλ3t is stable (λ3 > 2γ for Λ > γ).

The incoherent fluorescent field from the N independent atoms is determined by the fluctuating collective spin-flip operator δσ12/N. The fluorescence spectrum [1–5, 51–53] S(ω)=Relimt0dτei(ωω21)τδσ21(t+τ)δσ12(t)/N is derived in the explicit form

S(ω)=η22ω¯2+λ12+η12ω¯2+λ22+2η1η2(ω¯2+λ1λ2)(ω¯2+λ12)(ω¯2+λ22),
where ω̄ = ωω21, ηl > 0, and ηl2=(Γ/2λlλ1λ2)2ΓP2, (l = 1, 2). It should be noted that the calculation given in [45] excludes two latter terms, which can only describe the spectrum appropriately in a limited region. The first two Lorentz lineshapes in Eq. (7) describe the spontaneous emission at different rates λ1,2, while the third term originates from the constructive interference of the two emission processes. However, this interference is weak due to the opposite disparity of λ1,2 and η1,2 near threshold: λ1λ2 and η1η2. As a comparison, we recover the standard spectrum [1–5] S(ω)=ΓP2ω¯2+(Γ/2)2 from Eq. (7) for the case without the cavity coupling, where κ → ∞ and 4CP → 0 lead to λ1 = Γ/2 and λ2 = κ/2 → ∞. Shown in Fig. 1(a) is the spectrum S(ω) for the chosen parameters in the figure caption. These parameters give 4CP = 0.99 and describe the operation below and near threshold. The spectral linewidth is 7.6 × 10−3γ, which is about two thousandth of the natural width Γ = 3.2γ. The peak height is S(ω21) = 6.7 × 103, which is about 1.0 × 104 times as much as the natural peak height 4P2/Γ ∼ 0.625 (for P2 ≳ 0.5). Shown in Fig. 1(b) is the smallest decay rate λ1. As the system is very close to threshold, we have λ1κΓ(14CP)2(κ+Γ)0+, λ212(κ+Γ), η12(Γκ+Γ)2ΓP2, and η22(κκ+Γ)2ΓP2. The spectral linewidth is about d ≐ 2λ1 and the peak height is h=S(ω21)=(η2λ1+η1λ2)2(η2λ1)2. Thus the spectrum is extremely narrow and high even though the involved natural linewidths κ and Γ are large. The spectrum integration gives the total intensity 1πS(ω)dω=η12λ2+η22λ1+2η1η2λ1+λ2η22λ1, which is extremely enhanced. Clearly it is attributed to the amplification of the cavity field below threshold and the feedback to the atoms.

2.2. Linewidth limitation by quantum fluctuation

Although the expression of the spectrum in Eq. (7) is different from that in early studies [45–50], one common problem still remains. The intensity of the spontaneous emission goes to infinity as the linewidth approaches zero unboundedly (η22λ1). This is obviously unphysical. However, it should be emphasized that the above narrowing and heightening of the spectrum are not unbounded, but ultimately limited by quantum fluctuations [3–5]. It is the population fluctuations that set a nonzero limit λ1min even for λ1 → 0+. From Eq. (6) we calculate the z2 variance (δz2)2=NγP2/(γ+Λ)=NγP22/Λ, and then derive the ultimate limit

λ1min4CκΓP2κ+ΓγNΛ,
This limits the spectral width and height to
dmin=2λ1minandhmax(η2λ1min)2,
respectively, which are proportional to CN and NC2, respectively. As an example, for Λ ≳ γ, P212, κ ∼ Γ, C ∼ 1, and N ∼ 108, we have the ultimate limits dmin/Γ ∼ 10−4 and hmax/S0(ω21) ∼ 104.

3. Spontaneous feedback in resonance fluorescence

3.1. Dressed two-level atom in a single-mode cavity

We first show that the narrowing and elevation mechanism holds for resonant fluorescence of two-level atoms. In this case, the ensemble of N-atoms is driven by a classical field with frequency ω0. The total Hamiltonian of the coupled system in the rotating frame of the driving field reads as H = H0 + HI, where H0 = ħΔσ22 + ħΩ(σ12 + σ21) describes the interaction between the atoms and one strong classical field, and HI = ħΔcaa + ħg(21 + σ21a) denotes the interaction of cavity field with the atoms. Of the new parameters, Ω is half Rabi frequency and Δ = ω21ω0 and Δc = ωcω0 are respectively the detunings of the atomic and cavity frequencies ω21 and ωc from the dressing field frequency ω0. It is most convenient to transform to the dressed state picture to show the essential mechanism. We assume the generalized Rabi frequency to be large, Ω˜=Δ2+4Ω2(γ,κ). After diagonalizing H0, we obtain the dressed states as [9]

|1˜=cosθ|1sinθ|2,|2˜=sinθ|1+cosθ|2,
where we have assumed tan(2θ) = 2Ω/Δ, 0 < θ < π/2. The dressed states have well separated eigenvalues ħ(Δ ∓ Ω̃)/2 since Ω̃ ≫ (γ, κ). The fluorescence radiation occurs at ω = ω0, ω0 ± Ω̃, respectively. In the secular approximation, we can rewrite the atomic damping terms into the same form as for the bare atoms, except for substitutions of (γ̃ = γ cos4 θ, Λ̃ = γ sin4 θ, γ˜p=12γsin2(2θ)) for (γ, Λ, γp), respectively. With such substitutions, the dressed state populations 1,2, the decay rate Γ̃, the eigenvalues − λ̃1−3, the parameters η̃1,2, and the cooperativity parameter contained in λ̃1,2 have the same meanings and forms as above. The explicit expressions for 1,2 are P˜1=cos4θcos4θ+sin4θ and 2 = 1 − 1. For Δ < 0 we have 2 > 1. We tune the cavity field Δc = Ω̃ such that the atoms spontaneously emit photons into the cavity mode below threshold. Similarly, for Δ > 0 we tune the cavity field Δc = − Ω̃. Here we exemplify the former case. Performing the rotating transformation associated with the cavity detuning, we can describe explicitly the resonant interaction of the dressed atoms with the cavity field.

The interaction Hamiltonian is derived as

HI=g˜(aσ˜21+σ˜12a),
where σ˜kl=μ=1Nσ˜klμ (σ˜klμ=|kμlμ|) represents the collective spin-flip and projection operators for the dressed atoms, = g cos2 θ is the effective interaction strength. It is seen that the interaction Hamiltonian is of the same form as Hamiltonian (1) except for the dependence of , σ̃12 and σ̃21 on the normalized detuning Δ/Ω. It is such a parameter dependence that manipulates the dressed state populations and thus makes the fluorescence spectra controllable. Using the same techniques as above for the dressed transitions, we can derive the total fluorescence spectrum as
S(ω)=S1(ω¯+Ω˜)sin4θ+S2(ω¯Ω˜)cos4θ+S3(ω¯)sin2(2θ),
where ω̄ = ωω0. S1,2 originate respectively from the dressed transitions at ω0 ∓ Ω̃ and have the same forms as (7) except for the substitutions of the tilde parameters and the respective central frequencies. S3(ω)=4γ˜P˜2ω¯2+λ˜32 stems from dressed transitions at ω0. The linewidths of the sidebands at ω0 ∓ Ω̃ are 2λ̃1 when λ̃1 → 0+. The height ratio of the sidebands is
S1(ω0Ω˜)/S2(ω0+Ω˜)=tan4θ,
which is in agreement with the theoretical prediction [61,62] and experimental observation [63] although no sharp lines are involved there. For the numerical calculation we rewrite the cooperation parameter as C˜=C0cos4θ1+sin2(2θ)/2, in which C0=g2Nκγ is independent of the normalized detuning Δ/Ω. Figure 2 shows the extremely narrow and high sidebands of the fluorescence spectrum below and near threshold. The nearest three-peaked spectrum locates at Δ/Ω = −2.68531, and its sidebands have their common width 2λ̃1 ≐ 1.4 × 10−2γ and their respective heights 5.6 × 103 and 33.

 figure: Fig. 2

Fig. 2 Fluorescence spectrum S(ω) of dressed two-level atoms in a cavity for κ = γ and C0 = 30. We take Ω̃ = 5γ to fix the sidebands for clear show.

Download Full Size | PDF

3.2. Dressed three-level atom in a two-mode cavity

It is straightforward to generalize to the fluorescence spectra of three-level atoms. The atom-field interactions are shown in the insets of Figs. 3(a) and 3(b). The V-type atoms are excited from one common ground state |3〉 to two excited states |1, 2〉, from which they decay back to |3〉. The Λ-type atoms are excited from two ground states |1, 2〉 to one common excited state |3〉, from which they decay back to |1, 2〉. Exemplifying the V-type atoms we perform the numerical verification. The system Hamiltonian reads as

H=H0+HI,
with
H0=l=1,2[Δlσll+Ω(σl3+σ3l)],HI=l=1,2[Δclalal+gl(alσl3+σ3lal)],
where a1,2 and a1,2 are the annihilation and creation operators for the cavity fields, respectively. The two Rabi frequencies are assumed to be equal, Ω1,2 = Ω. We have defined detunings Δl = ω3lωl and Δcl = ωclωl (l = 1, 2). We assume that Δ1 = −Δ2 = Δ and Ω˜=Δ2+2Ω2κ,γ. Diagonalizing H0 gives the dressed states [64]
|+=1+sinθ2|1+1sinθ2|2+cosθ2|3,|0=cosθ2|1+cosθ2|2+sinθ|3,|=1sinθ2|1+1+sinθ2|2cosθ2|3,
which have equally spaced eigenvalues 0, ±Ω̃, respectively. Thus, the fluorescence spectra have the five peaks, which locate symmetrically at ω = ω1, ω1 ± Ω̃, ω1 ± 2Ω̃, respectively. The cavity fields are tuned resonant with the dressed transitions at inner sidebands, respectively, i.e., Δc1 = −Δc2 = Ω̃. The spontaneous photons from the inverted dressed transitions enter the cavity modes below threshold for ΔΩ=(,1). The Λ-type atoms have similar behaviour for ΔΩ=(1,+). Figure 3 displays the sharpening and boosting of the inner sidebands of fluorescence spectra S(ω) for the |1〉 ↔ |3〉 transitions of the two systems. We have assumed equal parameters g1,2 = g, κ1,2 = κ, and γ1,2 = γ, take κ = γ and C0 = 30, and fix Ω̃ = 5γ. In Fig. 3(a), for the nearest five-peaked spectrum at Δ/Ω = −1.1355, its inner sidebands have their common width 4.6 × 10−3γ and their respective heights 7.8 × 103 and 4.1 × 102. In Fig. 3(b), for the nearest five-peaked spectrum at Δ/Ω = 1.1915, its inner sidebands have their common width 4.0 × 10−3γ and their respective heights 9.2 × 103 and 4.6 × 102. Although we show in Fig. 3 the spectra from the |1〉 ↔ |3〉 transitions of the three-level atoms, the same features appear for the spectra from |2〉 ↔ |3〉 transitions (but not shown here) because of the symmetry of the parameters.

The cascade configuration of three-level atoms is also suitable for the present mechanism. In this system, more often than not, we can have one inverted transition between dressed states. So long as conditions are satisfied, the narrowing and elevation of the spontaneous lines are predictable. However, the system has no longer the symmetry as for the V-type or Λ-type cases. Thus, in more cases, we only have narrow lines for one transition channel. In fact, so is the case if we break the symmetry between the two transitions in V-type or Λ-type atoms by not choosing symmetric Rabi frequencies or symmetric detunings.

 figure: Fig. 3

Fig. 3 Fluorescence spectra from the |1〉 ↔ |3〉 transitions of dressed three-level atoms in (a) V and (b) Λ configurations.

Download Full Size | PDF

3.3. Molecule system

Finally, it is possible to apply the above spontaneous feedback to narrow and heighten fluorescence spectra of molecules. Typically, within electronic states there are a number of vibrational levels, which are successively labeled by v = 0, 1, 2, · · · in ascending order. Thus molecular electronic spectra comprise a number of vibrational sub-bands. Non-radiative vibration relaxation, which takes place within the excited electronic state (v′ = 0, 1, 2, · · ·), is fast compared to the rate of emission. It is the case for large molecules in solution such as the dyes, often used in fluorescence spectroscopy [65]. The so-called relaxed fluorescence originates from the v′ = 0 excited vibrational level. Meanwhile, those levels above the v″ = 0 level within the vibrational ground state tend not to be occupied at room temperature for not too low vibrational frequencies. The incoherent pump starts from v″ = 0. It is advantageous for creating the population inversion for the electronic transitions v′ = 0 ↔ v″ > 0. When the cavity bandwidth κ is much less than the adjacent vibrational frequencies within the ground electronic state, the cavity will selectively couple one fluorescence transition v′ = 0 → v″ > 0 and thus leads to fluorescence spectrum narrowing and heightening.

4. Conclusion

In conclusion, we have proposed the spontaneous feedback mechanism for narrowing and heightening the spontaneous spectra. Spontaneous emission itself is used as the control resource. Ultimately, both the linewidth and the peak height are limited by the quantum fluctuations. The mechanism is valid for the single spectral line of the bare atoms, and also for the Mollow-like sidebands of the dressed atoms and the relaxed fluorescence spectra of molecules. In addition to the cavity, there is no aid of any extra added coherent resource. The spontaneous feedback control has loose conditions for experimental realization and practical applications. As our proposal, the dressed two-level atoms have more advantages for the experimental realization compared with the bare two-level atoms and the three-level atoms. For the former, varying the driving field amplitude and/or frequency is a controllable way to create population inversion on one dressed transition and select the cavity frequency [66,67]. The three-level schemes are relatively complicated because two driving fields and two cavity fields are employed.

Funding

National Natural Science Foundation of China (Grants No. 11474118 and No. 61178021); National Basic Research Program of China (Grant No. 2012CB921604).

References and links

1. R. Loudon, The Quantum Theory of Light, 3rd ed. (Oxford University, 2000).

2. D. F. Walls and G. J. Milburn, Quantum Optics (Springer, 1994).

3. M. O. Scully and M. S. Zubairy, Quantum Optics (Cambridge University, 1997). [CrossRef]  

4. P. Meystre and M. Sargent III, Elements of Quantum Optics, 4th ed. (Springer, 2007).

5. C. W. Gardiner and P. Zoller, Quantum Noise, 2nd ed. (Springer, 2000). [CrossRef]  

6. H. J. Carmichael, Statistical Methods in Quantum Optics, 2nd ed. (Springer, 2002).

7. R. W. Boyd, Nonlinear Optics, 3rd ed. (Elsevier, 2008).

8. B. R. Mollow, “Power spectrum of light scattered by two-level systems,” Phys. Rev. 188, 1969–1975 (1969). [CrossRef]  

9. C. Cohen-Tannoudji, J. Dupont-Roc, and G. Grynberg, Atom-photon Interactions (Wiley, 1992).

10. P. Goy, J. M. Raimond, M. Gross, and S. Haroche, “Observation of cavity-enhanced single-atom spontaneous emission,” Phys. Rev. Lett. 50, 1903–1906 (1983). [CrossRef]  

11. M. Lewenstein, T. W. Mossberg, and R. J. Glauber, “Dynamical suppression of spontaneous emission,” Phys. Rev. Lett. 59, 775–778 (1987). [CrossRef]   [PubMed]  

12. D. J. Heinzen, J. J. Childs, J. E. Thomas, and M. S. Feld, “Enhanced and inhibited visible spontaneous emission by atoms in a confocal resonator,” Phys. Rev. Lett. 58, 1320 (1987). [CrossRef]   [PubMed]  

13. For an overview, see Cavity Quantum Electrodynamics, edited by P. R. Berman, ed. (Academic, 1994).

14. E. Arimondo, Progress in Optics, Vol. 35, E. Wolf, ed. (Elsevier Science, 1996), pp. 257–354. [CrossRef]  

15. S. Y. Zhu and M. O. Scully, “Spectral line elimination and spontaneous emission cancellation via quantum interference,” Phys. Rev. Lett. 76, 388–391 (1996). [CrossRef]   [PubMed]  

16. H. R. Xia, C. Y. Ye, and S. Y. Zhu, “Experimental observation of spontaneous emission cancellation,” Phys. Rev. Lett. 77, 1032–1034 (1996). [CrossRef]   [PubMed]  

17. L. Li, X. Wang, J. Yang, G. Lazarov, J. Qi, and A. M. Lyyra, “Comment on ”Experimental observation of spontaneous emission cancellation”,” Phys. Rev. Lett. 84, 4016 (2000). [CrossRef]  

18. S. M. Barnett and R. Loudon, “Sum rule for modified spontaneous emission rates,” Phys. Rev. Lett. 77, 2444–2446 (1996). [CrossRef]   [PubMed]  

19. P. Mataloni, E. De Angelis, and F. De Martini, “Bose-Einstein partition statistics in superradiant spontaneous emission,” Phys. Rev. Lett. 85, 1420–1423 (2000). [CrossRef]   [PubMed]  

20. M. Bayer, T. L. Reinecke, F. Weidner, A. Larionov, A. McDonald, and A. Forchel, “Inhibition and enhancement of the spontaneous emission of quantum dots in structured microresonators,” Phys. Rev. Lett. 86, 3168–3171 (2001). [CrossRef]   [PubMed]  

21. E. Frishman and M. Shapiro, “Complete suppression of spontaneous decay of a manifold of states by infrequent interruptions,” Phys. Rev. Lett. 87, 253001 (2001). [CrossRef]   [PubMed]  

22. E. Frishman and M. Shapiro, “Suppression of the spontaneous emission of atoms and molecules,” Phys. Rev. A 68, 032717 (2003). [CrossRef]  

23. D. G. Norris, L. A. Orozco, P. Barberis-Blostein, and H. J. Carmichael, “Observation of ground-state quantum beats in atomic spontaneous emission,” Phys. Rev. Lett. 105, 123602 (2010). [CrossRef]   [PubMed]  

24. T. Quang, M. Woldeyohannes, S. John, and G. S. Agarwal, “Coherent control of spontaneous emission near a photonic band edge: a single-atom optical memory device,” Phys. Rev. Lett. 79, 5238–5241 (1997). [CrossRef]  

25. E. P. Petrov, V. N. Bogomolov, I. I. Kalosha, and S. V. Gaponenko, “Spontaneous emission of organic molecules embedded in a photonic crystal,” Phys. Rev. Lett. 81, 77–80 (1998). [CrossRef]  

26. M. D. Leistikow, A. P. Mosk, E. Yeganegi, S. R. Huisman, A. Lagendijk, and W. L. Vos, “Inhibited spontaneous emission of quantum dots observed in a 3D photonic band gap,” Phys. Rev. Lett. 107, 193903 (2011). [CrossRef]   [PubMed]  

27. M. R. Jorgensen, Jeremy W. Galusha, and M. H. Bartl, “Strongly modified spontaneous emission rates in diamond-structured photonic crystals,” Phys. Rev. Lett. 107, 143902 (2011). [CrossRef]   [PubMed]  

28. P. Campagne-Ibarcq, S. Jezouin, N. Cottet, P. Six, L. Bretheau, F. Mallet, A. Sarlette, P. Rouchon, and B. Huard, “Using spontaneous emission of a qubit as a resource for feedback control,” Phys. Rev. Lett. 117060502 (2016). [CrossRef]   [PubMed]  

29. H. Freedhoff and T. Quang, “Ultrasharp lines in the absorption and fluorescence spectra of an atom in a cavity,” Phys. Rev. Lett. 72, 474–477 (1994). [CrossRef]   [PubMed]  

30. P. Zhou and S. Swain, “Ultranarrow spectral lines via quantum interference,” Phys. Rev. Lett. 77, 3995–3998 (1996). [CrossRef]   [PubMed]  

31. C. H. Keitel, “Narrowing spontaneous emission without intensity reduction,” Phys. Rev. Lett. 83, 1307–1310 (1999). [CrossRef]  

32. A. González-Tudela and D. Porras, “Mesoscopic entanglement induced by spontaneous emission in solid-state quantum optics,” Phys. Rev. Lett. 110, 080502 (2013). [CrossRef]   [PubMed]  

33. M. O. Scully, “Single Photon Subradiance: Quantum control of spontaneous emission and ultrafast readout,” Phys. Rev. Lett. 115, 243602 (2015). [CrossRef]   [PubMed]  

34. A. Asenjo-Garcia, M. Moreno-Cardoner, A. Albrecht, H. J. Kimble, and D. E. Chang, “Exponential improvement in photon storage fidelities using subradiance and selective radiance in atomic arrays,” Phys. Rev. X 7, 031024 (2017).

35. S. J. Masson, M. D. Barrett, and S. Parkins, “Cavity QED engineering of spin dynamics and squeezing in a spinor gas,” Phys. Rev. Lett. 119, 213601 (2017). [CrossRef]   [PubMed]  

36. I. M. Mirza, S. J. van Enk, and H. J. Kimble, “Single-photon time-dependent spectra in coupled cavity arrays,” J. Opt. Soc. Am. B 30, 2640–2649 (2013). [CrossRef]  

37. I. M. Mirza and T. Begzjav, “Fano-Agarwal couplings and non-rotating wave approximation in single-photon timed Dicke subradiance,” Europhys. Lett. 114, 24004 (2016). [CrossRef]  

38. S. Zeeb, C. Noh, A. S. Parkins, and H. J. Carmichael, “Superradiant decay and dipole-dipole interaction of distant atoms in a two-way cascaded cavity QED system,” Phys. Rev. A 91, 023829 (2015). [CrossRef]  

39. M. Bradford and Jung-Tsung Shen, “Spontaneous emission in cavity QED with a terminated waveguide,” Phys. Rev. A 87, 063830 (2013). [CrossRef]  

40. J. C. A. Carvalho, A. Laliotis, M. Chevrollier, M. Ori, and D. Bloch, “Backward-emitted sub-Doppler fluorescence from an optically thick atomic vapor,” Phys. Rev. A 96, 043405 (2017). [CrossRef]  

41. W. Guerin, M. O. Araújo, and R. Kaiser, “Subradiance in a large cloud of cold atoms,” Phys. Rev. Lett. 116, 083601 (2016). [CrossRef]   [PubMed]  

42. M. O. Araújo, I. Krešié, R. Kaiser, and W. Guerin, “Superradiance in a large and dilute cloud of cold atoms in the linear-optics regime,” Phys. Rev. Lett. 117, 073002 (2016). [CrossRef]   [PubMed]  

43. M. Fleischhauer, C. H. Keitel, L. M. Narducci, M. O. Scully, S. Y. Zhu, and M. S. Zubairy, “Lasing without inversion: interference of radiatively broadened resonances in dressed atomic systems,” Opt. Commun. 94, 599–608 (1992). [CrossRef]  

44. X. M. Hu and J. S. Peng, “Quantum interference from spontaneous decay in Lambda systems: realization in the dressed-state picture,” J. Phys. B: At. Mol. Opt. Phys. 33, 921–931 (2000). [CrossRef]  

45. H. Haken, “A nonlinear theory of laser noise and coherence. I,” Z. Phys. 181, 96–124 (1964). [CrossRef]  

46. W. H. Louisell, A. Yariv, and A. E. Siegman, “Quantum fluctuations and noise in parametric processes. I.,” Phys. Rev. 124, 1646–1654 (1961). [CrossRef]  

47. J. P. Gordon, W. H. Louisell, and L. R. Walker, “Quantum fluctuations and noise in parametric processes. II,” Phys. Rev. 129, 481–485 (1963). [CrossRef]  

48. M. Lax and W. H. Louisell, “Quantum noise. XII. Density-operator treatment of field and population fluctuations,” Phys. Rev. 185, 568–591 (1969). [CrossRef]  

49. W. H. Louisell, Quantum Statistical Properties of Radiation (Wiley, 1990).

50. M. Sargent III, M. O. Scully, and W. E. Lamb Jr., Laser Physics (Addison-Wesley, 1974).

51. R. Bonifacio and L. A. Lugiato, “Optical bistability and cooperative effects in resonance fluorescence,” Phys. Rev. A 18, 1129–1144 (1978). [CrossRef]  

52. H. J. Carmichael, D. F. Walls, P. D. Drummond, and S. S. Hassan, “Quantum theory of optical bistability. III. Atomic fluorescence in a low-Q cavity,” Phys. Rev. A 27, 3112–3128 (1983). [CrossRef]  

53. M. D. Reid, “Quantum theory of optical bistability without adiabatic elimination,” Phys. Rev. A 37, 4792–4818 (1988). [CrossRef]  

54. A. Dantan and M. Pinard, “Quantum-state transfer between fields and atoms in electromagnetically induced transparency,” Phys. Rev. A 69, 043810 (2004). [CrossRef]  

55. R. H. Dicke, “Coherence in spontaneous radiation processes,” Phys. Rev. 93, 99–110 (1954). [CrossRef]  

56. M. Tavis and F. W. Cummings, “Exact solution for an N-moleculeradiation-field hamiltonian,” Phys. Rev. 170, 379–384 (1968). [CrossRef]  

57. M. Tavis and F. W. Cummings, “Approximate solutions for an N-molecule-radiation-field hamiltonian,” Phys. Rev. 188, 692–695 (1969). [CrossRef]  

58. C. M. Savage, “Resonance fluorescence spectrum of an atom strongly coupled to a cavity,” Phys. Rev. Lett. 63, 1376–1379 (1989). [CrossRef]   [PubMed]  

59. Y. Mu and C. M. Savage, “One-atom lasers,” Phys. Rev. A 46, 5944–5954 (1992). [CrossRef]   [PubMed]  

60. J. McKeever, A. Boca, A. D. Boozer, J. R. Buck, and H. J. Kimble, “Experimental realization of a one-atom laser in the regime of strong coupling,” Nature 425, 268–271 (2003). [CrossRef]   [PubMed]  

61. P. Zhou and S. Swain, “Dynamics of a driven two-level atom coupled to a frequency-tunable cavity,” Phys. Rev. A 58, 1515–1530 (1998). [CrossRef]  

62. H. Freedhoff and T. Quang, “Steady-state resonance fluorescence spectrum of a two-level atom in a cavity,” J. Opt. Soc. Am. B 10, 1337–1346 (1993). [CrossRef]  

63. A. Lezama, Y. Zhu, S. Morin, and T. W. Mossberg, “Cavity-perturbed strong-field resonance fluorescence,” Phys. Rev. A 39, R2754–R2757 (1989). [CrossRef]  

64. G. L. Cheng, X. M. Hu, W. X. Zhong, and Q. Li, “Two-channel interaction of squeeze-transformed modes with dressed atoms: Entanglement enhancement in four-wave mixing in three-level systems,” Phys. Rev. A 78, 033811 (2008). [CrossRef]  

65. J. L. McHale, Molecular Spectroscopy (Prentice-Hall, 1999).

66. D. J. Gauthier, Q. Wu, S. E. Morin, and T. W. Mossberg, “Realization of a continuous-wave, two-photon optical laser,” Phys. Rev. Lett. 68, 464–467 (1992). [CrossRef]   [PubMed]  

67. W. Lange and H. Walther, “Observation of dynamic suppression of spontaneous emission in a strongly driven cavity,” Phys. Rev. A 48, 4551–4554 (1993). [CrossRef]   [PubMed]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (3)

Fig. 1
Fig. 1 (a) Spontaneous spectrum S(ω) of inverted two-level atoms in a cavity for κ = γ, Λ = 1.2γ, γp = 0.5γ, and C = 2.7225. Note that the vertical axis is in logarithmic coordinate. (b) The smallest decay rate λ1 versus the cooperative parameter C for κ = 0.2γ (solid), γ (dashed), and 5γ (dotted).
Fig. 2
Fig. 2 Fluorescence spectrum S(ω) of dressed two-level atoms in a cavity for κ = γ and C0 = 30. We take Ω̃ = 5γ to fix the sidebands for clear show.
Fig. 3
Fig. 3 Fluorescence spectra from the |1〉 ↔ |3〉 transitions of dressed three-level atoms in (a) V and (b) Λ configurations.

Equations (16)

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

H = g ( a σ 21 + σ 12 a ) ,
α ˙ = ( κ / 2 ) α i g v + F α ( t ) ,
v ˙ = ( Γ / 2 ) v + i g ( z 2 z 1 ) α + F v ( t ) ,
z ˙ 2 = γ z 2 + Λ z 1 i g ( a v * v a * ) + F z 2 ( t ) ,
λ 1 , 2 = κ + Γ 4 [ 1 1 4 κ Γ ( 1 4 CP ) ( κ + Γ ) 2 ] ,
( d / d t ) δ z 2 = ( γ + Λ ) δ z 2 + F z 2 ( t ) ,
S ( ω ) = η 2 2 ω ¯ 2 + λ 1 2 + η 1 2 ω ¯ 2 + λ 2 2 + 2 η 1 η 2 ( ω ¯ 2 + λ 1 λ 2 ) ( ω ¯ 2 + λ 1 2 ) ( ω ¯ 2 + λ 2 2 ) ,
λ 1 min 4 C κ Γ P 2 κ + Γ γ N Λ ,
d min = 2 λ 1 min and h max ( η 2 λ 1 min ) 2 ,
| 1 ˜ = cos θ | 1 sin θ | 2 , | 2 ˜ = sin θ | 1 + cos θ | 2 ,
H I = g ˜ ( a σ ˜ 21 + σ ˜ 12 a ) ,
S ( ω ) = S 1 ( ω ¯ + Ω ˜ ) sin 4 θ + S 2 ( ω ¯ Ω ˜ ) cos 4 θ + S 3 ( ω ¯ ) sin 2 ( 2 θ ) ,
S 1 ( ω 0 Ω ˜ ) / S 2 ( ω 0 + Ω ˜ ) = tan 4 θ ,
H = H 0 + H I ,
H 0 = l = 1 , 2 [ Δ l σ l l + Ω ( σ l 3 + σ 3 l ) ] , H I = l = 1 , 2 [ Δ c l a l a l + g l ( a l σ l 3 + σ 3 l a l ) ] ,
| + = 1 + sin θ 2 | 1 + 1 sin θ 2 | 2 + cos θ 2 | 3 , | 0 = cos θ 2 | 1 + cos θ 2 | 2 + sin θ | 3 , | = 1 sin θ 2 | 1 + 1 + sin θ 2 | 2 cos θ 2 | 3 ,
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.