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

Four-wave-mixing between the upper excited states in a ladder-type atomic configuration

Open Access Open Access

Abstract

Four-wave-mixing (FWM) radiation is generated between the hyperfine structures of the 5D and 5P states in a thermally broadened rubidium atomic vapor using resonant atomic coherence. Background-free unidirectional signals having narrow spectral linewidths are isolated and experimentally studied in the frequency domain, and the effects of the driving beam parameters on the properties of the radiation are discussed. The radiation has several new properties compared to traditional FWM radiations generated between the 5P and 5S states. The high-resolution signals obtained in this method could make it favorable in spectroscopic procedures that rely on two-photon fluorescence.

©2012 Optical Society of America

1. Introduction

Multilevel atoms have been extensively used to host nonlinear wave mixing processes. The atomic coherence established between dipole-forbidden energy levels by two coherent optical fields gives rise to important phenomena such as coherent population trapping and electromagnetically induced transparency (EIT) [13]. Depending upon the relative position of the intermediate energy level, such coherence creates a polarization in the medium at the sum or difference frequency of the coherence-inducing waves. The driven oscillation efficiently mixes with additional fields, giving rise to higher-order such as four- and six-wave mixing processes [48]. The efficient matter-field coupling and the suppressed absorption by virtue of the EIT make it possible to perform nonlinear optical processes using low-power continuous-wave (CW) beams. This feature is important in obtaining high-resolution spectra, and has been an important alternative to using high-powered pulses that, while enhancing the higher-order nonlinearities, can have large frequency bandwidths and can significantly alter the atomic energy levels via effects such as power-broadening and Autler-Townes (AT) splitting.

Wave-mixing processes facilitated by atomic coherence have been explored and characterized in many energy-level configurations such as lambda, double-lambda, ladder, Y and inverted-Y type systems [48]. Many important and useful properties have been identified, including multi-mode quantum correlations and entanglement, photonic memory and phase-controllable interference between different higher-order nonlinear processes [911]. In these processes, the signal is always generated at the transition between the first excited state and the ground state.

In the current experimental work, we parametrically amplify the vacuum mode between two excited states in a three-level ladder-type configuration (Fig. 1(a) ) and study it in the frequency domain. Atomic coherence and EIT mechanism are implemented, allowing the amplification to occur using low-power CW beams. We have investigated the spectral response of the generated waveform to the various contributing parameters, such as the spectroscopic properties (and multi-level structures) of the atomic energy levels, the vicinity of the laser frequencies to the various atomic resonances, and the powers of the driving beams. We consider their contributions to the properties of the generated radiations (such as the efficiency, line shape and linewidth) and discuss the optimum conditions suitable for this process. The generated radiation, containing high-resolution narrow-linewidth spectroscopic information, is background-free. These features can make this process desirable over other spectroscopic methods relying on two-photon fluorescence where the signals are typically very weak since the photons are scattered in all spatial directions and in any given detection direction, the background noise can be of comparable intensity with the signal.

 figure: Fig. 1

Fig. 1 (a) Simplified three-level ladder-type configuration in 85Rb that is coherently driven in the FWM process; |g>, |i> and |e> stand for the ground, intermediate and excited states, respectively. E1, E1 and E2 are external driving beams from laser sources, while Ef is the atom-radiated FWM signal that is parametrically amplified from the vacuum mode. (b) Schematic of the experimental configuration showing the directions and polarizations of the four fields. (c)-(f) Realistic energy level diagram showing the hyperfine (hf) levels of each driven state, as well as the incoherent decay channels (wavy arrows) between various combinations of driven hf levels: [f, f′, f′′] = (c) [3,2,2], (d) [3,3,4], (e) [3,4,3], and (f) [3,4,4]. The single-resonance decay channels are drawn first, followed by the double-resonance decay channels. The number of single-resonance and double-resonance optical pumping channels is different in the various cases. The spacings between the energy levels are not drawn to scale.

Download Full Size | PDF

Recently, it was shown that the FWM radiation generated in this atomic configuration can be effectively phase-modulated in the frequency domain by implementing interferometric control in the driving beams [12]. The system was used to demonstrate various features such as line shape symmetrization, linewidth narrowing and bandwidth switching. The versatile potential of this FWM configuration for Fourier-domain applications is one of our main motivations for experimentally characterizing it, and we believe that the results will be of general interest in the study of nonlinear wave mixing phenomena in multi-level atoms. An energy level configuration similar to the current work was studied in the time domain in Ref [13]. Analytical solutions to the system were derived in [14] with approximations such as very weak driving beams and small hyperfine coupling, both of which are different from our current experimental conditions.

In Fig. 1(a), we have only shown the closed three-level atomic system consisting of the fine structures being driven by the laser beams. In reality, each of these structures consists of a myriad of hyperfine (hf) levels due to coupling with the nuclear magnetic moment, most of which can radiatively decay to energy levels not being driven in the FWM process, making this a mixed system (i.e. consisting of closed as well as open sub-systems.) How open a driven sub-system is depends on the selection rules for the associated hf levels. The hf levels f, f′ and f′′ of the ground, intermediate and excited states, respectively, as well as the various decay channels associated with four different three-level subsystems, are shown in Figs. 1(c)-1(f). In each diagram, the decay channels of the single-photon transition are drawn first, followed by the decay channels due to the two-photon process. From these examples, it can be seen that some FWM subsystems have more decay channels than the others, and that various channels exist via which the atomic population gets optically pumped into the undriven ground state f = 2. The single-photon transition f = 3 → f′ = 4 is closed, whereas f = 3 → f′ = 2, 3 can also radiatively decay to f = 2. A peculiar feature of the ladder-type configuration consisting of multiple sublevels is the so-called double resonance optical pumping (DROP) effect [15]. Beacause of DROP, even when the single-photon transition is closed, the two-photon process opens various optical pumping channels as shown in Figs. 1(e)-1(f). The 5D state also has decay channels via the 6P state, which are not shown in Fig. 1.

Furthermore, due to the close proximity of the hf energies in the 5D and 5P states, many subsystems are simultaneously resonant for a given pair of driving beam frequencies. The hf levels further consist of varying numbers of Zeeman sublevels with different transition strengths. As will be shown below, the multi-level nature as well as the fact that most of the driven subsystems can radiatively decay and lose population to the environment have important consequences to the Fourier-domain waveforms of the atom-radiated coherent FWM signal including the line shape and efficiency at various spectral positions, giving rise to a rich array of spectra. Accurately reproducing the line shapes analytically or numerically thus involves the bookkeeping of all the subsystem parameters, and is beyond the scope of this current experimental work. The intention of this current article is to illuminate the experimentally observed properties of this interesting FWM system and to qualitatively assess the underlying causes of the most important features.

2. Experimental methods

The energy level configuration and the experimental geometry are shown in Fig. 1. The probe beam E1 (frequency ω1, wavelength λ1, wave vector k1) is generated by a CW diode laser DL. The wavelength λ1 is scanned around 780 nm in order to probe the Doppler-broadened spectral bandwidth of the 85Rb isotope’s D2 transition. The vapor cell is 5 cm long, and is magnetically shielded and heated to 60° C. The transmitted probe beam intensity is monitored by a photodiode PD. A strong beam E2 (frequency ω2, wavelength λ2, wave vector k2) from a CW Ti-Sapphire laser is aligned to counterpropagate with E1. The wavelength of E2 is fixed but can be tuned around 776.158 nm, the wavelength of the upper transition in the cascade scheme. When both one-photon resonance and two-photon resonance (TPR) are satisfied, the coupling beam renders the atomic medium transparent for the probe beam by virtue of EIT. Even though the D2 absorption line is Doppler broadened in the hot atomic medium, the two-beam counterpropagating geometry allows the two-photon EIT process to be basically Doppler-free for the cascade configuration [16]. When the probe beam’s frequency detuning is larger than the Doppler-broadened linewidth of the D2 transition, the single photon absorption is vanishingly small. Here, when the strong coupling beam is present satisfying TPR, we no longer have EIT. Instead, a direct two-photon transition is driven between |g> and |e> resulting in a two-photon absorption (TPA) peak. This Doppler-free TPA resonance has a much narrower linewidth than the Doppler-broadened D2 line’s absorption linewidth, and the TPA depth can be tuned via the coupling beam’s intensity. For intermediate frequency detunings lying between the EIT and TPA regimes, both stepwise (via |i>) and direct transitions from |g> to |e> are driven. At these frequency detunings, we observe a convolution of EIT and TPA in the transmission of the probe beam. The analytical solution showing the evolution of the TPR from EIT to TPA as the intermediate frequency detuning is increased can be found in [16].

The output of the diode laser DL is split to create a third beam E1 (frequency ω1, wavelength λ1, wave vector k1′) that intersects with E1 and E2 inside the vapor cell at a small angle of θ (typically 0.4°) with k1. The polarizations and powers of the beams can be altered independently. The third-order nonlinearity of the atomic medium, made efficient by the induced resonant coherences, leads to the generation of a new FWM radiation Ef χ(3)E1′E2E1 which counterpropagates with E1 (due to conservation of linear momentum satisfying kf = k2 + k1 - k1) but has the frequency of E2 (due to conservation of energy satisfying ωf = ω2 + ω1- ω1). In the EIT regime, the FWM process is enhanced because the transitions can be driven near the atomic resonances, while dissipation from |i> as well as |e> are vanishingly small. This allows the χ(3) nonlinear optical process to be driven and measured at low intensities. In the TPA regime, the dissipation from |i> is small due to the large intermediate frequency detuning. Here, when only E1 and E2 are present but E1 is absent, the TPR coherence between |g> and |e> radiatively decays causing incoherent fluorescence scattering. The presence of E1 stimulates coherent FWM radiation in the phase-matched direction. As will be shown below, the spectral linewidth of the FWM radiation is similar to the TPA linewidth, governed basically by the linewidth of the |e> state.

By appropriate choices of the polarizations of the driving beams, the polarization of the FWM signal Ef is made to be orthogonal to that of E1. This allows for an effective isolation of the weak signal Ef using a polarization beam splitter, and is monitored with an avalanche photodiode APD. The voltage measurements of PD and APD are monitored simultaneously using a multi-channel oscilloscope, along with a reference Fabry-Perot cavity signal used for frequency calibration of the scanned DL output.

3. Experimental observations and discussions

Figure 2 shows the line-shapes of the generated FWM signal at various frequency detunings of the intermediate resonance, Δ1, where Δ1 = ω1 – ωig, with ωig being the transition frequency between the ground state and the first excited state. The Doppler-broadened absorption profile of the probe beam is also shown in the figure for reference, and gives information about the spectral position of the intermediate resonance used in the two-photon and FWM processes. Here, the absorption and FWM signals correspond to the f = 3 ground state of the 85Rb isotope. We note that while the intermediate-state detuning is different in each signal, all of them are two-photon-resonant (TPR); that is, the signal occurs only when Δ1 + Δ2 = 0, where Δ2 = ω2 – ωei and ωei is the transition frequency between the two excited states in the cascade configuration. As a result, even though the absorption profile is Doppler-broadened for the heated atomic ensemble, the generated FWM radiation has a line-shape that is Doppler-free. For convenience, the cases Δ1 < 0, Δ1 = 0 and Δ1 > 0 will be referred to as “red-detuned”, “zero-detuned” and “blue-detuned”, respectively.

 figure: Fig. 2

Fig. 2 FWM signal line shape, linewidth and efficiency at four different intermediate frequency detunings, placed at intervals of 500 MHz. All other experimental parameters are constant in the four cases. The Doppler-broadened absorption linewidth of the corresponding ground state (hf = 3 of 85Rb, 5S1/2) is also shown for reference.

Download Full Size | PDF

In obtaining the four signals shown in Fig. 2, the only experimental parameter being varied is the value of ω2, which causes TPR to occur at different values of Δ1 as ω1 is being scanned. Except for their spectral positions, occurring at 500 MHz intervals, all the other experimental conditions, such as the vapor cell temperature (60 °C), beam powers (P1 = 7.4 mW, P1′ = 11.6 mW and P2 = 40 mW) and beam geometry (θ = 0.4°), are identical. The importance of the intermediate frequency detuning is evident in the line shape, linewidth and efficiency of the FWM process. Far from intermediate state resonance, the signal has a narrow linewidth. As the condition Δ1 = 0 is approached, the signal’s linewidth becomes broader and the line shape becomes significantly convoluted. In the region slightly red-detuned from center, the FWM signal intensity also sharply decreases, experiencing a local minima.

FWM signals occurring at 100 MHz intervals are presented in Fig. 3(a) , showing the frequency-detuning dependent trends in more detail. All other experimental parameters are the same as those used in Fig. 2. At each spectral position, the signal is a convolution of a sharp “right” peak and a broad “left” peak. Each pair of dots connected by a line corresponds to the maximum intensities of the sharp and broad peaks occurring within a signal at a given detuning. Below, we will discuss the dependence of the FWM signal’s (1) linewidth, (2) line shape and (3) efficiency upon the driving beam parameters as well as upon the internal structure of the atoms, and also (4) consider the dual role of the driving beam E1.

 figure: Fig. 3

Fig. 3 (Color online) TPR FWM signals corresponding to different values of Δ1, separated by 100 MHz each. The energy levels driven are |g> = 5S1/2, hf = 3, |i> = 5P3/2 and |e> = 5D3/2. At each value of the frequency detuning, the FWM signal is a convolution of a sharp, strong peak and a broad, weak peak, the maximum intensities of which are denoted by a blue square and a red dot, respectively. Each of the two signal peak trends are connected by lines to aid the eye. The two peak values corresponding to a given convolution at a given frequency detuning are connected by a solid black line. The horizontal (vertical) dotted lines identify the intensities (intermediate frequency detunings) of the maximas and minimas of the signal convolution’s two peaks. (a) P1 = 7.4 mW, P1′ = 11.6 mW, P2 = 38 mW (b) P1 = 3 mW, P1′ = 4 mW, P2 = 55 mW. Note the change of scale in the intensity axes.

Download Full Size | PDF

3.1. Linewidth variations

Towards the center of the Doppler-broadened linewidth, as the condition Δ1 = 0 is approached, power-related effects, such as power broadening and AT splitting of the atomic energy levels, become dominant [17]. These effects are also revealed in the broadening and splitting of the FWM signal in the zero-detuned region. For large powers of E1 and E1, when Δ1 = 0 is satisfied, the signal occurs at the AT-satellites [18] of the energy levels. There is a decrease in signal intensity in the spectral region occurring between the power-broadened AT satellites; that is, the signal maximum is displaced around Δ1 + Δ2 = 0. When |Δ1| >> 0 as in the edges of the Doppler-broadened absorption linewidth and outside it, the EIT evolves into a two-photon-absorption (TPA) [16] having a narrow linewidth since power broadening is substantially reduced. Here the signal occurs within the linewidth of the TPA resonance and the signal maxima occurs at Δ1 + Δ2 = 0. In particular, the power- broadening or splitting of the intermediate level is minimal in this two-photon resonant condition, and the linewidth of the FWM signal is mainly limited by the linewidth of the upper-excited state. At intermediate detunings |Δ1| > 0, the probe beam experiences a convolution of EIT and TPA effects, and the generated FWM signal also displays the contributions due to these two mechanisms. Here, both direct two-photon transition from |g> to |e>, as well as stepwise transitions via |i> exist; the direct two-photon transition’s linewidth is narrower as it depends on the relatively long-lived 5D state (natural linewidth 0.97 MHz), whereas the stepwise transition is broader because it also depends on the 5P state (natural linewidth 6 MHz).

3.2. Line shape asymmetries

In order to understand the asymmetries in the spectral line-shape of the generated FWM radiation, one needs to consider the multi-level structure of the atoms as shown in Figs. 1(c)-1(f). For different values of ω2, different hf levels of the intermediate state are closest to satisfy TPR and contribute to the FWM process most effectively. The hf levels of the upper excited state are sufficiently close and lie within the power-broadened linewidth of the intermediate level, and all contribute to the TPR, whereas the intermediate state hf levels are further apart and dispersed within the Doppler-broadened absorption window. The hf levels have varying multiplicities and disparate transition strengths; it is these spectroscopic characteristics of the atomic energy levels that contribute to the sharp asymmetries in the signal line-shape.

The TPR effects involving f′ = 4 has the biggest contribution to the convoluted FWM line-shape, due to its large multiplicity and larger Clebsch-Gordon coefficients. More importantly, as shown in Fig. 1, the FWM pathways involving f′ = 4 have the fewest decay channels. This is why the signal generation (Fig. 2 and Fig. 3) is strongest in the blue-detuned region of the Doppler width, as this is where the f′ = 4 level lies. This is also why the sharp peak lies towards the right edge of each signal convolution for the chosen spin levels in this configuration. The effects of the branching ratios of the energy levels in the ladder-type system have been analyzed by Noh and Moon [19], showing signal convolutions due to the presence of closed and open subsystems.

To make these facts more evident, we have also generated signals by using other spin-levels having different constraints. First, in Figs. 4(a) -4(b), we change |g> to the other ground state hf level 2, while using the same |i> and |e> fine structures as used for Fig. 3. Here, we observe that the position of the sharp peak within the signal convolution occurs at the red-detuned side. Also, the position within the Doppler width where signal intensity is at a maximum, is in the red-detuned region. These changes occur because here, it is the transition f = 2 → f′ = 1 that is closed. Next, we use the same |g> and |i> as that used in Fig. 3, but change the upper excited state |e> to the other fine structure of 5D, i.e. 5D5/2 (Figs. 4(c)-4(d)). Here, similar to Fig. 3, the maximum signal intensity occurs in the blue-detuned region of the Doppler-width since the transition f = 3 → f′ = 4 is closed. However, in this case, the maximum peak within the signal convolution occurs in the red-detuned side. This change occurs because the energies of the hf levels in 5D5/2 are inverted; that is, in this fine structure, the hf levels with higher values have lower energies [20]. This causes the contribution to the FWM signal from the higher f′′ levels, which have allowed transitions from f′ = 4 of |i> as well as larger Zeeman multiplicities, to shift to the lower-energy side of the signal convolution. This is in contrast to Fig. 3, where |e> corresponded to 5D3/2 in which the hf levels are not inverted. The stronger FWM signal intensity observed when |e> is 5D5/2 may be attributed to it having fewer decay channels; while 5D3/2 can decay to both J = 3/2 and 1/2 of the 5P and 6P levels, the 5D5/2 fine structure can decay only to J = 3/2, due to selection rules. From Figs. 3 and 4, it is clear that the FWM efficiency is largest at frequency detunings where the single-resonance and double-resonance optical pumping effects are the weakest.

 figure: Fig. 4

Fig. 4 (The meanings of the dots, squares and lines in (b) and (d) are the same as in Fig. 3.) The beam powers are P1 = 7.4 mW, P1′ = 11.6 mW, P2 = 38 mW. (a) and (b) The ground state used is hf = 2 of 85Rb, 5S1/2, with |i> = 5P3/2 and |e> = 5D3/2. (c) and (d) The ground state hf = 3 of 85Rb, 5S1/2 is used with |i> = 5P3/2, but with |e> = 5D5/2 and ω2 = 775.978 nm, where the hf levels are inverted. Note the change of scale in the intensity axes. In (a) and (c), the FWM transitions with the least number of decay channels are shown.

Download Full Size | PDF

3.3. Variations in the FWM signal efficiency

At high beam powers for the lower transitions, saturation effects begin to occur, reducing the FWM efficiency. The associated power-broadening effects also contribute to the reduction of the maximum signal intensity. The minima in the signal intensity towards the red-detuned region deserves some attention. The decrease is the largest when the powers of E1 and E1 are large. For instance, as shown in Fig. 3(a) for the conditions of P1 = 7.4 mW, P1′ = 11.6 mW, P2 = 38 mW, the percent decrease is 1030% (480%) for the strong (weak) peak of the FWMsignal convolution. When the beam powers are changed to P1 = 3 mW and P1′ = 4 mW, the percent decrease is only 230% (180%) for the strong (weak) peak (Fig. 3(b)). These values remained constant as P2 was changed from 20 mW to 60 mW. We note that the hf levels 2 and 3 of the intermediate state lie in the red detuned region of the Doppler-broadened linewidth, as can be observed from saturation absorbtion spectroscopy. At high powers of E1 and E1, these hf levels are power-broadened and there is a significant overlap between them. When Δ1 lies in this overlapped region, the contributions of these hf levels to the total two-photon transition amplitude becomes comparable in strength but with opposite signs. The sign of the individual phases has contributions from the signs of the dispersions due to the opposite detunings. Such destructive interference due to multiple intermediate states [2123] causes the total transition amplitude to decrease, suppressing the FWM efficiency. Moreover, in the high intensity regime of the ground-state coupling beams driving to f’ = 2 or 3, the atomic population gets optically pumped out of the system to the f = 2 ground state as shown in Figs. 1(c)-1(d), leading to reduced signal generation.

3.4. Dual role of the driving beam E1

Finally, we note the dual role of the coupling beam E1 in this configuration. As a stimulant to the FWM process, it gains a photon whenever a photon is generated in the FWM signal, as shown in Fig. 1(a). However, because it has access to the ground state population, E1 also contributes to the depletion of the ground state population into incoherent channels, which becomes especially important at large beam powers. Also, because both E1 and E1 haveaccess to the ground state population, the coherences induced by these two beams between |g> and |i> are both significant. A mismatch between the strengths of these two coherences is detrimental to the FWM efficiency [24, 25]. As a result, increasing P1′ indefinitely does not help the FWM efficiency. In fact, increasing the strength of the coherence due to E1 beyond the coherence due to E1 begins to extinguish the FWM intensity. The rates of initial growth and subsequent extinction of the FWM radiation with increasing P1′ is different at different values of Δ1, and is shown in Fig. 5 . This behavior distinguishes this FWM process from the traditional cascade FWM configuration in which the signal is generated in the lower transition with two coupling beams in the upper transition and one probe beam in the lower transition. There, as the coupling beam intensity is increased, the FWM intensity grows until it reaches a maximum value where it remains constant, and signal extinction does not occur.

 figure: Fig. 5

Fig. 5 Dependence of the FWM signal strength on the power P1′ at three different values of Δ1 within the Doppler-broadened D2 absorption linewidth. The powers of the other two beams are held fixed at P1 = 3 mW and P2 = 22 mW. The three chosen values of Δ1, also shown in the inset, correspond to where (i) signal maxima occurs at the blue detuned region (blue dots), (ii) signal maxima occurs at the red detuned region (red triangles), and (iii) signal minima occurs towards the center-red detuned region (black squares). The EIT peak visible in the inset corresponds to case (iii). The three signal trends are connected by lines to aid the eye.

Download Full Size | PDF

4. Summary

The vacuum mode between the upper excited states in a ladder-type configuration is parametrically amplified using atomic coherence mechanisms to enhance the third-order nonlinear response, and studied in the frequency domain. The generated radiation is background free, and its Doppler-free spectral waveform contains high-resolution information about the spectroscopic properties of the atomic energy levels. The line-shape, linewidth and intensity of the generated FWM radiation are determined by various factors such as the beam powers and associated power-broadening effects, the multilevel nature of the atoms and selection rules, the frequency detunings, the destructive interference effects due to contributions by multiple intermediate states, and the dual role of one of the coupling beams. The new radiation could find use in FWM-based applications, and the method can be used to improve procedures using two-photon fluorescence that typically have weak signals and large background noises.

Acknowledgments

We acknowledge partial funding support from the National Science Foundation.

References and links

1. S. E. Harris, “Electromagnetically Induced Transparency,” Phys. Today 50(7), 36–42 (1997). [CrossRef]  

2. E. Arimondo, Progress in Optics, E. Wolf, ed. (Elsevier Science, Amsterdam, 1996), 257–354.

3. M. Xiao, Y. Li, S. Jin, and J. Gea-Banacloche, “Measurement of dispersive properties of electromagnetically induced transparency in rubidium atoms,” Phys. Rev. Lett. 74(5), 666–669 (1995). [CrossRef]   [PubMed]  

4. S. E. Harris, J. E. Field, and A. Imamoglu, “Nonlinear optical processes using electromagnetically induced transparency,” Phys. Rev. Lett. 64(10), 1107–1110 (1990). [CrossRef]   [PubMed]  

5. P. R. Hemmer, D. P. Katz, J. Donoghue, M. Cronin-Golomb, M. S. Shahriar, and P. Kumar, “Efficient low-intensity optical phase conjugation based on coherent population trapping in sodium,” Opt. Lett. 20(9), 982–984 (1995). [CrossRef]   [PubMed]  

6. 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]  

7. A. S. Zibrov, M. D. Lukin, and M. O. Scully, “Nondegenerate parametric self-oscillation via multiwave mixing in coherent atomic media,” Phys. Rev. Lett. 83(20), 4049–4052 (1999). [CrossRef]  

8. Y. Zhang, U. Khadka, B. Anderson, and M. Xiao, “Controlling four-wave and six-wave mixing processes in multilevel atomic systems,” Appl. Phys. Lett. 91(22), 221108 (2007). [CrossRef]  

9. J. Wen, S. Du, Y. Zhang, M. Xiao, and M. H. Rubin, “Nonclassical light generation via a four-level inverted-Y system,” Phys. Rev. A 77(3), 033816 (2008). [CrossRef]  

10. V. Boyer, A. M. Marino, R. C. Pooser, and P. D. Lett, “Entangled images from four-wave mixing,” Science 321(5888), 544–547 (2008). [CrossRef]   [PubMed]  

11. Y. Zhang, U. Khadka, B. Anderson, and M. Xiao, “Temporal and spatial interference between four-wave mixing and six-wave mixing channels,” Phys. Rev. Lett. 102(1), 013601 (2009). [CrossRef]   [PubMed]  

12. U. Khadka, H. Zheng, and M. Xiao, “Interferometric control of parametrically amplified waveforms,” Phys. Rev. A 84(4), 043814 (2011). [CrossRef]  

13. E. F. McCormack and E. Sarajlic, “Polarization effects in quantum coherences probed by two-color, resonant four-wave mixing in the time domain,” Phys. Rev. A 63(2), 023406 (2001). [CrossRef]  

14. S. Williams, E. A. Rohlfing, L. A. Rahn, and R. Zare, “Two-color resonant four-wave mixing: Analytical expressions for signal intensity,” J. Chem. Phys. 106(8), 3090–3102 (1997). [CrossRef]  

15. H. S. Moon, W. K. Lee, L. Lee, and J. B. Kim, “Double resonance optical pumping spectrum and its application for frequency stabilization of a laser diode,” Appl. Phys. Lett. 85(18), 3965–3967 (2004). [CrossRef]  

16. J. Gea-Banacloche, Y. Li, S. Jin, and M. Xiao, “Electromagnetically induced transparency in ladder-type inhomogeneously broadened media: Theory and experiment,” Phys. Rev. A 51(1), 576–584 (1995). [CrossRef]   [PubMed]  

17. R. R. Moseley, S. Shepherd, D. J. Fulton, B. D. Sinclair, and M. H. Dunn, “Two-photon effects in continuous-wave electromagnetically- induced transparency,” Opt. Commun. 119(1-2), 61–68 (1995). [CrossRef]  

18. 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]  

19. H. R. Noh and H. S. Moon, “Diagrammatic analysis of multiphoton processes in a ladder-type three-level atomic system,” Phys. Rev. A 84(5), 053827 (2011). [CrossRef]  

20. R. Gupta, S. Chang, C. Tai, and W. Happer, “Cascade radio-frequency spectroscopy of excited S and D states of rubidium; Anomalous D-state hyperfine structure,” Phys. Rev. Lett. 29(11), 695–698 (1972). [CrossRef]  

21. J. E. Bjorkholm and P. F. Liao, “Resonant enhancement of two-photon absorption in sodium vapor,” Phys. Rev. Lett. 33(3), 128–131 (1974). [CrossRef]  

22. B. Chatel, J. Degert, S. Stock, and B. Girard, “Competition between sequential and direct paths in a two-photon transition,” Phys. Rev. A 68(4), 041402 (2003). [CrossRef]  

23. M. C. Stowe, A. Pe’er, and J. Ye, “Control of four-level quantum coherence via discrete spectral shaping of an optical frequency comb,” Phys. Rev. Lett. 100(20), 203001 (2008). [CrossRef]   [PubMed]  

24. B. Lü, W. H. Burkett, and M. Xiao, “Nondegenerate four-wave mixing in a double-Lambda system under the influence of coherent population trapping,” Opt. Lett. 23(10), 804–806 (1998). [CrossRef]   [PubMed]  

25. H. Kang, G. Hernandez, and Y. Zhu, “Resonant four-wave mixing with slow light,” Phys. Rev. A 70(6), 061804 (2004). [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 (5)

Fig. 1
Fig. 1 (a) Simplified three-level ladder-type configuration in 85Rb that is coherently driven in the FWM process; |g>, |i> and |e> stand for the ground, intermediate and excited states, respectively. E1, E1 and E2 are external driving beams from laser sources, while Ef is the atom-radiated FWM signal that is parametrically amplified from the vacuum mode. (b) Schematic of the experimental configuration showing the directions and polarizations of the four fields. (c)-(f) Realistic energy level diagram showing the hyperfine (hf) levels of each driven state, as well as the incoherent decay channels (wavy arrows) between various combinations of driven hf levels: [f, f′, f′′] = (c) [3,2,2], (d) [3,3,4], (e) [3,4,3], and (f) [3,4,4]. The single-resonance decay channels are drawn first, followed by the double-resonance decay channels. The number of single-resonance and double-resonance optical pumping channels is different in the various cases. The spacings between the energy levels are not drawn to scale.
Fig. 2
Fig. 2 FWM signal line shape, linewidth and efficiency at four different intermediate frequency detunings, placed at intervals of 500 MHz. All other experimental parameters are constant in the four cases. The Doppler-broadened absorption linewidth of the corresponding ground state (hf = 3 of 85Rb, 5S1/2) is also shown for reference.
Fig. 3
Fig. 3 (Color online) TPR FWM signals corresponding to different values of Δ1, separated by 100 MHz each. The energy levels driven are |g> = 5S1/2, hf = 3, |i> = 5P3/2 and |e> = 5D3/2. At each value of the frequency detuning, the FWM signal is a convolution of a sharp, strong peak and a broad, weak peak, the maximum intensities of which are denoted by a blue square and a red dot, respectively. Each of the two signal peak trends are connected by lines to aid the eye. The two peak values corresponding to a given convolution at a given frequency detuning are connected by a solid black line. The horizontal (vertical) dotted lines identify the intensities (intermediate frequency detunings) of the maximas and minimas of the signal convolution’s two peaks. (a) P1 = 7.4 mW, P1′ = 11.6 mW, P2 = 38 mW (b) P1 = 3 mW, P1′ = 4 mW, P2 = 55 mW. Note the change of scale in the intensity axes.
Fig. 4
Fig. 4 (The meanings of the dots, squares and lines in (b) and (d) are the same as in Fig. 3.) The beam powers are P1 = 7.4 mW, P1′ = 11.6 mW, P2 = 38 mW. (a) and (b) The ground state used is hf = 2 of 85Rb, 5S1/2, with |i> = 5P3/2 and |e> = 5D3/2. (c) and (d) The ground state hf = 3 of 85Rb, 5S1/2 is used with |i> = 5P3/2, but with |e> = 5D5/2 and ω2 = 775.978 nm, where the hf levels are inverted. Note the change of scale in the intensity axes. In (a) and (c), the FWM transitions with the least number of decay channels are shown.
Fig. 5
Fig. 5 Dependence of the FWM signal strength on the power P1′ at three different values of Δ1 within the Doppler-broadened D2 absorption linewidth. The powers of the other two beams are held fixed at P1 = 3 mW and P2 = 22 mW. The three chosen values of Δ1, also shown in the inset, correspond to where (i) signal maxima occurs at the blue detuned region (blue dots), (ii) signal maxima occurs at the red detuned region (red triangles), and (iii) signal minima occurs towards the center-red detuned region (black squares). The EIT peak visible in the inset corresponds to case (iii). The three signal trends are connected by lines to aid the eye.
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.