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

Tuning the phase sensitivity of a double-lambda system with a static magnetic field

Open Access Open Access

Abstract

We study the effect of a DC magnetic field on the phase sensitivity of a double-lambda system coupled by two laser fields, a probe and a pump. It is demonstrated that the gain and the refractive index of the probe can be controlled by either the magnetic field or the relative phase between the two laser fields. More interestingly, when the system reduces to a single-lambda system, turning on the magnetic field transforms the system from a phase-insensitive process to a phase-sensitive one. In the pulsed-probe regime, we observed switching between slow and fast light when the magnetic field or the relative phase was adjusted. Experiments using a coated 87Rb vapor cell produced results in good agreement with our numerical simulation. This work provides a novel and simple means to manipulate phase sensitive electromagnetically-induced-transparency or four-wave mixing, and could be useful for applications in quantum optics, nonlinear optics and magnetometery based on such systems.

© 2013 Optical Society of America

1. Introduction

Quantum interference between different excitation path ways is an intriguing phenomenon in quantum optics, laser and atomic physics. Notable examples are electromagnetically induced transparency (EIT) [13], multi-wave mixing [4] etc. Among these, an interesting category is the phase-sensitive process formed by closed-loop interactions, where the phases of the optical fields can dramatically change the steady state of the atoms and the optical susceptibility. There has been a large amount of work in this field since the 1980s [5, 6], and the interests boosted after the experimental demonstration of phase-sensitive EIT [7]. Applications of phase sensitive processes include large nonlinearity [810], slow and fast light [1113], and optical switch [14] etc. To the best of our knowledge, it is a tradition to use oscillating electromagnetic fields, i.e., either all optical fields [7, 1518], or a combination of microwave fields and optical fields [1921], to form a closed loop.

In this paper, we propose to use a static magnetic field to close the interaction loop. This approach is suitable as long as the ground states can be coupled by the magnetic field. For example, ground states formed by superpositions of Zeeman sublevels can be mixed if their spin orientations are not parallel to the external magnetic field. In this case, the magnetic field can in effect coherently couple the two ground states through Larmor precession. For instance, as we will show below, in a lambda system, without the magnetic field the system is a normal EIT configuration, whose steady state susceptibility is not influenced by the optical phases; with the magnetic field, a closed loop is formed and the field absorption becomes dependent on the relative phase. Similar idea also applies to a double-lambda system. Although its level configuration naturally allows a closed-cycle four-wave-mixing (FWM) which is already phase sensitive, applying a magnetic field can still alter the degree of phase sensitivity, as also shown below both theoretically and experimentally.

Adding a magnetic field as a new knob to the double-lambda system with Zeeman sub-level ground states might be particularly useful, since such a system has attracted considerable amount of interests recently for use in polarization self-rotation [22], light squeezing [23, 24], light entanglement [25] and amplified slow and stored light [26]. The following advantages of this double-lambda system are probably responsible for its popularity: (1) It corresponds to the energy configuration of the D1 line of the alkali atoms, and can be easily implemented by experiments. (2) Its relevance to magnetic fields allows various magnetometer schemes to be realized [27, 28]. (3) Laser fields with only one frequency are sufficient to address this system due to the Zeeman ground states, which greatly simplifies experiments. (4) The long lifetime of the ground states enables coherence-enhanced nonlinear optics [3].

This paper is organized as follows. In section 2, we introduce the system theoretically, and lay out key steps in our numerical simulation. Then in section 3, we describe our experimental setup and measurement methods. In section 4, we present the experiment results along with corresponding numerical results, which mainly include: (i) The presence of a magnetic field increases the dependence of the optical absorption on the relative phase between laser fields; (ii) When the magnetic field reverses and the relative phase changes by π, the optical responses remain the same; (iii) In the dynamical regime, the probe pulse can experience either slow light or fast light depending on the magnetic field and the relative phase. Finally in section 5, we conclude.

2. Theory

The double-lambda system under consideration is shown in Fig. 1, where (a) and (b) use different ground state basis. |3〉 and |4〉 are degenerate Zeeman sublevels (with opposite magnetic quantum numbers) and are eigenstates of the atomic angular momentum component along the direction of the static magnetic field. |1〉 and |2〉 are the two hyperfine excited states separated by Δ. The optical fields are nearly resonant with the lower excited state. We assume that the selection rules and Clebsch-Gordan coefficients determine that a sigma plus (minus) field E1 (E2) can couple |3〉 (|4〉) to both |1〉 and |2〉, with Rabi frequency Ω1 (−Ω2) and Ω12) respectively. Since in the experiment we use two orthogonally linearly polarized laser fields, a weak probe Ep and a strong control Ec, it is more convenient to use the linear basis in Fig. 1(b), where |X=(|3+|4)/2, and |Y=(|3|4)/2. In this basis, the probe field couples |Y〉 → |2〉 and |X〉 → |1〉 transitions with the same Rabi frequency Ωp=(Ω1Ω2)/2, and the control field couples |X〉 → |2〉 and |Y〉 → |1〉 transitions with the same Rabi frequency Ωc=(Ω1+Ω2)/2. When a magnetic field along the light propagation direction is applied, |3〉 and |4〉 have zeeman shift δB/2, − δB/2 respectively; and in the linear basis, |X〉 and |Y〉 are coherently coupled by the magnetic field with an effective Rabi frequency equal to δB. Here, δB = BmB/h with μB the Bohr magneton, g the Landé g factor, and m the magnetic quantum number. For 87Rb, we have δB = 1.4 MHz for B = 1 Gauss.

 figure: Fig. 1

Fig. 1 Level configurations for a double-lambda system coupled by two orthogonally polarized laser fields resonant with the lower excited state, represented (a) in the circular basis where the two circularly polarized fields are E1 and E2, and (b) in the linear basis where the two linearly polarized fields are Ec and Ep. Here, |X=(|3+|4)/2, |Y=(|3|4)/2. Due to the selection rules and Clebsch-Gordan coefficients of Rb, E1 couples |3〉 → |2〉, |1〉 both with Rabi frequency Ω1, and E2 couples |4〉 → |2〉, |1〉 with Rabi frequency Ω2, −Ω2 respectively. Similarly, Ep couples |Y〉 → |2〉, |X〉 → |1〉 both with Rabi frequency Ωp, and Ec couples |X〉 → |2〉, |Y〉 → |1〉 both with Rabi frequency Ωc.

Download Full Size | PDF

The Hamiltonian of the system takes the following form in the circular basis [Fig. 1(a)]:

H^=2πδB(|33||44|)/2+Δ|11|+Ω1|23|+Ω2|24|+Ω1|13|Ω2|14|+H.c..
Here, Ω1 =32E1/ and Ω2 =42E2/, withij the transition dipole moment between level i and j. The two circularly polarized laser fields E1 and E2 are connected to the linearly polarized fields Ep and Ec by
(E1E2)=12(1i1i)(EcEp)
Various decay channels can be taken into account by the master equation
ρ^˙=1ih¯[H^,ρ^]+Γ^excΓ^rel
where Γ̂exc and Γ̂rel are excitation matrix and relaxation matrix. We assume that both excited states have a decay rate Γ, and in the circular basis, the ground states population difference decay rate is γ1 and the coherence decay rate is γ2. In our model, when cold atoms are considered, we set Γ/2π = 6 MHz, and when a warm atomic vapor is considered, we set Γ/2π = 500 MHz to take into account of Doppler broadening. We consider the steady state of the atoms by setting ρ^˙=0, and then the propagation of the two circular polarized fields are described by the Maxwell equation in the slowly varying amplitude approximation:
Ω1z=iκ(ρ23+ρ13)
Ω2z=iκ(ρ24ρ14)
where κ = νN℘2/2ε0ch̄ is the coupling constant, with ν the laser frequency, the dipole moment of the transitions |2〉 → |3〉 and |2〉 → |4〉 (assumed to be equal), and N the atomic density. Numerically, we can solve for the fields’ complex amplitudes Ω1(L) and Ω2(L) at the vapor cell output (L the cell length), and then have the probe field’s output power as:
Wp|Ωp(L)|2=(|Ω1(L)|2+|Ω2(L)|22|Ω1(L)||Ω2(L)|×cosθ)/2
where θ is the relative phase between Ω1(L) and Ω2(L). It can be seen that the probe field’s transmission is related to both the amplitudes and relative phase of Ω1 and Ω2 at the output.

3. Experimental setup

We performed experiments (schematically shown in Fig. 2) using a cylindrical paraffin coated vapor cell containing isotropically enriched 87Rb. The cell (2.5 cm diameter, 7.5 cm length) was heated by a blown-air oven residing in a solenoid for magnetic field control, all within a four-layer magnetic shield. Tuned close to the 87Rb D1 line F = 2 → F′ = 1 transition, the laser passed through a polarized beam splitter (PBS) to generate the linearly polarized pump and the probe field. Then they each went through an acoustic-optical modulator (AOM) for independent power control and pulse shaping. The two identical AOMs shifted the laser frequency by 80 MHz, and the two 80 MHz RF sources were phase locked to ensure the nearly perfect phase coherence between the probe and the control. The first order beams from the AOMs were then combined by another PBS and directed into the vapor cell. At the cell output, a high quality PBS separated the probe and the pump, and they were independently detected by two photo-detectors (PD) with tunable gains. 30 dB gain was used for the probe detection in all the data presented below.

 figure: Fig. 2

Fig. 2 Schematics of the experimental setup. AOM: acoustic-optical modulator, PZT: piezoelectric transducer, PBS: polarization beam splitter, PD: photo-detector.

Download Full Size | PDF

One of the key elements of this experiment was to have good control on the relative phase between the control and the probe, which was tuned by varying the voltage on the piezoelectric transducer (PZT) attached to the mount of a mirror right before the combining PBS. To reduce fluctuations of the relative phase, the splitting and combining optics were covered by a box to minimize air flow. The maximal phase change the PZT could provide was about 4π. To measure the relative phase at the cell input, we used a flip mirror to direct the beam to a half wave plate followed by a PBS to interfere the control and the probe. Also, it was ensured that this calibration optics at the side gave the same phase with a temporary calibration optics at the cell output (added a half wave plate before the detection PBS) while the entire shield was pushed aside. To check the phase stability, the interference signal was taken before and after each absorption measurement for a particular phase.

In all the experiments reported below, the control field and probe field power were about 225 μW and 15 μW respectively (unless otherwise stated), and both of them were 3.5 mm in diameter. We used a single mode fiber at the laser output to ensure good beam profile. The overlap of the two beams after combination were checked by a beam profiler at several locations along the light stream. The temperature of the cell was about 55 °C, corresponding to an atomic density of about 2 × 1011 cm−3 and an optical depth about 15 [29].

4. Results and discussions

4.1. Effect of the magnetic field on phase sensitivity

In this section, we discuss the effect of the magnetic field on the phase sensitivity of the double-lambda system as shown in Fig. 1(b).

Let’s start from a simpler case which we call the cold atom regime, where the spacing (∼ 800 MHz) between the two excited states is much larger than the linewidth (∼ 6 MHz) of the excited states. Figure 3 shows the calculated probe transmission (normalized to the input value) vs the relative phase between the probe and control, for three magnetic field values (represented by δB). Since these curves have a period of 2π, only one period is plotted. It can be seen that, when the magnetic field is off, the probe transmission is insensitive to the phase. This is because the negligible effect of the upper excited state effectively reduces the system to a normal lambda system, which is not a closed loop without the magnetic field. When the magnetic field is on, the loop is formed. Indeed, when δB increases to 10 Hz and then to 80 Hz, the transmission becomes increasingly sensitive to the phase.

 figure: Fig. 3

Fig. 3 Calculated probe transmission (normalized to the input) vs the relative phase between the probe and the pump for different δB in the cold atom regime, where the excited state decay rate Γ/2π = 6 MHz. (a) The black dashed, grey solid and black solid curves are for δB = 0 Hz, 10 Hz and 80 Hz respectively. (b) The black dashed, grey solid and black solid curves are for δB = 0 Hz, −10 Hz and −80 Hz respectively. Simulation parameters: Ωc/2π = 20 kHz, Ωp/2π = 5 kHz, γ1/2π = γ2/2π = 10 Hz, and the optical depth was 0.15. Changes in the optical depth do not affect the main features of the curves as described in the text.

Download Full Size | PDF

Figure 3(b) shows the calculated probe transmission for magnetic fields opposite to that in Fig. 3(a). It was found that two curves with opposite δB values coincide if we translate one of them by π on the x-axis. This phenomenon can be explained using the linear basis [Fig. 1(b)]. In each lambda loop, for example, in the cycle of |2〉 → |X〉 → |Y〉 → |2〉, the coherence term ρ2Y is determined by the product of Ωc and δB which remains unchanged if the sign of δB and Ωc reverses simultaneously. Indeed, we found that the master equation remains valid when the following substitution is made: δB → −δB, Ωp → Ωp, Ωc → −Ωc, ρ1Xρ1X, ρ2Yρ2Y, ρ1Y → −ρ1Y, ρ2X → −ρ2X, ρXY → −ρXY, ρaaρaa (a = 1, 2, X, Y), which indicates these are two sets of identical solutions to the system. In other words, the probe field susceptibility should be the same if we change the sign of the magnetic field and change the relative phase from φ to φ + π simultaneously.

To verify above predictions, we measured the probe transmission vs the relative phase for δB = 0, ±40 Hz [Fig. 4(a)]. Within 50 ms, we swept the relative phase from 0 to 2π and the data shown in Fig. 4(a) were averaged for about 30 times. The transmission was normalized to the off-resonant transmission of the probe when the control was absent. We can see that, in contrast to the cold atom regime, the phase sensitivity was already very pronounced without the magnetic field. This is because that Doppler broadening makes the effects of the upper excited state not negligible, and thus the fields form a closed loop four-wave-mixing (even without the magnetic field) which is phase sensitive. When δB was nonzero, the phase sensitivity increased, as in the simulation for cold atoms. Also, the two curves for δB = ±40 Hz can almost overlap when one curve is shifted by π along the x-axis, consistent with above theoretical analysis. To account for the shape difference of the measured curves compared to the cold atom case, we performed simulations with Γ/2π = 500 MHz [Fig. 4(b)] and found good agreement with our experiment. The experiment results show that a double-lambda scheme has distinct phase dependence than a single lambda scheme.

 figure: Fig. 4

Fig. 4 (a) Measured output probe transmission vs the relative phase for different δB. The black dashed, black solid and grey solid curves are for δB = 0 Hz, 40 Hz and −40 Hz respectively. Experimental conditions are in section 3. (b) Corresponding theoretical results where the excited state decay rate Γ/2π = 500 MHz. Other simulation parameters: γ1/2π = 25 Hz, γ2/2π = 28 Hz, Ωc/2π = 240 kHz, Ωp/2π = 60 kHz, and the optical depth was 15.

Download Full Size | PDF

4.2. System manipulation via combination of relative phase and magnetic field

In this section, we investigate the combined effects of the magnetic field and the relative phase on the probe’s transmission and its group velocity.

We first measured the probe transmission vs δB for various relative phase φ. The magnetic field sweep in Fig. 5(a) took about 5 seconds and the curve shown was averaged for about 50 times. As shown in Fig. 5(a), the curves combine features of absorption and dispersion plots, which can be interpreted using the circular basis. Equation (6) shows that the probe transmission carries the information of both the absorption and refractive index (phase) of the circular laser beams. In particular, the first two terms of Eq. (6) represent the total transmission of the two circular fields, and should give a normal Lorentzian EIT spectra when δB is swept. As for the interference term of Eq. (6), cosθ can be written as cosθ = cos(θ0 + ε) = cosθ0ε sin(θ0) if ε is small. Here θ0 is the input relative phase between the two circular beams, and ε is the accumulated phase difference which has a dispersive feature. Similar lineshapes was found in other systems, such as the NCPT resonance system [30, 31] where the probe transmission also contains contribution from a Lorentzian and a dispersion function.

 figure: Fig. 5

Fig. 5 Probe transmission vs δB for several relative phases. (a) Measured results, with experimental conditions stated in section 3. (b) Calculated results, with identical simulation parameters as used in Fig. 4(b).

Download Full Size | PDF

Also, it is clearly seen from Fig. 5 that, the two curves with relative phase φ and φ + π are mirror images of each other. This again verifies that the probe transmission remains unchanged if we change the sign of δB and add π to the relative phase simultaneously. Figure 5(b) is the numerically calculated results, which agree well with the experiment. Remaining discrepancy between the experiment and the simulation results is mainly due to that in our simplified model, we did not include the effects of atomic motion in the coated vapor cell [3234], and Doppler broadening was not strictly taken into account by integrating over all velocities.

Next, we studied the dynamic behavior of the probe field, and found that the refractive index and the group velocity can be also controlled by the magnetic field and the relative phase. The probe pulse was from the AOM’s first-order beam. The driving field of the AOM was the output of a mixer which mixes a continuous wave 80 MHz RF signal and a pulse sequence generated by a function generator controlled by LabVIEW. Figure 6 shows the experiment data with 30 times average. For optimal slow and fast light effects, the full width half maximum (FWHM) of the probe pulse was chosen to be about 5 ms. The continuous control field power was 225 μW and the peak power of the probe pulse was 15 μW. When δB was set at −91 Hz, we could turn the slow light into fast light by changing the relative phase from about 1.75 π to 0 [Fig. 6(a)]. The base of the fast light or slow light curve was the result of the optical rotation of the control field due to the magnetic field. On the other hand, if the relative phase was set to be 0, we could also turn the slow light into fast light by changing δB from 0 Hz to −91 Hz [Fig. 6(b)]. The fractional delay and advance were about 25% and 10% respectively.

 figure: Fig. 6

Fig. 6 Measured slow and fast light results. Probe pulse transmission is normalized to the reference (black solid curve) pulse. The blue dashed curve is the fast light, and the red dashdotted curve is the slow light. (a) δB = −91 Hz. (b) relative phase φ = 0.

Download Full Size | PDF

To explain such phenomena, we calculated the refractive index experienced by the probe field while fixing the control field frequency and sweeping the probe’s one-photon detuning δ. For this calculation, we rewrote the Hamiltonian of the system in the linear basis as depicted in Fig. 1(b), and computed the real part of (ρ1X + ρ2Y)/Ep which is proportional to the refractive index for the probe. According to the group velocity formula

vg=cn+ωdndω,
the slope of the refractive index spectra of the probe near δ = 0 Hz in Fig. 7 determines the group velocity of the probe pulse. Figure 7(a) shows that when δB = −91 Hz, the slope near δ = 0 changes from positive to negative when the phase changes from 1.75 π to 0. Figure 7(b) shows that when the relative phase φ = 0, the slope near δ = 0 changes from positive to negative when δB varies from 0 Hz to −91 Hz. This verifies that switching between slow and fast light can be realized by either adjusting the magnetic field or the relative phase.

 figure: Fig. 7

Fig. 7 Calculated spectrum of the real part of (ρ1X +ρ2Y)/Ep vs the probe frequency, with the pump frequency fixed. (a) Relative phase manipulation for δB = −91 Hz. (b) Magnetic field manipulation for the relative phase φ = 0. Simulation parameters are the same as in Fig. 4(b).

Download Full Size | PDF

5. Conclusion

In conclusion, we demonstrated a new method to manipulate the phase sensitivity of a double-lambda system. A magnetic field coupling the two ground states can function as an effective oscillating electromagnetic field, and thus can either close an otherwise open interaction loop, rendering the system phase-sensitive, or alter the phase sensitivity of an initially phase-dependent system. We have theoretically and experimentally verified that a static magnetic field can drastically influence the absorption and refractive index of an optical field and its dependence on the relative phase. This work should be useful for group velocity manipulation, nonlinear optics and magnetometery.

Acknowledgments

We are grateful to Xiangming Hu and Lei Feng for useful discussions. This work was supported by the NBRPC (973 Program Grants No. 2011CB921604 and No. 2012CB921604), the NNSFC (Grant No. 61078013), and Fudan University.

References and links

1. K. J. Boiler, A. Imamoglu, and S. E. Harris, “Observation of electromagnetically induced transparency,” Phys. Rev. Lett. 66,2593–2596 (1991) [CrossRef]  .

2. M. D. Lukin, “Colloquium: trapping and manipulating photon states in atomic ensembles,” Rev. Mod. Phys. 75,457–472 (2003) [CrossRef]  .

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

4. Y. Zhang and M. Xiao, Multi-Wave Mixing Processes (Springer, Berlin, 2009) [CrossRef]  .

5. S. J. Buckle, S. M. Barnett, P. L. Knigt, M. A. Lauder, and D. T. Pegg, “Atomic interferometers: Phase-dependence in multilevel atomic transitions,” Opt. Acta 33,1129–1140 (1986) [CrossRef]  .

6. S. P. Krinitzky and D. T. Pegg, “Coherent irradiation of multilevel atoms in branched and cyclic configurations,” Phys. Rev. A 33,403–406 (1986) [CrossRef]   [PubMed]  .

7. E. A. Korsunsky, N. Leinfellner, A. Huss, S. Baluschev, and L. Windholz, “Phase-dependent electromagnetically induced transparency,” Phys. Rev. A 59,2302–2305 (1999) [CrossRef]  .

8. M. Sahrai, A. Maleki, R. Hemmati, and M. Mahmoudi, “Transient dispersion and absorption in a V-shaped atomic system,” Eur. Phys. J. D 56,105–112 (2010) [CrossRef]  .

9. T. M. Preethi, M. Manukumara, K. Asha, J. Vijay, D. A. Roshi, and A. Narayanan, “Phase-sensitive microwave optical double resonance in an N system,” Euro. Phys. Lett. 95,34005 (2011) [CrossRef]  .

10. J. Dimitrijević and D. Arsenović, “Simple analytical expressions for the analysis of the phase-dependent electromagnetically induced transparency in a double-Λ atomic scheme,” Phys. Scr. T149, 014007 (2012) [CrossRef]  .

11. D. Bortman-Arbiv, A. D. Wilson-Gordon, and H. Friedmann, “Phase control of group velocity: From subluminal to superluminal light propagation,” Phys. Rev. A 63,043818 (2001) [CrossRef]  .

12. H. Sun, H. Guo, Y. Bai, D. Han, S. Fan, and X. Chen, “Light propagation from subluminal to superluminal in a three-level Lambda-type system,” Phys. Lett. A 335,68–75 (2005) [CrossRef]  .

13. D. Jafari, M. Sahrai, H. Motavalli, and M. Mahmoudi, “Phase control of group velocity in a dielectric slab doped with three-level ladder-type atoms,” Phys. Rev. A 84,063811 (2011) [CrossRef]  .

14. J. Kou, R. Wan, Z. Kang, L. Jiang, L. Wang, Y. Jiang, and J. Gao, “Phase-dependent coherent population trapping and optical switching,” Phys. Rev. A 84,063807 (2011) [CrossRef]  .

15. W. Maichen, F. Renzoni, I. Mazets, E. Korsunsky, and L. Windholz, “Transient coherent population trapping in a closed loop interaction scheme,” Phys. Rev. A 53,3444–3448 (1996) [CrossRef]   [PubMed]  .

16. E. A. Korsunsky and D. V. Kosachiov, “Phase-dependent nonlinear optics with double-Λ atoms,” Phys. Rev. A 60,4996–5009 (1999) [CrossRef]  .

17. X. Fan, Z. Liu, Y. Liang, K. Jia, and D. Tong, “Phase control of probe response in a Doppler-broadened N-type four-level system,” Phys. Rev. A 83,043805 (2011) [CrossRef]  .

18. A. F. Huss, E. A. Korsunsky, and L. Windholz, “Phase control of electromagnetically induced transparency in a double-Λ system,” J. Mod. Opt. 49,141–155 (2002) [CrossRef]  .

19. H. Li, V. A. Sautenkov, Y. V. Rostovtsev, G. R. Welch, P. R. Hemmer, and M. O. Scully, “Electromagnetically induced transparency controlled by a microwave field,” Phys. Rev. A 80,023820 (2009) [CrossRef]  .

20. D. Kosachiov, B. Matisov, and Yu. Rozhdestvensky, “Coherent population trapping: sensitivity of an atomic system to the relative phase of exciting fields,” Opt. Commun. 85,209–212 (1991) [CrossRef]  .

21. M. S. Shahriar and P. R. Hemmer, “Direct excitation of microwave-spin dressed states using a laser-excited resonance Raman interaction,” Phys. Rev. Lett. 65,1865–1868 (1990) [CrossRef]   [PubMed]  .

22. T. Horrom, S. Balik, A. Lezama, M. D. Havey, and E. E. Mikhailov, “Polarization self-rotation in ultracold atomic 87Rb,” Phys. Rev. A 83, 053850 (2011) [CrossRef]  .

23. E. E. Mikhailov and I. Novikova, “Low-frequency vacuum squeezing via polarization self-rotation in Rb vapor,” Opt. Lett. 33,1213–1215 (2008) [CrossRef]   [PubMed]  .

24. S. Barreiro, P. Valente, H. Failache, and A. Lezama, “Polarization squeezing of light by single passage through an atomic vapor,” Phys. Rev. A 84,033851 (2011) [CrossRef]  .

25. W. Wasilewski, T. Fernholz, K. Jensen, L. S. Madsen, H. Krauter, C. Muschik, and E. S. Polzik, “Generation of two-mode squeezed and entangled light in a single temporal and spatial mode,” Opt. Express 17,14444–14457 (2009) [CrossRef]   [PubMed]  .

26. I. Novikova, D. F. Phillips, and R. L. Walsworth, “Slow light with integrated gain and large pulse delay,” Phys. Rev. Lett. 99,173604 (2007) [CrossRef]   [PubMed]  .

27. W. Wasilewski, K. Jensen, H. Krauter, J. J. Renema, M. V. Balabas, and E. S. Polzik, “Quantum noise limited and entanglement-assisted magnetometry,” Phys. Rev. Lett. 104,133601 (2010) [CrossRef]   [PubMed]  .

28. B. Patton, O. O. Versolato, D. C. Hovde, E. Corsini, J. M. Higbie, and D. Budker, “A remotely interrogated all-optical 87Rb magnetometer,” Appl. Phys. Lett. 101,083502 (2012) [CrossRef]  .

29. N. B. Phillips, A. V. Gorshkov, and I. Novikova, “Light storage in an optically thick atomic ensemble under conditions of electromagnetically induced transparency and four-wave mixing,” Phys. Rev. A 83,063823 (2011) [CrossRef]  .

30. M. Crescimanno and M. Hohensee, “N+CPT clock resonance,” J. Opt. Soc. Am. B 25,2130–2139 (2008) [CrossRef]  .

31. C. Hancox, M. Hohensee, M. Crescimanno, D. F. Phillips, and R. L. Walsworth, “Lineshape asymmetry for joint coherent population trapping and three-photon N resonances,” Opt. Lett. 33,1536–1538 (2008) [CrossRef]   [PubMed]  .

32. M. Klein, M. Hohensee, D. F. Phillips, and R. L. Walsworth, “Electromagnetically induced transparency in paraffin-coated vapor cells,” Phys. Rev. A 83,013826 (2011) [CrossRef]  .

33. I. Novikova, R. Walsworth, and Y. Xiao, “Electromagnetically induced transparency-based slow and stored light in warm atoms,” Laser Photonics Rev. 6,333–353 (2012) [CrossRef]  .

34. Z. Xu, W. Qu, R. Gao, X. Hu, and Y. Xiao, “Linewidth of electromagnetically induced transparency under motional averaging in a coated vapor cell,” Chin. Phys. B 22,033202 (2013) [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 (7)

Fig. 1
Fig. 1 Level configurations for a double-lambda system coupled by two orthogonally polarized laser fields resonant with the lower excited state, represented (a) in the circular basis where the two circularly polarized fields are E1 and E2, and (b) in the linear basis where the two linearly polarized fields are Ec and Ep. Here, | X = ( | 3 + | 4 ) / 2, | Y = ( | 3 | 4 ) / 2. Due to the selection rules and Clebsch-Gordan coefficients of Rb, E1 couples |3〉 → |2〉, |1〉 both with Rabi frequency Ω1, and E2 couples |4〉 → |2〉, |1〉 with Rabi frequency Ω2, −Ω2 respectively. Similarly, Ep couples |Y〉 → |2〉, |X〉 → |1〉 both with Rabi frequency Ωp, and Ec couples |X〉 → |2〉, |Y〉 → |1〉 both with Rabi frequency Ωc.
Fig. 2
Fig. 2 Schematics of the experimental setup. AOM: acoustic-optical modulator, PZT: piezoelectric transducer, PBS: polarization beam splitter, PD: photo-detector.
Fig. 3
Fig. 3 Calculated probe transmission (normalized to the input) vs the relative phase between the probe and the pump for different δB in the cold atom regime, where the excited state decay rate Γ/2π = 6 MHz. (a) The black dashed, grey solid and black solid curves are for δB = 0 Hz, 10 Hz and 80 Hz respectively. (b) The black dashed, grey solid and black solid curves are for δB = 0 Hz, −10 Hz and −80 Hz respectively. Simulation parameters: Ωc/2π = 20 kHz, Ωp/2π = 5 kHz, γ1/2π = γ2/2π = 10 Hz, and the optical depth was 0.15. Changes in the optical depth do not affect the main features of the curves as described in the text.
Fig. 4
Fig. 4 (a) Measured output probe transmission vs the relative phase for different δB. The black dashed, black solid and grey solid curves are for δB = 0 Hz, 40 Hz and −40 Hz respectively. Experimental conditions are in section 3. (b) Corresponding theoretical results where the excited state decay rate Γ/2π = 500 MHz. Other simulation parameters: γ1/2π = 25 Hz, γ2/2π = 28 Hz, Ωc/2π = 240 kHz, Ωp/2π = 60 kHz, and the optical depth was 15.
Fig. 5
Fig. 5 Probe transmission vs δB for several relative phases. (a) Measured results, with experimental conditions stated in section 3. (b) Calculated results, with identical simulation parameters as used in Fig. 4(b).
Fig. 6
Fig. 6 Measured slow and fast light results. Probe pulse transmission is normalized to the reference (black solid curve) pulse. The blue dashed curve is the fast light, and the red dashdotted curve is the slow light. (a) δB = −91 Hz. (b) relative phase φ = 0.
Fig. 7
Fig. 7 Calculated spectrum of the real part of (ρ1X +ρ2Y)/Ep vs the probe frequency, with the pump frequency fixed. (a) Relative phase manipulation for δB = −91 Hz. (b) Magnetic field manipulation for the relative phase φ = 0. Simulation parameters are the same as in Fig. 4(b).

Equations (7)

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

H ^ = 2 π δ B ( | 3 3 | | 4 4 | ) / 2 + Δ | 1 1 | + Ω 1 | 2 3 | + Ω 2 | 2 4 | + Ω 1 | 1 3 | Ω 2 | 1 4 | + H . c ..
( E 1 E 2 ) = 1 2 ( 1 i 1 i ) ( E c E p )
ρ ^ ˙ = 1 i h ¯ [ H ^ , ρ ^ ] + Γ ^ exc Γ ^ rel
Ω 1 z = i κ ( ρ 23 + ρ 13 )
Ω 2 z = i κ ( ρ 24 ρ 14 )
W p | Ω p ( L ) | 2 = ( | Ω 1 ( L ) | 2 + | Ω 2 ( L ) | 2 2 | Ω 1 ( L ) | | Ω 2 ( L ) | × cos θ ) / 2
v g = c n + ω d n d ω ,
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.