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

High resolution electromagnetically induced transparency spectroscopy of Rydberg 87Rb atom in a magnetic field

Open Access Open Access

Abstract

We have studied the high resolution spectroscopy of Rydberg state 87Rb in a ladder-type electromagnetically induced transparency (EIT) configuration at room temperature. A highly excited Rydberg atom is nearly degenerate for its hyperfine states but this degeneracy will be broken by an applied magnetic field, resulting in a spectral splitting in a coupled basis. In our ladder-type EIT experiment, we observed the high resolution spectral splitting of Rydberg atoms in an external magnetic field with two different optical polarization combinations of σ+σ+ and σ+σ for probe and coupling laser beams. A strict theory has been set up to explain the observed spectral line position and intensity accurately considering the Zeeman effects of three EIT-concerned states all in the coupled basis. Specially for the Rydberg state, we can transform its wavefunction back into the decoupled basis for the spectral line assignment.

© 2017 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

The optical phenomena based on quantum interference and atomic coherent effects have attracted a lot of attention in recent years. Electromagnetically induced transparency (EIT) is a quantum coherent effect that a weak probe field will experience increased transmission in an absorbing medium when a strong coupling field is added. It has already led to widespread applications in many fields, for example, slowing of the light [1], lasing without inversion [2, 3], quantum memories [4], precision spectroscopy [5] and magnetometry [6], especially in microwave electrometry using the highly excited Rydberg states [7,8]. Up to now, most of these phenomena have been widely studied in simple three-level schemes, such as Λ–type, V-type and ladder type using atomic vapors [9] or solid state materials [10,11].

Recently, Λ–type EIT technique has been widely used in studying the Zeeman effect of atom [12–14]. In a real atomic system, the atomic states usually have complex structure with Zeeman-splitting sublevels [15]. Studying the complex behavior of EIT resonances in the degenerate atomic system in magnetic fields has attracted great interests both theoretically and experimentally [12, 14, 16–19]. The energy levels are Zeeman shifted with different resonant frequencies and their transition probabilities are extremely sensitive to an external magnetic field [20,21]. For example, Gao et al. observed the EIT window of 87Rb split into three or four narrower subwindows due to the lifting of magnetic sublevels’ degeneracy [9]. Sargsyan et al. have also investigated the EIT of 85Rb and 87Rb in the magnetic field by using a nano-cell, a centimeter-long glass cell and a cell with a neon buffer gas for better spectral resolution [16,17,22,23].

The EIT can also be configured as a ladder type, where the highest energy state of the concerned three levels is a Rydberg state [10, 24–35]. This kind of EIT can explore the fine (hyperfine) structure of Rydberg states, owing to its narrow linewidth and high sensitivity [24,25,36]. In the Λ-type or V-type EIT where the wavelengths of probe and coupling lasers are usually almost the same, the proper choice of co- or counter-propagating beams can reduce the Doppler-broadened vapor to a virtually Doppler free medium. However, for the three-level ladder-type configuration, the coupling and probe wavelengths are usually very different and there exists a wavelength mismatch between them [24, 37, 38], where the multi-intermediate levels lead to a complex spectral structure. Moreover, these hyperfine splittings of the multi-intermediate state are not directly mapped to the observed spectrum, but scaled by a factor of 1 − λcp while the energy level splitting of the upper Rydberg state by λcp when scanning the probe laser [39].

If a magnetic field applied, this wavelength mismatching effect causes the spectral splitting more complex and it is very difficult to clarify the observed spectral lines [40] extensively compared with those in a Λ–type atomic system [34,41–43]. A splitting of EIT window of 133Cs in the cascade three-level system involved the nD or nS Rydberg state with a static magnetic field (several tens of Gauss) has been reported [41,42], where the experimental observation is roughly explained based on a simple linear Zeeman effect for the main spectral line positions. Similar spectral assignment for Rb atom in magnetic field has also been reported [35].

In fact, even in the magnetic field of tens of Gauss, the Zeeman splitting of the hyperfine levels of the high Rydberg atom is much larger than the hyperfine level gap itself and we have to describe the states in a decoupled basis rather than a coupled one [44], making the analysis of the spectrum more difficult, especially for the line intensity calculation. In this paper, we report an experimental spectral observation of Rydberg atoms in an external magnetic field. A theoretical calculation shows that these transitions are closely related to the hyperfine energy level couplings due to the introduction of the magnetic fields, which has not been extensively discussed in the previous experimental studies [35,41,42].

2. Experimental setup

Figure 1(a) shows the energy scheme associated with the three-level ladder-type EIT process of 87Rb atoms. EIT involves a weak probe field with its frequency scanning over transitions from the ground state 5S1/2(F = 2) to the intermediate state 5P3/2(F′ = 1, 2, 3), and an intense coupling field fixed at frequency of transition from the intermediate state 5P3/2(F′ = 3) to the highly excited Rydberg state nS1/2(F″ = 1, 2), where n = 50 in our case.

 figure: Fig. 1

Fig. 1 The energy level scheme (a) and experimental setup (b) of ladder-type EIT system for 87Rb. It should be noted that the hyperfine levels of Rydberg state are nearly degenerate in free fields for a vanishing coupling constant between the nuclear spin I and the total angular momentum J. The 780 nm probe (thin red line) and 480 nm coupling (thick blue line) beams are counter-propagated and overlapped in the vapor cell. PBS: polarizing beam splitter; DM: 45° dichroic mirror; BS: beam splitter; M: mirror; λ/2: half wave plate; λ/4: quarter wave plate; BPD: balanced photodetector.

Download Full Size | PDF

The schematic diagram of our experimental setup is also presented, as shown in Fig. 1(b). A filter-tuned, cateye refletor feedback external cavity diode laser (Moglabs) with linewidth < 1 MHz is used as a probe beam (λp ∼ 780 nm), which scans across the 87Rb 5S1/2 −5P3/2 transitions. While the coupling field with wavelength 480 nm, driving the transition of 5P3/2(F′ = 3) − nS/nD, is provided by a frequency-doubled laser system (TA-SHG Pro, Toptica) for the 960 nm fundamental light. The frequency of coupling field is measured by a wavelength meter (wavemaster) with resolution about 10 MHz.

The 780 nm laser beam splits into two parts by the beam splitter, where one serves as the probe beam, and the other for a reference signal. The probe and coupling lasers are overlapped almost completely and counter-propagated through a rubidium atom vapor cell with length of only 25 mm at room temperature. The small size avoids the inhomegeneity of the magnetic field generated by the coils to the maximum extent. The probe beam (coupling beam) has a waist of about 0.3 mm (1.5 mm) and power 100 μW (300 mW) in our experiments. The reference laser beam also transmits through the same rubidium vapor cell but without coupling laser overlapped, which is used to record only the absorption spectrum with Gaussian-broadening. The transmissions of the weak probe and reference beams are measured by a balanced photodiode (BPD) and a difference signal can be obtained. Thus we can remove the Gaussian-broadening and make the transmission appear on a flat background while scanning the probe laser frequency, which highlights the Rydberg EIT spectral lines. A current-controlled coil surrounding the cell can supply a magnetic field parallel to the laser beams, which is nearly homogenous in center part. The magnetic field is measured by a Gauss meter (Lakeshore 425), and the magnetic field generated by the coil is homogeneous within the area of a L × d = 70 mm ×10 mm cylinder in the center. The homogeneous magnetic filed can be up to 230 Gauss with an error less than 0.3%. Both beams pass through a half-wave plate (λ/2), a polarizing beam splitter (PBS), and a quarter-wave plate (λ/4) before entering the cell. We can configure the probe and coupling laser fields to different polarization combinations such as σ+σ, σ+σ+, where σ+ is defined as the electric field of laser rotating in a left hand sense with respect to the direction of wave traveling while σ in right hand sense.

3. Theoretical model for the atomic transition

The atom in an external magnetic field can be described by the following Hamiltonian [44,45]

H^¯=H^0+H^fs+H^hfs+H^B,
where Ĥ0 is the coarse atomic structure, Ĥfs and Ĥhfs describe the fine and hyperfine interactions, and ĤB represents the atomic interaction with the external magnetic field. Our optically concerned transition occurs between the hyperfine states in magnetic fields so that the first two terms are kept constant during the transition and we can omit them in a relative measurement for the transition frequency. Since the ground state is well described by the good quantum number F and mF, we will express the whole Hamiltonian in the coupled basis |(JI)FmF〉. However, in a very strong magnetic field when both angular momenta (J and I) are totally uncoupled, the most convenient is the uncoupled bases approach where the eigenfunctions of an atomic state can be represented as |JmJ〉|ImI〉. The coupled basis is correlated to the decoupled one by [44]
|(JI)FmF=(1)JI+mF2F+1×mJ,mI(JIFmJmImF)|JmJ|ImI,
where quantities in brackets are 3jm symbols.

Specially, the splittings associated with the hyperfine interaction around the zero-detuning energies can be written diagonally as

ΔEhfs=Ahfs2K+Bhfs432K(K+1)2I(I+1)J(J+1)I(2I1)J(2J1),
where Ahfs is the magnetic dipole constant, Bhfs is the electric quadrupole constant and K = F(F + 1) − I(I + 1) − J(J + 1). The interaction of atom with the applied external magnetic field has the form
HB=μBgJJB+μBgIIB,
where the two terms are expressed in the coupled basis |(JI)FmF〉 as
(JI)FimF|μBgJJB|(JI)FjmF=μBBgJ(1)J+I+Fi+FjmF+1×(2Fi+1)(2Fj+1)J(J+1)(2J+1)×(Fi1FjmF0mF){JFiIFjJ1}
and
(JI)FimF|μBgIIB|(JI)FjmF=μBBgI(1)J+I+Fi+FjmF+1×(2Fi+1)(2Fj+1)I(I+1)(2I+1)×(Fi1FjmF0mF){IFiIFjI1}
respectively. Here quantities in brackets are 3j symbols and in curled brackets 6j symbols. Note non-zero diagonal part is also included in the matrix for i = j. A diagonalization of the full matrix for the ground state, the intermediate state and the excited Rydberg state, can give the transition frequency and transition amplitude for a one-photon absorption.

The transition probability from the initial state to the final one can be written as σ = |ΨΨ′|2, where Ψ and Ψ′ are the eigen-function in the coupled basis for the initial and final states, which has been detailed in the spectral intensity analysis of Rb D2 line at weak/intermediate fields [46]. It should be noted that for the intermediate states in strong magnetic fields or high Rydberg state even in weak fields, we have to take into account the state-coupling [47,48]. The transition probability between pure states from |(JI)FmF〉 to |(JI)F′mF′〉 is given by

σmFmF=FmF|er|FmF=J|er|J(1)2F+J+I+mF(2F+1)(2F+1)(2J+1)×(F1FmFmFmFmF){JJ1FFI}.
Here the part 〈J||er⃗||J′〉 is a constant in the transitions and all energy levels are expressed in the coupled basis. While for the Rydberg state, it is better to describe it in the decoupled basis since mJ and mI are closely good quantum numbers there. We can project the calculated wavefunction of the Zeeman levels of high Rydberg atom back to the decoupled basis via Eq. 2 for spectral assignment in a usual way.

4. Results and discussion

There have been some works on the ladder-type EIT with the coupling laser resonant to the highly excited nD(n > 40) Rydberg state [49,50]. For the highly excited state 85Rb or 87Rb, the hyperfine splittings of the nD are neglectable and only the remarkable fine interaction remains (nD3/2 and nD5/2), thus resulting in six two-photon resonant lines, in combination with the hyperfine splittings of the intermediate state 5P3/2 [49]. For the Rydberg state nS, the spectrum is further reduced to three lines as there is no fine splitting for nS state. However, even for the case of the coupling laser resonant to the highly excited nS1/2 Rydberg states, the spectrum will get very complex if an external magnetic field applied. In this paper, we investigate the EIT spectral splitting of 87Rb atoms excited to 50S1/2 via the intermediate state 5P3/2. In zero field, the high Rydberg state has a neglectable hyperfine splitting, but an applied weak magnetic field can induce a strong interaction between different hyperfine energy levels, resulting in a remarkable spectral splitting.

Unlike the Λ-type EIT [45, 51,52], there is a serious wavelength mismatching effect due to the large discrepancy between the probe and coupling lasers in wavelength. Figure 2 shows the observed absorption spectra of the probe laser as a function of probe frequency shift at different coupling detunings. Owning a nearly degeneracy for the state 50S1/2, two spectral lines are observed, corresponding to the hyperfine-splitting of the intermediate state 5P3/2, namely, 5S1/2, F = 2 → 5P3/2, F′ = 2 → 50S1/2 and 5S1/2, F = 2 → 5P3/2, F′ = 3 → 50S1/2. Due to the wavelength mismatch between the probe and coupling lasers, the hyperfine splitting of the 5P3/2 state is scaled by a factor of 1 − λcp = 0.38, and the new interval between F′ = 2 and F′ = 3 is 0.38 × 267 MHz ≈ 101 MHz, which is in good agreement with the experiment.

 figure: Fig. 2

Fig. 2 Experimentally measured (blue lines) and theoretically calculated (red lines) absorptions of the probe beam as a function of probe beam frequency shift for different coupling detunings. The wavelength mismatching effect has been included in the calculation.

Download Full Size | PDF

As observed in the previous work [53–57], the spectral line amplitude is strongly dependent on the detuning of the coupling laser field. The transmittance spectrum of the 5S1/2(F = 2) → 5P3/2(F′ = 2, 3) → nS transition is a double-resonance optical pumping process. At the coupling detuning Δ = 0 MHz shown in Fig. 2(c), where the coupling laser is resonant to the 5P3/2, F′ = 2 → 50S1/2 transition, this spectral line has a maximum amplitude. When the coupling detuning increases from Δ = −180 MHz up to Δ = 215 MHz as shown in Figs. 2(a)–2(e), with the coupling laser approaching the 5P3/2, F′ = 3 → 50S1/2 resonant transition corresponding to an energy interval of 267 MHz between F′ = 2 and F′ = 3, the associated spectral line gets stronger and stronger significantly while the one with F′ = 2 vanishes. The variable detuning is accomplished by locking the coupling laser to an interferometric frequency control (IScan system, Germany). The maximum detuning value in Fig. 2(e) is about 215 MHz, still less than the energy interval 267 MHz. The theoretical simulation agrees well with the experimental result.

The introduction of magnetic field induces an energy level splitting for 50S1/2, very different from those for the intermediate state 5P3/2 and ground state 5S1/2. The energy diagram is shown in Fig. 3. For the ground state 5S1/2, the hyperfine splitting is about 6.8 GHz, far larger than the interaction induced by the applied magnetic field, which leads to a paramagnetic energy shift for every mF component. For every mF energy level, there is no remarkable interaction between different hyperfine levels and F is approximately a good quantum number. This approximation is also valid for the intermediate state 5P3/2 since its hyperfine splittings are also far larger than the Zeeman energy level shift in our case. The energy gap between F′ = 0 and F′ = 1 is rather small, about 72 MHz, comparable with the applied maximum magnetic field induced interaction, which is estimated to be on the order of 40 MHz at B = 40 Gauss. However, unfortunately, these levels do not participate in our ladder-EIT observation for weak transition probabilities. Therefore, principally, only the splittings for 50S1/2 should be considered seriously. The ground state and the intermediate state can be nearly diagonal in the coupled basis |(JI)FmF〉. While for the 50S1/2 term, the energy gap between hyperfine levels is almost ignorable, only 0.25 MHz [37], which makes even a weak magnetic field introduce a remarkable interaction between the degenerated hyperfine states, where the states are well described and then classified in the decoupled basis |JmJ〉|ImI〉.

 figure: Fig. 3

Fig. 3 Ladder-type EIT system of 87Rb atom in a magnetic field. The probe laser frequency ωp is scanned over the F = 2 to 5P3/2 transition, whereas the frequency ωc is resonant to the transition F′ = 3 → 50S1/2. ω0p and ω0c denote the frequencies of the probe and coupling lasers in zero magnetic field, respectively.

Download Full Size | PDF

We firstly present a spectral observation for the probe and coupling laser configured as σ+ and σ polarized at a magnetic field B = 37 Gauss as shown in Fig. 4. In the measurement, we lock the coupling laser to the resonance of 5P3/2, F′ = 3 to 50S1/2 for a better signal to noise ratio according to the discussion on the effect of laser detuning on the spectral line as shown in Fig. 2. We define ω0p and ω0c to denote the resonant frequency of transition from F = 2 to F′ = 3 and F′ = 3 to 50S1/2 state respectively at zero magnetic field for convenience. As discussed previously, the magnetic field induced interaction breaks the degeneracy of the Rydberg state 50S1/2 and splits the energy level. In this case, F is not a good quantum number any longer and the Zeeman interaction can not be taken as a perturbation. The coupled hyperfine states will mix seriously and the Zeeman energy level of Rydberg state can be expressed in the coupled basis as

Ψ=F|mF|CF|FmF
where CF is the expansion coefficient of the wavefunction Ψ in the coupled basis |FmF〉 while mF is a good quantum number and has a given value. We can project this state back into the decoupled basis |JmJ〉|ImI〉 since the eigen-state is well described by the quantum number mJ and mI. An example of the analysis for the spectral line A and A′ is shown in Fig. 5. All transition assignments are listed in Table 1, where the spectral lines are assigned in the coupled and decoupled basis at the same time for the Rydberg state. For the spectral line A(A′), B(B′) and C(C′), each pair is the result of interaction between different component F in the coupled basis for high Rydberg state, thus the spectral line should be a linear combination of different |FmF〉 as given in Eq. 8. Therefore, we have to assign the spectral lines with probabilities in the coupled basis. The only exception is for the spectral lines D and E where there is no interaction between F for the Rydberg state. As mF is a good quantum number, the magnetic field induced interaction can happen only between the same mF but all values of F satisfying F ≥ |mF|. For the state mF (|mF| = F), there is only one F value, corresponding to the cases of spectral lines D and E with F = 2, mF = −2 and F = 2, mF = 2, respectively. All states can be projected back into decoupled basis |JmJImI〉, which is also labeled in Table 1. In the decoupled basis, a pair of (mJ, mI) is enough for assignment of the spectral line as all other projection components are nearly zero, making little contribution.

 figure: Fig. 4

Fig. 4 The spectra for the probe and coupling laser σ+ and σ polarized in a magnetic field B = 37 Gauss. The red line is the experimental observation while the blue line is the calculated one. In the coupled basis, F is no longer a good quantum number for the Rydberg states except for the lines D and E.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 The assignment of the Rydberg state in the three-level ladder-type EIT configuration for the spectral line A and A′ shown in Fig. 4. In the coupled basis, different F-component of Rydberg state mixes each other, making the coupling laser can also transit from F′ = 3 to F″ = 1 (in dotted olive lines), but they can be described by the good quantum numbers mJ and mI in the decoupled basis |JmJ〉|ImI〉 as shown.

Download Full Size | PDF

Tables Icon

Table 1. Assignment for the spectral lines shown in Fig. 4

We also investigate EIT spectrum via polarization combination σ+ and σ+, respectively for probe and coupling beams at magnetic field B = 40 Gauss. It is shown in Fig. 6. An analysis similar to Fig. 4 is made considering the state mixture. The assignment of the spectral lines is shown in Table 2 where the quantum number in the decoupled basis and the mixing coefficients in the coupled basis are also listed.

 figure: Fig. 6

Fig. 6 The spectra for the probe and coupling laser σ+ and σ+ polarized in a magnetic field B = 40 Gauss. Similar to that shown in Fig. 4, the Rydberg state has to been described by a decoupled basis |JmJ, ImI〉 rather than a coupled basis.

Download Full Size | PDF

Tables Icon

Table 2. Assignment for the spectral lines shown in Fig. 6

It should be noted that the spectra of Rydberg EIT for different polarization combinations of σ+σ+ and σ+σ are completely different even at the same magnetic fields. We calculate these spectra at B = 40 Gauss as shown in Fig. 7, where the coupling laser has the same detuning as that in Fig. 6. In the combination of σ+ and σ, the initial and final states in the Raman process have the same mF while mF is changed by 2 in the combination of σ+ and σ+, corresponding to very different energy levels, therefore resulting in different spectral line positions.

 figure: Fig. 7

Fig. 7 The calculated spectra of Rydberg EIT at B = 40 Gauss for the polarization combinations of σ+σ and σ+σ+. They have completely different spectral patterns.

Download Full Size | PDF

The dependence of Rydberg EIT spectrum on the applied magnetic fields is also investigated as shown in Fig. 8. The polarizations of the probe and coupling fields are configured as σ+ and σ+, respectively. Figures 8(a)–8(b) are the normalized experimental and calculated spectra, respectively. To clarify the spectral lines clearly, a narrow line broadening (3.5 MHz) is used for the calculated transitions. In the experiment, a high power of the coupling laser has to be used to record a Rydberg EIT spectrum with high signal-to-noise ratio, but at the same time it causes the broadening of the peaks (4.3 MHz). We define the frequency shifts of the Zeeman levels in 5S1/2, 5P3/2 and 50S1/2 states relative to their frequencies at zero magnetic field as Δ1, Δ2 and Δ3, respectively. Then the frequency detunings of probe and coupling lasers in magnetic fields relative to ω0p and ω0c are Δp = Δ2 − Δ1 and Δc = Δ3 − Δ2. Considering the wavelength mismatch, the detunings satisfy the relation Δp+Δc=(Δ2Δ1)+λcλp(Δ3Δ2). When the coupling laser is on the resonant transition of 5P3/2, F′ = 3 → nS1/2, the EIT subpeaks exist at Δp=(Δ2Δ1)+λcλp(Δ3Δ2)=Δ1+(1λcλp)Δ2+λcλpΔ3. Our calculated results explains the experimental peculiarities well.

 figure: Fig. 8

Fig. 8 The experimental and calculated spectra of Rydberg EIT at various values of magnetic field from 0 to 47 Gauss. A narrow linewidth is used for resolving the lines.

Download Full Size | PDF

As discussed previously, the ground state 5S1/2 has a large energy gap (6.8 GHz) between the hyperfine levels and the interaction between them can be neglected so that the Zeeman effect shows a linear behavior vs. the magnetic field below 1000 Gauss. On the other extreme that the hyperfine levels are nearly degenerated for the high Rydberg state 50S1/2, the Zeeman is completely linear beyond a nonzero magnetic field (> 10 Gauss) as well. This feature implies that the dependence of the spectral transition frequency on the applied magnetic field relies on the intermediate state 5P3/2 mainly. For the intermediate state 5P3/2, the hyperfine structure splittings are about several hundreds of MHz, comparable with the magnetic field induced interaction, which results in a non-linear dependence of the total transition frequency on the magnetic field. However, for the cases whether in a weak magnetic field or in a strong one, our calculation based on the Hamiltonian in the coupled basis works well, even for the line intensity. Specially, for the Rydberg state, we can project the wavefunction back into the decoupled basis for spectral line assignment in a usual way.

5. Conclusion

In summary, we have investigated the double-resonance in a 5S1/2 → 5P3/2 → 50S1/2 ladder-type EIT system for 87Rb atom in magnetic fields. For the ground state 5S1/2 and intermediate state 5P3/2, the hyperfine levels are well separated by an energy gap much larger than the Zeeman interaction and the Zeeman interaction can be taken as a perturbation for the hyperfine energy levels in the coupled basis. However, for the nearly degenerated high Rydberg state 50S1/2, this Zeeman interaction can not be taken as a perturbation any more and the coupled hyperfine states will mix seriously. The eigen state of 50S1/2 in magnetic field has to be viewed as an linear combination of eigen wavefunction |FmF〉 in the coupled basis. For a better understanding for the Zeeman-splitting of Rydberg state in a usual way, we project the wavefunction back into the decoupled basis where the spectral line is well assigned by a pair of (mJ, mI). All the experimental observations are well explained by the combination of coupled and decoupled basis.

Funding

National Natural Science Foundation of China (NSFC) (91421305, 91121005, 11674359); Program 973 (2013CB922003).

Disclosures

The authors declare that there are no conflicts of interest related to this article.

References and links

1. A. Peng, M. Johnsson, W. P. Bowen, P. K. Lam, H. A. Bachor, and J. J. Hope, “Squeezing and entanglement delay using slow light,” Phys. Rev. A 71, 033809 (2005). [CrossRef]  

2. A. S. Zibrov, M. D. Lukin, D. E. Nikonov, L. Hollberg, M. O. Scully, V. L. Velichansky, and H. G. Robinson, “Experimental demonstration of laser oscillation without population inversion via quantum interference in rb,” Phys. Rev. Lett. 75, 1499–1502 (1995). [CrossRef]   [PubMed]  

3. J. Mompart and R. Corbalán, “Lasing without inversion,” J. Opt. B 2, R7 (2000). [CrossRef]  

4. C. Liu, Z. Dutton, C. H. Behroozi, and L. V. Hau, “Observation of coherent optical information storage in an atomic medium using halted light pulses,” Nature 409, 490–493 (2001). [CrossRef]   [PubMed]  

5. S. M. Iftiquar, G. R. Karve, and V. Natarajan, “Subnatural linewidth for probe absorption in an electromagnetically-induced-transparency medium due to doppler averaging,” Phys. Rev. A 77, 063807 (2008). [CrossRef]  

6. L. Karpa, G. Nikoghosyan, F. Vewinger, M. Fleischhauer, and M. Weitz, “Frequency matching in light-storage spectroscopy of atomic raman transitions,” Phys. Rev. Lett. 103, 093601 (2009). [CrossRef]   [PubMed]  

7. H. Fan, S. Kumar, J. Sedlacek, H. Kübler, S. Karimkashi, and J. P. Shaffer, “Atom based rf electric field sensing,” J. Phys. B 48, 202001 (2015). [CrossRef]  

8. C. L. Holloway, J. A. Gordon, S. Jefferts, A. Schwarzkopf, D. A. Anderson, S. A. Miller, N. Thaicharoen, and G. Raithel, “Broadband rydberg atom-based electric-field probe for si-traceable, self-calibrated measurements,” Ieee. T. Antenn. Propag. 62, 6169–6182 (2014). [CrossRef]  

9. X.-G. Wei, J.-H. Wu, G.-X. Sun, Z. Shao, Z.-H. Kang, Y. Jiang, and J.-Y. Gao, “Splitting of an electromagnetically induced transparency window of rubidium atoms in a static magnetic field,” Phys. Rev. A 72, 023806 (2005). [CrossRef]  

10. Vladan Pavlović and Ljiljana Stevanović, “Group velocity of light in a three level ladder-type spherical quantum dot with hydrogenic impurity,” Superlattice. Microst. 100, 500–507 (2016). [CrossRef]  

11. Z. Wang, “Control of the optical multistability in a three-level ladder-type quantum well system,” Opt. Commun. 282, 4745–4748 (2009). [CrossRef]  

12. A. Sargsyan, R. Mirzoyan, T. Vartanyan, and D. Sarkisyan, “Determination of the structure of hyperfine sublevels of rb in strong magnetic fields by means of the coherent population trapping technique,” J. Exp. Theor. Phys. 118, 359–364 (2014). [CrossRef]  

13. X. Jiang, H. Zhang, and Y. Wang, “Electromagnetically induced transparency in a zeeman-sublevels λ-system of cold 87rb atoms in free space,” Chin. Phys. B 25, 034204 (2016). [CrossRef]  

14. A. M. Akulshin, S. Barreiro, and A. Lezama, “Electromagnetically induced absorption and transparency due to resonant two-field excitation of quasidegenerate levels in rb vapor,” Phys. Rev. A 57, 2996–3002 (1998). [CrossRef]  

15. D. Ya-Bin, G. Jiang-Rui, and D. You-Er, “Quantum coherent effects in multi-zeeman-sublevel atomic systems,” Chin. Phys. B 17, 3306–3312 (2008). [CrossRef]  

16. A. Sargsyan, R. Mirzoyan, and D. Sarkisyan, “Splitting of the electromagnetically induced transparency resonance on 85rb atoms in strong magnetic fields up to the paschen-back regime,” JETP Lett. 96, 303–307 (2012). [CrossRef]  

17. A. Sargsyan, D. Sarkisyan, and A. Papoyan, “Dark-line atomic resonances in a submicron-thin rb vapor layer,” Phys. Rev. A 73, 033803 (2006). [CrossRef]  

18. P. Valente, H. Failache, and A. Lezama, “Comparative study of the transient evolution of hanle electromagnetically induced transparency and absorption resonances,” Phys. Rev. A 65, 023814 (2002). [CrossRef]  

19. Y. Dong, H. Wang, J. Gao, and J. Zhang, “Quantum coherence effects in quasidegenerate two-level atomic systems,” Phys. Rev. A 74, 063810 (2006). [CrossRef]  

20. L. Weller, K. S. Kleinbach, M. A. Zentile, S. Knappe, C. S. Adams, and I. G. Hughes, “Absolute absorption and dispersion of a rubidium vapour in the hyperfine paschen-back regime,” J. Phys. B 45, 215005 (2012). [CrossRef]  

21. R. Wynands, A. Nagel, S. Brandt, D. Meschede, and A. Weis, “Selection rules and line strengths of zeeman-split dark resonances,” Phys. Rev. A 58, 196–203 (1998). [CrossRef]  

22. A. Sargsyan, R. Mirzoyan, and D. Sarkisyan, “A study of dark resonance splitting for the d1 line of 87rb in strong magnetic fields,” Opt. Spectrosc. 113, 456–462 (2012). [CrossRef]  

23. A. Sargsyan, C. Leroy, Y. Pashayan-Leroy, R. Mirzoyan, A. Papoyan, and D. Sarkisyan, “High contrast d1 line electromagnetically induced transparency in nanometric-thin rubidium vapor cell,” Appl. Phys. B 105, 767–774 (2011). [CrossRef]  

24. Z.-S. He, J.-H. Tsai, Y.-Y. Chang, C.-C. Liao, and C.-C. Tsai, “Ladder-type electromagnetically induced transparency with optical pumping effect,” Phys. Rev. A 87, 033402 (2013). [CrossRef]  

25. M. Mack, F. Karlewski, H. Hattermann, S. Höckh, F. Jessen, D. Cano, and J. Fortágh, “Measurement of absolute transition frequencies of 87rb to ns and nd rydberg states by means of electromagnetically induced transparency,” Phys. Rev. A 83, 052515 (2011). [CrossRef]  

26. Z.-S. He, J.-H. Tsai, M.-T. Lee, Y.-Y. Chang, C.-C. Tsai, and T.-J. Whang, “Determination of the cesium 11s2s1/2 hyperfine magnetic coupling constant using electromagnetically induced transparency,” J. Phys. Soc. Jpn. 81, 124302 (2012). [CrossRef]  

27. A. Sargsyan, D. Sarkisyan, U. Krohn, J. Keaveney, and C. Adams, “Effect of buffer gas on an electromagnetically induced transparency in a ladder system using thermal rubidium vapor,” Phys. Rev. A 82, 045806 (2010). [CrossRef]  

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

29. B. Yang, J. Gao, T. Zhang, and J. Wang, “Electromagnetically induced transparency without a doppler background in a multilevel ladder-type cesium atomic system,” Phys. Rev. A 83, 013818 (2011). [CrossRef]  

30. H.-R. Noh and H. S. Moon, “Transmittance signal in real ladder-type atoms,” Phys. Rev. A 85, 033817 (2012). [CrossRef]  

31. R. Sobolewski, J. Wang, J. Wang, H. Liu, B. Yang, J. He, and J. Fiurásek, “Measurement of hyperfine splitting and determination of hyperfine structure constant of cesium 8s1/2 state by using of ladder-type eit,” Proc. SPIE 8773, 877311 (2013). [CrossRef]  

32. H. S. Moon and T. Jeong, “Three-photon electromagnetically induced absorption in a ladder-type atomic system,” Phys. Rev. A 89, 033822 (2014). [CrossRef]  

33. D. E. Jones, J. D. Franson, and T. B. Pittman, “Ladder-type electromagnetically induced transparency using nanofiber-guided light in a warm atomic vapor,” Phys. Rev. A 92, 043806 (2015). [CrossRef]  

34. D. J. Whiting, E. Bimbard, J. Keaveney, M. A. Zentile, C. S. Adams, and I. G. Hughes, “Electromagnetically induced absorption in a nondegenerate three-level ladder system,” Opt. Lett. 40, 4289–4292 (2015). [CrossRef]   [PubMed]  

35. L. Ma, D. A. Anderson, and G. Raithel, “Paschen-back effects and rydberg-state diamagnetism in vapor-cell electromagnetically induced transparency,” Phys. Rev. A 95, 061804 (2017). [CrossRef]  

36. A. Sargsyan, M. G. Bason, D. Sarkisyan, A. K. Mohapatra, and C. S. Adams, “Electromagnetically induced transparency and two-photon absorption in the ladder system in thin columns of atomic vapors,” Opt. Spectrosc. 109, 529–537 (2010). [CrossRef]  

37. A. Tauschinsky, R. Newell, H. B. van Linden van den Heuvell, and R. J. C. Spreeuw, “Measurement of 87rb rydberg-state hyperfine splitting in a room-temperature vapor cell,” Phys. Rev. A 87, 042522 (2013). [CrossRef]  

38. J. D. Pritchard, D. Maxwell, A. Gauguet, K. J. Weatherill, M. P. A. Jones, and C. S. Adams, “Cooperative atom-light interaction in a blockaded rydberg ensemble,” Phys. Rev. Lett. 105, 193603 (2010). [CrossRef]  

39. A. K. Mohapatra, T. R. Jackson, and C. S. Adams, “Coherent optical detection of highly excited rydberg states using electromagnetically induced transparency,” Phys. Rev. Lett. 98, 113003 (2007). [CrossRef]   [PubMed]  

40. J. Naber, R. Spreeuw, A. Tauschinsky, and B. van Linden van den Heuvell, “Electromagnetically induced transparency with rydberg atoms across the breit-rabi regime,” SciPost Phys. 2, 015 (2017). [CrossRef]  

41. S. Bao, W. Yang, H. Zhang, L. Zhang, J. Zhao, and S. Jia, “Splitting of an electromagnetically induced transparency window of a cascade system with 133cs rydberg atoms in a static magnetic field,” J. Phys. Soc. Jpn. 84, 104301 (2015). [CrossRef]  

42. S. Bao, H. Zhang, J. Zhou, L. Zhang, J. Zhao, L. Xiao, and S. Jia, “Polarization spectra of zeeman sublevels in rydberg electromagnetically induced transparency,” Phys. Rev. A 94, 043822 (2016). [CrossRef]  

43. S. X. Bao, H. Zhang, J. Zhou, L. J. Zhang, J. M. Zhao, L. T. Xiao, and S. T. Jia, “Tunable frequency stabilization to zeeman sublevel transitions between an intermediate state and rydberg states,” Laser Phys. 27, 015701 (2017). [CrossRef]  

44. A. Sargsyan, G. Hakhumyan, C. Leroy, Y. Pashayan-Leroy, A. Papoyan, D. Sarkisyan, and M. Auzinsh, “Hyperfine paschen-back regime in alkali metal atoms: consistency of two theoretical considerations and experiment,” J. Opt. Soc. Am. B 31, 1046 (2014). [CrossRef]  

45. H. Cheng, H.-M. Wang, S.-S. Zhang, P.-P. Xin, J. Luo, and H.-P. Liu, “Electromagnetically induced transparency of 87rb in a buffer gas cell with magnetic field,” J. Phys. B 50, 095401 (2017). [CrossRef]  

46. M. U. Momeen, G. Rangarajan, and P. C. Deshmukh, “Variations of intensity in rb d2 line at weak/intermediate fields,” J. Phys. B 40, 3163–3172 (2007). [CrossRef]  

47. A. Sargsyan, A. Tonoyan, G. Hakhumyan, A. Papoyan, E. Mariotti, and D. Sarkisyan, “Giant modification of atomic transition probabilities induced by a magnetic field: forbidden transitions become predominant,” Laser Phys. Lett. 11, 055701 (2014). [CrossRef]  

48. G. Hakhumyan, C. Leroy, R. Mirzoyan, Y. Pashayan-Leroy, and D. Sarkisyan, “Study of “forbidden” atomic transitions on d2 line using rb nano-cell placed in external magnetic field,” Eur. Phys. J. D 66119 (2012). [CrossRef]  

49. A. K. Mohapatra, T. R. Jackson, and C. S. Adams, “Coherent optical detection of highly excited rydberg states using electromagnetically induced transparency,” Phys. Rev. Lett. 98, 113003 (2007). [CrossRef]   [PubMed]  

50. J. Zhao, X. Zhu, L. Zhang, Z. Feng, C. Li, and S. Jia, “High sensitivity spectroscopy of cesium rydberg atoms using electromagnetically induced transparency,” Opt. Express 17, 15821–15826 (2009). [CrossRef]   [PubMed]  

51. A. Javan, O. Kocharovskaya, H. Lee, and M. O. Scully, “Narrowing of electromagnetically induced transparency resonance in a doppler-broadened medium,” Phys. Rev. A 66, 013805 (2002). [CrossRef]  

52. H. Cheng, S.-S. Zhang, P.-P. Xin, Y. Cheng, and H.-P. Liu, “Theoretical simulation of 87rb absorption spectrum in a thermal cell,” Chin. Phys. B 25, 114203 (2016). [CrossRef]  

53. H. Cheng, H.-M. Wang, S.-S. Zhang, P.-P. Xin, J. Luo, and H.-P. Liu, “High quality electromagnetically induced transparency spectroscopy of 87rb in a buffer gas cell with a magnetic field,” Chin. Phys. B 26, 74204 (2017). [CrossRef]  

54. A. Krishna, K. Pandey, A. Wasan, and V. Natarajan, “High-resolution hyperfine spectroscopy of excited states using electromagnetically induced transparency,” Europhys. Lett. 72, 221–227 (2005). [CrossRef]  

55. Y. Li, S. Jin, and M. Xiao, “Observation of an electromagnetically induced change of absorption in multilevel rubidium atoms,” Phys. Rev. A 51, R1754–R1757 (1995). [CrossRef]   [PubMed]  

56. K. Kim, M. Kwon, H. D. Park, H. S. Moon, H. S. Rawat, K. An, and J. B. Kim, “Electromagnetically induced absorption spectra depending on intensities and detunings of the coupling field in cs vapour,” J. Phys. B 34, 4801 (2001). [CrossRef]  

57. M. Wang, X.-G. Lu, J.-H. Bai, L.-Y. Pei, X.-X. Miao, Y.-L. Gao, L.-A. Wu, P.-M. Fu, S.-P. Yang, Z.-G. Pang, R.-Q. Wang, and Z.-C. Zuo, “Kramers-kronig relation in a doppler-broadened λ-type three-level system,” Chin. Phys. B 24, 114205 (2015). [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 (8)

Fig. 1
Fig. 1 The energy level scheme (a) and experimental setup (b) of ladder-type EIT system for 87Rb. It should be noted that the hyperfine levels of Rydberg state are nearly degenerate in free fields for a vanishing coupling constant between the nuclear spin I and the total angular momentum J. The 780 nm probe (thin red line) and 480 nm coupling (thick blue line) beams are counter-propagated and overlapped in the vapor cell. PBS: polarizing beam splitter; DM: 45° dichroic mirror; BS: beam splitter; M: mirror; λ/2: half wave plate; λ/4: quarter wave plate; BPD: balanced photodetector.
Fig. 2
Fig. 2 Experimentally measured (blue lines) and theoretically calculated (red lines) absorptions of the probe beam as a function of probe beam frequency shift for different coupling detunings. The wavelength mismatching effect has been included in the calculation.
Fig. 3
Fig. 3 Ladder-type EIT system of 87Rb atom in a magnetic field. The probe laser frequency ωp is scanned over the F = 2 to 5P3/2 transition, whereas the frequency ωc is resonant to the transition F′ = 3 → 50S1/2. ω0p and ω0c denote the frequencies of the probe and coupling lasers in zero magnetic field, respectively.
Fig. 4
Fig. 4 The spectra for the probe and coupling laser σ+ and σ polarized in a magnetic field B = 37 Gauss. The red line is the experimental observation while the blue line is the calculated one. In the coupled basis, F is no longer a good quantum number for the Rydberg states except for the lines D and E.
Fig. 5
Fig. 5 The assignment of the Rydberg state in the three-level ladder-type EIT configuration for the spectral line A and A′ shown in Fig. 4. In the coupled basis, different F-component of Rydberg state mixes each other, making the coupling laser can also transit from F′ = 3 to F″ = 1 (in dotted olive lines), but they can be described by the good quantum numbers mJ and mI in the decoupled basis |JmJ〉|ImI〉 as shown.
Fig. 6
Fig. 6 The spectra for the probe and coupling laser σ+ and σ+ polarized in a magnetic field B = 40 Gauss. Similar to that shown in Fig. 4, the Rydberg state has to been described by a decoupled basis |JmJ, ImI〉 rather than a coupled basis.
Fig. 7
Fig. 7 The calculated spectra of Rydberg EIT at B = 40 Gauss for the polarization combinations of σ+σ and σ+σ+. They have completely different spectral patterns.
Fig. 8
Fig. 8 The experimental and calculated spectra of Rydberg EIT at various values of magnetic field from 0 to 47 Gauss. A narrow linewidth is used for resolving the lines.

Tables (2)

Tables Icon

Table 1 Assignment for the spectral lines shown in Fig. 4

Tables Icon

Table 2 Assignment for the spectral lines shown in Fig. 6

Equations (8)

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

H ^ ¯ = H ^ 0 + H ^ fs + H ^ hfs + H ^ B ,
| ( JI ) F m F = ( 1 ) J I + m F 2 F + 1 × m J , m I ( J I F m J m I m F ) | J m J | I m I ,
Δ E hfs = A hfs 2 K + B hfs 4 3 2 K ( K + 1 ) 2 I ( I + 1 ) J ( J + 1 ) I ( 2 I 1 ) J ( 2 J 1 ) ,
H B = μ B g J J B + μ B g I I B ,
( JI ) F i m F | μ B g J J B | ( JI ) F j m F = μ B B g J ( 1 ) J + I + F i + F j m F + 1 × ( 2 F i + 1 ) ( 2 F j + 1 ) J ( J + 1 ) ( 2 J + 1 ) × ( F i 1 F j m F 0 m F ) { J F i I F j J 1 }
( JI ) F i m F | μ B g I I B | ( JI ) F j m F = μ B B g I ( 1 ) J + I + F i + F j m F + 1 × ( 2 F i + 1 ) ( 2 F j + 1 ) I ( I + 1 ) ( 2 I + 1 ) × ( F i 1 F j m F 0 m F ) { I F i I F j I 1 }
σ m F m F = F m F | e r | F m F = J | e r | J ( 1 ) 2 F + J + I + m F ( 2 F + 1 ) ( 2 F + 1 ) ( 2 J + 1 ) × ( F 1 F m F m F m F m F ) { J J 1 F F I } .
Ψ = F | m F | C F | F m F
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.