This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Paper The following article is Open access

Transition from electromagnetically induced transparency to Autler–Townes splitting in cold cesium atoms

, , , , , and

Published 17 July 2018 © 2018 The Author(s). Published by IOP Publishing Ltd on behalf of Deutsche Physikalische Gesellschaft
, , Citation Liping Hao et al 2018 New J. Phys. 20 073024 DOI 10.1088/1367-2630/aad153

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

1367-2630/20/7/073024

Abstract

Electromagnetically induced transparency (EIT) and Autler–Townes splitting (ATS) are two similar yet distinct phenomena that modify the transmission of a weak probe field through an absorption medium in the presence of a coupling field, featured in a variety of three-level atomic systems. In many applications it is important to distinguish EIT from ATS splitting. We present EIT and ATS spectra in a three-level cascade system, involving cold cesium atoms in the $35{S}_{1/2}$ Rydberg state. The EIT linewidth, γEIT, defined as the full width at half maximum of the transparency window, and the ATS splitting, γATS, defined as the peak-to-peak distance between AT absorption peaks, are used to delineate the EIT and ATS regimes and to characterize the transition between the regimes. In the cold-atom medium, in the weak-coupler (EIT) regime γEIT ≈ A + B(${{\rm{\Omega }}}_{c}^{2}$ + ${{\rm{\Omega }}}_{p}^{2})/{{\rm{\Gamma }}}_{{eg}}$, where Ωc and Ωp are the coupler and probe Rabi frequencies, Γeg is the spontaneous decay rate of the intermediate 6P3/2 level, and parameters A and B that depend on the laser linewidth. We explore the transition into the strong-coupler (ATS) regime, which is characterized by the relation γATS ≈ Ωc. The experiments are in agreement with numerical solutions of the Master equation. Our analysis accounts for non-ideal conditions that exist in typical realizations of Rydberg-EIT, including laser-frequency jitter, Doppler mismatch of the utilized two-color Rydberg EIT system, and strong probe fields. The obtained criteria to distinguish cold-atom EIT from ATS are readily accessible and applicable in practical implementations.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Electromagnetically induced transparency (EIT) [1] is a quantum interference effect in which the absorption of a weak probe laser, interacting resonantly with an atomic transition, is reduced in the presence of a coupling laser, which (near-)resonantly couples the upper probe level to a third level. EIT is, for instance, crucial in optically controlled slowing of light [2] and optical storage [3]. Autler–Townes splitting (ATS) [4], proposed by Autler and Townes, is a linear (resonant) AC Stark effect in a two-level system that is strongly driven at resonance, and can be measured via a weak probe from a third level. ATS was observed originally in the microwave and later the light domain. EIT and ATS have been extensively investigated experimentally and theoretically in Λ−, V− and cascade-type three-level atoms [1, 510].

EIT and ATS may phenomenologically look similar, but they are different in nature, one being a quantum interference and the other a linear AC Stark effect. This leads to an interest in the establishment of criteria to discern them. In [9], threshold values for distinguishing EIT and ATS, Ωt, are defined via the coherence decay rates in the Master equation of the various systems, according to which ATS is observed in four different three-level systems in a strong-coupling-field regime (Ωc / Ωt > 1), whereas EIT is observed only in Λ- and cascade EIT configurations in a weak-coupling-field regime (Ωc/ Ωt < 1). Anisimov et al [10] also propose an objective method based on Akaike's information criterion (AIC), to discern ATS from EIT in experimental data of a Λ-type three-level atom. Recently, a series of experiments [1113] has been performed in which the AIC has been used to discriminate EIT from ATS in Λ-type three-level systems free of Doppler effects. Zhang et al [14] provided a theoretical model using a five-level system and investigated EIT and ATS based on the AIC. Another area of interest, in which EIT and ATS are important, is spectroscopy of highly exited Rydberg levels using all-optical methods. Rydberg-EIT, a cascade EIT case, was first observed in a vapor cell [15] and later in a rubidium MOT [16]. In Rydberg-EIT work, the phenomena under investigation are often due to strong Rydberg-atom interactions that are distinct from the optical couplings that give rise to EIT or ATS. For instance, Rydberg-EIT has been used to realize a single-photon transistor [17] and a single-photon source [18, 19] by employing a blockade effect [2022], which results from the strong Rydberg-atom interactions. ATS has also been investigated in systems that have involved Rydberg states of rubidium [23, 24] and cesium [25, 26]. Exploiting the fact that Rydberg atoms strongly couple to radio-frequency (RF) fields, Rydberg EIT and ATS effects are used to realize precise measurements of such fields [2731]. Holloway et al [32] have investigated the relationship between the Rabi frequency of resonant RF transitions between Rydberg states and the resultant ATS splitting in Rydberg-EIT spectra measured in room-temperature atomic vapor. In related work, the enhancement and suppression of multi-wave mixing (MWM) in five-level systems [33] has been studied, and the effect of Rydberg-atom interactions was explored [34]. Spectra and spatial images of probe and MWM signals were obtained for six-wave [35] and eight-wave implementations [36]. The role of Rydberg interactions as well as the co-occurrence of Rydberg-EIT and ATS were discussed in these works.

While numerous efforts have already dealt with discriminating EIT from ATS [914], as reported in the above overview, many of these earlier works are limited to Λ-type systems, deal with Doppler-effect-free cold-atom systems or all-optical equivalents of those, assume that the probe is in the weak-field limit, assume that the system is in steady-state, assume a vanishing laser linewidth, and/or assume that there is no interaction between the atoms (or their analogs) that exhibit EIT or ATS. In the present work we explore the EIT/ATS transition in a case where none of these ideal conditions are necessarily satisfied. We use cascade EIT and ATS, with a Rydberg upper-level state, in a cesium magneto-optical trap (MOT) as an experimental test system. We introduce two parameters, the Rydberg-EIT linewidth, γEIT, and the AT spilitting, γATS, to characterize the EIT to ATS transition. In our model, we calculate spectra using methods provided in [32] to study the dependence of γEIT and γATS on the Rabi frequencies of both the probe and coupling transitions. The transition between EIT and ATS effect is linked to a qualitative change in behavior of γEIT and γATS as a function of coupler and probe Rabi frequencies. In our analysis, we expand the above cited previous studies to a cascade Rydberg EIT/ATS system. In the numerical methods used, we do not make any weak-field approximations. Also, the Doppler effect, residual Doppler mismatch, averaging over the Maxwell velocity distribution, and inhomogeneous broadening due to laser-frequency jitter are accounted for. The model does not include Rydberg-atom interactions, which are not relevant under the conditions (Rydberg quantum numbers, atom densities and interaction times) used in the experiment. We obtain readily accessible criteria to distinguish EIT from ATS.

2. Theoretical model

We consider the cesium cascade three-level system shown in figure 1(a). The coupling laser drives the upper transition, $| 6{P}_{3/2}$, F$=\,5\rangle $ ($| 2\rangle $) $\to | 35{S}_{1/2}\rangle $ ($| 3\rangle $). The weak probe laser couples the lower transition, $| 6{S}_{1/2}$, ${F}=4\rangle $ ($| 1\rangle $) $\to | 6{P}_{3/2}$, F$=\,5\rangle $ ($| 2\rangle $). The respective wavelengths and Rabi frequencies are λc and Ωc, and λp and Ωp. In the rotating-wave approximation and the field picture, the Hamiltonian of the three-level atom represented in the space $\{| 1\rangle ,| 2\rangle ,| 3\rangle \}$ is

Equation (1)

where Δc and Δp are the detunings of the coupling and probe beams, respectively. To account for decay and dephasing, the system is described using the Lindblad equation for the density matrix ρ

Equation (2)

where ${\mathfrak{L}}$ is the Lindblad operator that accounts for the decay processes in the atom. In the space $\{| 1\rangle ,| 2\rangle ,| 3\rangle \}$, ${\mathfrak{L}}$ becomes [37]

Equation (3)

where γ2 and γ3 are the dephasing rates of the intermediate and Rydberg states, respectively. It is γ2 = γe + Γeg, where γe is a collision-induced dephasing rate of level $| 2\rangle $, and γe ≪ Γeg = 2π × 5.2 MHz. Further, γ3 = γr + Γre. For the Rydberg level the population decay rate Γre is, typically, smaller than the dephasing γr, because Rydberg-atom lifetimes are long (lifetimes are ∼n3 and on the order of 100 μs) and interactions between Rydberg atoms in cold-atom clouds are often strong (van der Waals interactions scale as n11).

Figure 1.

Figure 1. (a) Scheme of the cascade three-level atom. The coupling laser is resonant with the Rydberg transition, $| 6{P}_{3/2}$, F'$\,=\,5\rangle $ ($| 2\rangle $) → $| 35{S}_{1/2}\rangle $ ($| 3\rangle $) (wavelength 510 nm, Rabi frequency Ωc). The weak probe beam (wavelength 852 nm, Rabi frequency Ωp) is referenced to the transition $| 6{S}_{1/2},F=4\rangle $ ($| 1\rangle $)  → $| 6{P}_{3/2}$, F$=\,5\rangle $ ($| 2\rangle $) using a Doppler-free polarization spectroscopy setup, and is scanned over the resonance. (b) Sketch of the experimental setup. The laser beams are separated via a dichroic mirror (not shown), and the probe light is detected using a single-photon counter module (SPCM).

Standard image High-resolution image

The spectrum is given by the probe-power transmission, $P={P}_{0}\exp (-\alpha L)$, with the probe-laser absorption coefficient, α = 2πIm(χ)/λp, the MOT size, L, and the susceptibility of the medium seen by the probe laser, χ. The susceptibility, χ, is

Equation (4)

where N is the average atomic density, μ12 is the dipole moment of transition $| 1\rangle \to | 2\rangle $, Ep is the amplitude of the probe, epsilon0 is the vacuum permittivity, and ρ12 is the density matrix element between $| 1\rangle $ and $| 2\rangle $.

We numerically solve the equations (1)–(3) to obtain the absorption coefficient α for a range of values of Ωc and Ωp, including the case of a strong probe, Ωp ≳ Γeg. The result is averaged over the thermal velocity distribution in the gas [32]; under our conditions (T = 100–200 μK) the thermal motion is marginally important. In the calculation we also assume γr = 0, which is admissible due to our very low experimental probe intensities and the low principal quantum number of the utilized Rydberg state. Further, in our initial discussion we assume zero laser linewidth; this assumption is dropped later-on.

In figure 2(a) we show α as a function of Ωc and Δp, for the case of a fixed Ωp = 2π × 1.05 MHz and Δc = 0. The shape of the spectrum clearly indicates two qualitatively different regimes, a regime of low Ωc in which the overall linewidth is fixed at ≈Γeg and there is a very narrow transparency window whose width scales quadratically in Ωc. This regime is the EIT regime. As Ωc increases, the overall linewidth increases and the system enters the ATS regime. There, the line breaks up into two features that are narrower than the coupler-free absorption line and the peak-to-peak separation scales linearly in Ωc. The wide transparency region between the AT lines is almost as wide as the peak-to-peak separation itself.

Figure 2.

Figure 2. (a) Calculation of the probe absorption coefficient of cold cesium atoms in a MOT, α, for Ωc ranging from 2π × 0.5 MHz to 2π × 16 MHz. The cuts along the horizontal dashed lines correspond to the spectra shown in (b) and (c). (b) Probe absorption coefficient for Ωp = 2π × 1.05 MHz and Ωc = 2π × 2.03 MHz, atom density 1010 cm−3, and vanishing laser linewidth. This case is in the EIT regime. The FHWM of the EIT window, γEIT, and the ATS, γATS, are defined as shown. (c) Same as (b), but for Ωc = 2π × 13.01 MHz. This case is in the AT regime.

Standard image High-resolution image

For a quantitative analysis, we extract the full width at half maximum (FWHM) of the dip in the absorption, γEIT, as well as the peak-to-peak spacing of the line pair, γATS, from the absorption coefficient data. Both parameters γEIT and γATS can be calculated in both the EIT and the ATS regime. However, as detailed in the following, γATS yields little useful information in the EIT regime and vice versa. In figure 2(b), both parameters are shown for a typical spectrum in the EIT regime, where Ωp ≲ Ωc = 2π × 2.03 MHz < Γeg. In EIT, the main interest lies in how narrow the EIT transparency window is, in comparison with the natural linewidth Γeg. Therefore, in the EIT regime the parameter γEIT is most important (hence the name); it measures the bandwidth over which the EIT quantum interference mechanism leading to reduced probe absorption is experimentally observable. In the EIT regime, the separation between the absorption maxima, γATS, carries little useful information, as it just returns a value on the order of the overall coupler-free linewidth. Figure 2(c) shows the spectrum calculated for Ωc = 2π × 13.01 MHz > Γeg, which is the ATS regime. Since ATS is a linear AC Stark effect, the interest mainly is in how far the AT-split line centers are separated from each other. Therefore, in the ATS regime a better parameter to quantify the behavior is the peak-to-peak spacing of the line pair, γATS. As seen in figures 2(a) and (c), in the weak probe limit and for Δc = 0, in the ATS regime the peaks tend to have a width of Γeg/2 = 2π × 2.6 MHz, and a spacing γATS ≈ Ωc. (The Rydberg states are very long-lived (Γre ≪ Γeg) and do not contribute to the width of the AT peaks here).

Figure 3(a) shows the calculated γEIT as a function of both Ωc and Ωp. In the EIT domain, characterized by Ωc, Ωp ≲ Γeg and outlined by the dashed line, the separation between the contour lines decreases as a function of distance from the origin, reflecting that in the EIT regime ${\gamma }_{{\rm{EIT}}}\approx ({{\rm{\Omega }}}_{c}^{2}+{{\rm{\Omega }}}_{p}^{2})/{{\rm{\Gamma }}}_{{eg}}$. Further out, as the system enters the ATS regime, the separation between the contour lines tends to become fixed. Figure 3(b) shows that in the AT regime, characterized by Ωc ≳ Γeg, the γATS contour lines are equidistant circles, reflecting the fact that ${\gamma }_{{\rm{ATS}}}\approx \sqrt{{{\rm{\Omega }}}_{c}^{2}+{{\rm{\Omega }}}_{p}^{2}}$ in that regime. In figure 3(b), near the x-axis, and in figure 2(c) it is Ωp ≪ Ωc and therefore γATS ≈ Ωc. These trends for γATS are also borne out by calculating the separation between the pair of non-dark eigenstates of the (decay-free) Hamiltonian in equation (1), when setting all detunings to zero.

Figure 3.

Figure 3. (a) Calculated γEIT as a function of Ωp and Ωc. Within the displayed parameter space, the EIT regime is found in the lower-left corner. In that regime, ${\gamma }_{{\rm{EIT}}}\approx ({{\rm{\Omega }}}_{c}^{2}+{{\rm{\Omega }}}_{p}^{2})/{{\rm{\Gamma }}}_{{eg}}\lesssim {{\rm{\Gamma }}}_{{eg}}$. (b) Calculated γATS as a function of Ωp and Ωc. Except in a region of coupler Rabi frequencies considerably less than Γeg, it is ${\gamma }_{{\rm{ATS}}}\approx \sqrt{{{\rm{\Omega }}}_{c}^{2}+{{\rm{\Omega }}}_{p}^{2}}$.

Standard image High-resolution image

Calculations similar to those shown in figures 2 and 3 can be performed at any atom temperature (see [32] and references therein), and for both cascade and Λ-type three-level atoms. In the hot-atom cases (i.e., in vapor cells), the difference between coupler and probe wavelengths determines the size of residual Doppler shifts, which in turn determines the visibilities and detailed shapes of EIT and ATS spectra. In hot gases γEIT generally tends to be larger than in the cold-atom case, in particular in weak coupler and probe fields, and follows a different scaling (see figure 7 in [32]).

3. Experimental measurement

The EIT and ATS experiments are performed in a standard MOT with temperature ∼100 μK and atomic density up to ∼1010 cm−3. The coupling and probe lasers have linear and parallel polarizations and counter-propagate through the cold-atom cloud, as seen in figure 1(b). The details of the experiment have been described previously [38]. The probe beam is derived from a diode laser (DLpro, Toptica) that is locked to the ground-state transition, $| 1\rangle $ → $| 2\rangle $, using polarization spectroscopy [39]; the beam has a Gaussian waist ωp0 = 10 μm at the MOT center. The strong coupling laser (Toptica TA-SHG110) has a Gaussian waist ωc0 = 30 μm and drives the Rydberg transition $| 2\rangle \to | 3\rangle $. The frequency of the coupling laser is stabilized to the Rydberg transition using a Rydberg-EIT signal obtained from a cesium room-temperature vapor cell [40]. In each experimental cycle, after turning off the trap beams, we switch on the coupling and probe lasers for 25 μs. During the probe pulse the probe-laser frequency is swept across the $| 6{S}_{1/2},F=4\rangle $ → $| 6{P}_{3/2},{F}^{{\rm{{\prime} }}}=5 \rangle $ transition using a double-pass acousto-optic modulator (AOM) over a range of ±10 MHz relative to the transition center.

In order to avoid Rydberg excitation blockade and interaction effects, and to be able to reach high coupling Rabi frequencies, Ωc, we have chosen a state with a low principal quantum number n (35S1/2). Also, the MOT density is reduced to about ∼108 cm−3 by reducing the repumping laser power. Further, to ensure that the effect of radiation pressure [38, 41] on the EIT and ATS spectral profiles is negligible, we use a single-photon counter module (SPCM) for probe-light detection, allowing us to use a very low probe power, P852 = 200 pW. Under these conditions, the radiation-pressure-induced velocity change during the probe pulse has an upper limit of ≈7 cm s−1, corresponding to Doppler shifts <100 kHz (which is negligible in the present work). The EIT and ATS spectra are recorded using a data acquisition card (NI-PCI-6542) and processed with a Labview program.

The absorption coefficient, α, is obtained from the measured SPCM count number, Pp), and the off-resonant (absorption-free) count number, P0, using the relation $\alpha L=-\mathrm{ln}(P({{\rm{\Delta }}}_{p})/{P}_{0})$, where L is the effective MOT diameter along the probe beam path. In figure 4, we present experimental EIT and ATS sample spectra with indicated linewidth parameters γEIT and γATS. We have determined γEIT and γATS from experimental spectra covering a range of the coupling Rabi frequency Ωc, for fixed Ωp. In figure 5 we show the measured (hollow symbols) and calculated (lines) results for γEIT and γATS versus Ωc.

Figure 4.

Figure 4. Measurements of EIT and ATS absorption spectra for Ωp = 2π × 1.05 MHz and the indicated Rabi frequencies of the coupling laser, Ωc = 2π × 0 MHz (top), 3.48 MHz (middle) and 13.8 MHz (bottom). The solid lines show the results of Lorentzian multi-peak fits. The coupler-free linewidth (top curve) is 2π × (4.5 ± 0.3) MHz, which is close to the expected value of 2π × 5.2 MHz. The FWHM EIT linewidth, γEIT, and the ATS, γATS, are obtained from the fit functions, as indicated. The depth of the EIT dip (or, in the ATS regime, the depth of the valley between the AT peaks), H, is defined as the difference between the peak absorption, averaged over the two peaks, and the absorption minimum between the peaks.

Standard image High-resolution image
Figure 5.

Figure 5. Measurements (symbols) and calculations (solid lines) of γATS and γEIT as a function of Ωc for Ωp = 2π × 1.05 MHz and probe/coupling duration 25 μs (hollow) and 100 μs (filled). In the calculation, the laser linewidth is 2π × 1.5 MHz (bold solid lines) and 0 (thin solid lines). The black thin dashed line shows a fit to the experimental data; the fit function is ${\gamma }_{{\rm{EIT}}}\approx A+B({{\rm{\Omega }}}_{c}^{2}+{{\rm{\Omega }}}_{p}^{2})/{{\rm{\Gamma }}}_{{eg}}$ with A = 2π × (1.42 ± 0.10) MHz and B = 0.44 ± 0.03.

Standard image High-resolution image

To compare experimental spectra with calculated ones, it is important to account for the laser linewidth in the model. The utilized diode laser systems exhibit a frequency jitter on the order of 1 MHz FWHM, whereas high-frequency laser phase noise is negligible. The frequency jitter causes inhomogeneous broadening of the spectra, because in every realization of the experiment the lasers have a slightly different frequency, that does not change significantly over the duration of a single realization. To account for the inhomogeneous broadening, subsequent to solving the Master equation and obtaining calculated spectra such as the ones in figure 2, the spectra are convoluted with a Gaussian, whose width is given by the laser-frequency jitter. It is noted that high-frequency laser noise would cause homogeneous broadening, which would have to be accounted for differently, namely via additional dephasing terms in the off-diagonal terms in equation (3). The γEIT and γATS-values are obtained from the convoluted calculated spectra. In the present case, the calculations for 1.5 MHz FWHM of the laser-frequency jitter (bold solid lines in figure 5) reproduce the measurements best (also see discussion of figure 6 below).

Figure 6.

Figure 6. Measurements (symbols) and calculations (black solid lines) of the depth of the transparency window, H, defined in figure 4, versus Ωc for Ωp = 2π × 1.05 MHz. The FWHM of the laser jitter in the calculations ranges from 0 to 2.5 MHz. The curve for 1.5 MHz laser linewidth reproduces the experimental data best.

Standard image High-resolution image

We first discuss the EIT linewidth γEIT. In the EIT domain, characterized by Ωc, Ωp ≲ Γeg, the value of γEIT in figure 5 exhibits a quadratic behavior, as expected from figure 3(a). For the experimental data in figure 5, we find that in the EIT domain the data are fit quite well by an equation ${\gamma }_{{\rm{EIT}}}\approx A+B({{\rm{\Omega }}}_{c}^{2}+{{\rm{\Omega }}}_{p}^{2})/{{\rm{\Gamma }}}_{{eg}}$, as shown by the black thin dashed line. The fitting parameters, which are introduced to account for the inhomogeneous line broadening caused by the laser-frequency jitter, are found to be A = 2π × (1.42 ± 0.10) MHz and B = 0.44 ± 0.03. Without the inhomogeneous broadening, it would be A ≈ 0 and B ≈ 1. It is seen that the laser jitter reduces the pre-factor B to a value significantly less than its ideal value of 1, and it adds an additive constant A on the order of the laser linewidth. In the AT regime, characterized by Ωc ≳ Γeg, the γEIT-value follows a trend ${\gamma }_{{\rm{EIT}}}\approx {{\rm{\Omega }}}_{c}-2\pi \times 3.2\,\mathrm{MHz}$. (In the absence of laser-line broadening, in the AT regime it would be ${\gamma }_{{\rm{EIT}}}\approx {{\rm{\Omega }}}_{c}-{{\rm{\Gamma }}}_{{eg}}/2$ = Ωc − 2π × 2.6 MHz). Over the entire Ωc-range studied, the measured dependence γEITc) agrees quite well with a calculation in which a Gaussian laser-frequency jitter of 1.5 MHz FWHM is assumed.

Next we discuss γATS. In the AT regime, Ωc ≳ Γeg, and for Ωp ≪ Ωc, it is γATS ≈ Ωc. In the EIT regime, Ωc, Ωp ≲ Γeg, the value of γATS tends to be larger than Ωc; for Ωc ≲ 2π × 2 MHz γATS cannot be determined because the spectrum ceases to split into a separated line pair (also see figure 6). Inhomogeneous broadening due to the laser-frequency jitter generally increases the measured γATS, because the outer wings of the split lines are wider than the inside wings, hence the spectral averaging will pull the line centers in figure 2(b) outward. The absorption minimum at Δp = 0 disappears entirely when Ωc drops below a critical value, which is related to the laser linewidth and other broadening mechanisms. In our case this critical value is ≈2π × 2 MHz. Similar nonlinear behavior of the AT splitting is observed in [32], where an RF-induced ATS is used to measure a microwave electric field.

We note that Ωc has a different significance in the EIT and AT domains. In the AT regime, Ωc splits the lines but does not broaden them, whereas in the EIT regime Ωc broadens the EIT line, as γEIT increases with (Ωc2 + Ωp2). Conversely, laser jitter and other inhomogeneous broadening mechanisms also have a different significance in the two domains. In the EIT domain the spectra are very sensitive to inhomogeneous broadening, because even moderate laser jitter already strongly affects the parameters A and B in the fit function γEIT ≃ AB(${{\rm{\Omega }}}_{c}^{2}+{{\rm{\Omega }}}_{p}^{2}$)/Γeg. (In the absence of broadening, it would be A = 0 and B = 1.) In the ATS domain, the broadening has a comparatively small effect on γEIT, and an even smaller one on γATS. Generally, the agreement between theory and experiment in figure 5 is slightly better for γATS than it is for γEIT. We believe this is because the details of the line broadening are harder to model than the AT splitting of the line centers.

In figure 5 we have included a similar measurement with 100 μs probe- and coupler-pulse sweeps (filled diamonds in figure 5). It is seen that γATS for the longer pulses is less than in the 25 μs case. We attribute the reduction in γATS in part to dephasing caused by Rydberg-atom interactions, which could be considered via the dephasing terms in equation (3). Longer pulses will generally lead to a higher Rydberg-atom number in the atom-field interaction volume, causing dephasing by Rydberg–Rydberg interactions. The dependence of the ATS splitting on Rydberg interactions is detailed in [26]; in our present work interactions are avoided by using a low principal quantum number, a low atom density and a relatively short interaction time. Longer times will also increase the likelihood of Penning and thermal ionization. Any ions in the sample would contribute to Rydberg-level dephasing via the ion electric fields. With increasing pulse duration, the spectra are also increasingly affected by radiation-pressure-induced Doppler shifts, causing changes in the γATS-values derived from the spectra.

We finally consider the depth of the transparency window, which is an important measure for any practical application of EIT and ATS. The depth of the transparency window, H, is defined as the difference between the absorption coefficients on the peaks and at the center of the spectrum, Δp = 0 (see middle curve in figure 4). The depth H increases with Ωc in the EIT regime and plateaus at a laser-jitter-dependent maximal value in the ATS regime, as shown by the calculated curves in figure 6. Assuming a laser jitter with 1.5 MHz FWHM, we achieve good agreement between experiment and calculation. This best-fit laser linewidth agrees well with manufacturer estimates for the utilized laser systems.

4. Conclusion

We have presented measured and calculated cold-atom EIT and AT absorption spectra of a cascade three-level atom involving the $35{S}_{1/2}$ Rydberg state. The measurements show good agreement with calculations. The spectra exhibit two regimes, an EIT regime for weak Ωc and Ωp, and an ATS regime for large Ωc. While similar in appearance, the EIT and AT spectra are different in physical interpretation [9, 10]. We have defined widths γEIT and γATS that facilitates making the distinction between EIT and ATS. In the cold-atom EIT regime, Ωc, Ωp ≲ Γeg, we find ${\gamma }_{{\rm{EIT}}}=A+B(({{\rm{\Omega }}}_{c}^{2}+{{\rm{\Omega }}}_{p}^{2})/{{\rm{\Gamma }}}_{{eg}})$, where the parameters A and B depend on inhomogeneous broadening due to laser jitter. The MOT magnetic field, which is always left on, may add to the inhomogeneous broadening. In the ATS regime, Ωc ≳ Γeg, it is γATS ≈ Ωc and γEIT = γATS − C, with a constant C ≳ Γeg/2. The lower limit of C is realized in the absence of inhomogeneous broadening. An emphasis of our analysis has been to include the effects of laser jitter, saturation of the probe transition, and Doppler mismatch caused by non-zero temperature and unequal probe and coupling laser wavelengths.

Practical criteria to distinguish between EIT and ATS, as developed in this work, are valuable in a wide variety of atomic-physics, quantum-optics and quantum information applications of these schemes. For instance, the details of ATS splitting and line broadening effects are important in applications that deal with quantitative, atom-based microwave field measurements using Rydberg-EIT and microwave-coupled ATS [27, 32]. Cold-atom, narrow-linewidth EIT and ATS are helpful in improving the accuracy and resolution of the atom-based field measurements.

Acknowledgments

The work was supported by the National Key R&D Program of China (Grant No. 2017YFA0304203), the National Natural Science Foundation of China (Grants No. 61475090, and No. 61675123, and No. 61775124), Changjiang Scholars and Innovative Research Team in University of Ministry of Education of China (Grant No. IRT13076), the State Key Program of National Natural Science of China (Grant No. 11434007) and 1331 KSC. GR acknowledges support by the NSF (PHY-1506093) and BAIREN plan of Shanxi province.

Please wait… references are loading.
10.1088/1367-2630/aad153