ArticlePDF Available

Millimeter Wave Detection via Autler-Townes Splitting in Rubidium Rydberg Atoms

Authors:

Abstract and Figures

In this paper we demonstrate the detection of millimeter waves via Autler-Townes splitting in 85Rb Rydberg atoms. This method may provide an independent, atom-based, SI-traceable method for measuring mm-wave electric fields, which addresses a gap in current calibration techniques in the mm-wave regime. The electric- field amplitude within a rubidium vapor cell in the WR-10 waveguide band is measured for frequencies of 93 GHz, and 104 GHz. Relevant aspects of Autler-Townes splitting originating from a four-level electromagnetically induced transparency scheme are discussed. We measure the E-field generated by an open-ended waveguide using this technique. Experimental results are compared to a full-wave finite element simulation.
Content may be subject to copyright.
Millimeter Wave Detection via Autler-Townes Splitting in Rubidium Rydberg
Atomsa)
Joshua A. Gordon,1, b) Christopher L. Holloway,1Andrew Schwarzkopf,2Dave A. Anderson,2Stephanie Miller,2
Nithiwadee Thaicharoen,2and Georg Raithel2
1)National Institute of Standards and Technology (NIST), Electromagnetics Division, U.S. Department of Commerce,
Boulder Laboratories, Boulder, CO 80305
2)Department of Physics, University of Michigan, Ann Arbor, MI 48109
(Dated: 16 June 2014)
In this paper we demonstrate the detection of millimeter waves via Autler-Townes splitting in 85Rb Rydberg
atoms. This method may provide an independent, atom-based, SI-traceable method for measuring mm-wave
electric fields, which addresses a gap in current calibration techniques in the mm-wave regime. The electric-
field amplitude within a rubidium vapor cell in the WR-10 wave guide band is measured for frequencies of
93 GHz, and 104 GHz. Relevant aspects of Autler-Townes splitting originating from a four-level electromag-
netically induced transparency scheme are discussed. We measure the E-field generated by an open-ended
waveguide using this technique. Experimental results are compared to a full-wave finite element simulation.
The detection of millimeter waves (mm-waves) has
proven useful for a broad range of applications, includ-
ing weapons stand-off detection1, aeronautics2, remote
sensing3, next generation wireless communications4and,
stand-off human vital sign monitoring5to mention a few.
Each of these applications may be appropriately suited
to one or more of the myriad of sensor types available.
Traceability for both electric field and power measure-
ments at these frequencies is through power. Typical
sensors are Schottky diodes, bolometers, and calorime-
ters, all of which are traceable to calorimeter measure-
ments. Calorimeter and bolometer measurements are
traceable to DC voltage and resistance measurements6.
The sensors and the calorimeters measure power at a ref-
erence plane (often a connector) of a rectangular wave
guide. If a direct measurement of the electric field is
desired, then models and or measurements, such as near-
field antenna pattern techniques7, are needed to ob-
tain the relationship between the desired electric field
and the power at the reference plane. However, spec-
ifying reliable models and performing antenna pattern
measurements becomes difficult at mm-wave frequencies
because the mechanical tolerances and repeatability of
such components may vary such that they are a signif-
icant fraction of the operating wavelength. For these
reasons we are investigating more direct and traceable
techniques for mm-wave electric field and power calibra-
tions. Here we present on an atomic-based technique
which allows direct measurement of the magnitude of the
electric field, |E|, at mm-wave frequencies via Autler-
Townes (AT) splitting in Rydberg atoms. This split-
ting is inversely proportional to Planck’s constant, ~pro-
viding a link to the SI. We obtain data for a range of
electric field levels in the WR-10 band (75-110 GHz) at
a)This work was partially supported by DARPA’s QuASAR pro-
gram. Publication of the U.S. government, not subject to U.S.
copyright.
b)Electronic mail: josh.gordon@nist.gov
93.71 GHz and 104.77 GHz. Fundamentally this tech-
nique also lends itself to measurements beyond 110 GHz,
which may address a current lack of traceable methods
for calibrating mm-wave systems above 110 GHz. Fur-
thermore, it does not rely on a priori knowledge of an
antenna pattern for determining the electric field. In ad-
dition this technique has many novel properties useful
for measurements of |E|that we have recently reported
on, such as extremely large bandwidth8(1-500 GHz),
sub wavelength imaging9and two-photon AT interac-
tions at microwave frequencies for potential use in high
power microwave sensing10
Rydberg atoms have a single valence electron in a
highly excited state, where the principal quantum num-
ber is typically n>10. The dipole moment, , in Ryd-
berg atoms scales as n2, and at the nrequired for a mm-
wave transition (n30), it can be several orders of mag-
nitude greater than for a ground state atom 1000ea0,
where eis the electron charge and a0is the Bohr radius.
Therefore Rydberg atoms can have significant response
to mm-wave electric fields. In this paper we will focus on
the 85Rb isotope of rubidium excited to n= 28,29, be-
tween the nD5/2to (n+ 1)P3/2manifolds, corresponding
to Rydberg transitions in the WR-10 mm-wave band. As
has been well established in the literature11 the energy,
W(n), of Rydberg states may be modeled as,
W(n) = RRb
(n)2,(1)
where RRb is the Rydberg constant for the reduced elec-
tron mass in rubidium, n=nδis the effective prin-
ciple quantum number determined using the quantum
defect11,13,δ. The quantum defects from13 were used
to determine the specific mm-wave frequencies for each
transition.
Although a thorough discussion of electromagnetic in-
duced transparency (EIT) is beyond the scope of this
paper, a brief description of the phenomena is give. In
a gas of rubidium atoms an incident probe laser beam
arXiv:1406.2936v1 [physics.atom-ph] 11 Jun 2014
2
FIG. 1. Energy levels used in the experiment for generating
Autler-Townes splitting from 85Rb Rydberg atoms.
experiences large absorption when tuned to the D2tran-
sition, at λp= 780.241 nm. However, in the presence
of a second coupling laser at, λc480 nm this gas
will be rendered partially transparent to the probe laser
and a transmission peak will result in the spectrum of
the probe laser. This quantum interference effect be-
tween the ground state and states excited by the probe
and coupling laser is known as Electromagnetic Induced
Transparency14. This has been widely studied for both
Rydberg atoms as well as alkali atoms at lower nand was
first demonstrated by Boiler et. al. . Later the effects of
adding a fourth level to the EIT scheme (see Figure 1)
were theoretically investigated18. In this scheme the the
transition to this fourth level is taken to be a radio fre-
quency (RF) transition of either a hyper fine transition or
a Rydberg state transition18. In the case we present here,
the transition to the fourth level is a mm-wave Rydberg
transition. To be clear, in keeping with popular nomen-
clature in the literature, the use of the term RF will be
used interchangeably to mean mm-wave frequencies in
the rest of this paper. When Rydberg atoms are used in
a four level system, the strength of the RF transition at
modest electric field strengths is sufficient to transition
this four level system from the EIT regime into the AT
regimes15,16 where the RF electric field causes the EIT
peak to split into two peaks.
With the coupling laser on resonance, and the probe
frequency swept, the EIT peak is observed to split into
two equal peaks separated by the Rabi frequency, ΩRF ,
of the RF transition,
RF =RF |ERF |
~(2)
In actuality the frequency splitting, ∆fprobe, that is
measured on the probe laser EIT spectrum, must be
scaled by the ratio of laser wavelengths. This is in or-
der to take into account the effects of Doppler mismatch,
which occurs due to the different wavelengths of the
counter propagating probe and coupling laser beams in-
teracting with the thermal vapor (room temperature) of
atoms17. The measured splitting on the probe laser is
FIG. 2. Experimental setup for generating EIT and AT spec-
tra. Counter-propagating 780 nm probe laser (dotted line)
and 480 nm coupling beam (solid line) are shown, as well as
the Rb vapor cell, dichroic beam splitter and bandpass filter.
thus related to ΩRF in terms of the wavelengths of the
coupling laser, λc, and probe laser, λp, by
fprobe =λc
λp
RF
2π(3)
From (2) we see that this splitting is linearly pro-
portional to the RF electric field strength, |ERF |, the
dipole matrix element, RF of the Rydberg RF transi-
tion, and ~. This direct relationship of the measured
Rabi frequency to the electric field, the dipole matrix el-
ement and Planck0s constant is at the heart of the trace-
ability of this technique. Sedlacek et. al.19 used this
technique for the 53D5/254P3/2Rydberg transition in
87Rb to measure the electric field strength at 14.23 GHz
inside a vapor cell. In this paper we extend this technique
for measuring electric fields in the mm-wave regime.
Frequencies in the WR-10 band of f0= 93.71 GHz, and
f0= 104.77 GHz corresponding to the, 29D5/230P3/2,
and 28D5/229P3/2transitions respectively, are mea-
sured over a range of electric field strengths. Data
are presented comparing the electric field determined
via this AT splitting technique to numerical simulations
performed using a three dimensional finite element ap-
proach.
Our experimental setup is shown in Figure 2.
Two counter-propagating lasers beams were used, the
probe laser tuned to the D2 transition of 85Rb at
780.241 nm and the coupling laser tuned to approxi-
mately 480 nm for exciting Rydberg states are incident
on the room temperature vapor cell. This is depicted in
Figure 2. The full-width half-max beam diameters at the
center of the vapor cell for the probe and the coupling
laser are 80 µm and 100 µm respectively. The beam pow-
ers were nominally 28 mW for the coupling laser and 100
nW for the probe laser. The line widths for both probe
and coupling lasers were 1 MHz.The probe laser does
not need to be broadband tunable because it is always
probing the same transition (i.e. 85 Rb D2 line). For the
coupling laser it is advantageous to have broadband tun-
ability over a range of at least several hundred GHz to
be able to optically excite a selection of Rydberg levels.
The mm-waves were produced using an RF signal
3
generator (SigGen) with 0.1 Hz resolution to drive a
WR-10 6x frequency multiplier. The output of the fre-
quency converter was coupled to a WR-10 open ended
rectangular wave guide (OEG) placed approximately 140
mm from the vapor cell. The cell is a 25 mm x 75 mm
hollow glass cylinder containing 85Rb vapor commonly
used in saturation absorption spectroscopy. The vapor
cell was mounted on a low permittivity foam block to iso-
late mm-wave scattering from surrounding metal optics
mounts and microwave absorber was used to cover ex-
posed surfaces of the optics bench. A variable in-line at-
tenuator was used to vary the mm-wave power. The mm-
wave power was verified for each dial position of the vari-
able attenuator using a WR-10 power meter connected
directly to the output of the OEG. Because the attenu-
ator uses a mechanical vane to achieve attenuation, the
equal dial settings did not necessarily correspond to equal
steps in mm-wave power. Therefore, the power output
from the OEG, POEG , was calibrated using the WR-10
power meter for each frequency, and at each increment on
the variable attenuator. The reflection coefficient, |S11|,
between the OEG aperture and free space was measured
on a vector network analyzer. Since the power meter is
impedance matched to the OEG, the power actually leav-
ing the OEG when coupled to free space is determined
by modifying the power meter reading by (1 |S11 |2) so
as to take in account the aperture reflection missing in
the matched power meter reading.
The full range of the variable attenuation was used,
however the mm-wave power range at each frequency
was not the same because the power produced by the
mixer decreased as the frequency increases. Therefore the
power range at 93.71 GHz is larger than at 104.77 GHz.
This results in a variation of achievable dynamic range
between the mm-wave frequencies. The maximum power
measured at the OEG aperture was -0.83 dBm at 104.77
GHz, and +1.95 dBm at 93.71 GHz. The minimum power
measured at the OEG aperture which gave unambigu-
ous AT splitting was -11.58 dBm and -12.71 dBm, at
104.77 GHz and 93.71 GHz respectively. For each level
of POEG the AT signal was measured on the probe laser
using a silicon photodiode and lock-in amplifier. The
probe laser was separated from the coupling laser us-
ing a dichroic beam splitter followed by a 10 nm wide
line filter in front of the photo diode see Figure 2. The
lock-in signal was generated by chopping the coupling
laser beam using an acousto-optic modulator to produce
a 30 KHz square wave modulation. With the probe laser
sweeping across the Doppler spectrum of the D2 tran-
sition, the coupling laser was tuned to the wavelength
for the desired Rydberg state. The coupling laser wave-
length was determined using the 85Rb ionization energy
and D2 transition energy from12, and the calculated Ry-
dberg state energy using (1), fine-tuning was done by
observing the three-level EIT signature (with mm-wave
power off). Once the EIT signal was established the cou-
pling laser was locked to a stabilizing cavity.
To produce the AT signal, the SigGen was tuned to
FIG. 3. EIT peak with mm-wave power off and and AT split-
ting for POEG =2.43 dBm for the 28D5/229P3/2transi-
tion, f0= 104.77 GHz.
the mm-wave frequency that was determined again using
(1) for the desired Rydberg transition. With the variable
attenuator set to half of the power range, the mm-wave
frequency was fine-tuned so as to result in AT peaks of
equal heights. Both the laser fields and mm-waves were
(linear) π-polarized and aligned so as to minimize excita-
tion of the 3-level EIT pathway21. The mm-wave power
was then varied using equal increment dial settings on
the attenuator. The Doppler-free saturation absorption
spectrum of the probe laser was obtained simultaneously
with the AT spectrum. This was used to calibrate the
measured AT splitting from the known frequency spac-
ing of the Doppler-free hyperfine features present in the
D2 saturation spectrum. Equation(2) was then used to
determine the ΩRF , where the dipole moment RF , was
calculated using the methods described in11 and the ap-
propriate Clebsch-Gordan coefficients. Figure 3 shows
scans taken for the 28D5/229P3/2transition for the
EIT signal with mm-wave power off, along with the AT
signal with the power at the OEG set to -2.43 dBm.
Electric fields were using the experimental setup de-
scribed above. For each power setting of the vari-
able attenator the splitting was calculated using (2)
and (3) and compared to simulated results. At a
distance of 140 mm from the aperture of the OEG,
the vapor cell was well beyond the farthest far-
field distance calculated for the mm-wave frequencies
that were measured (i.e. 5.63 mm, the value at
104.77 GHz). The far-field distance was calculated us-
ing the dimension of the OEG aperture diagonal and the
conventional definition given in20. Simulations were per-
formed to determine the electric field radiated by the
OEG using the far-field calculator in the electromagnetic
4
FIG. 4. AT splitting in MHz versus POEG for the 29D5/2
30P3/2,f0= 93.71 GHz, and 28D5/229P3/2f0= 104.77
GHz transitions. Linear fits are shown as solid lines.
finite element solver HFSS (mention of this software is
not an endorsement but is only intended to clarify what
was done in this work). These field values were then used
to compare with those measured using the vapor cell.
First, we established that the measured splitting scales
linearly with the electric field as expected from (2). Given
that the electric field at the vapor cell is proportional to
the POEG , if the splitting indeed follows the behavior
in (2), then a linear relationship would be apparent by
plotting ΩRF versus POEG. This is clearly shown in
Figure 4 for both frequencies. Figures 5 and 6 show a
comparison of the electric field values determined from
the vapor cell measurements to those obtained from the
HFSS far-field simulation. The error bars in these plots
show the expected range of electric field values within
the vapor cell due to field variations that are present be-
cause of the dielectric boundary of the cell. We discuss
this further next.
From Figures 5 and 6, we see that there is a
noticeable difference between the measured and simu-
lated electric field values. Also, the agreement of the
measured electric field to simulated results is not con-
sistent between frequencies. A strongly observable effect
which alters the electric field inside the vapor cell results
from standing waves set up by the dielectric boundary of
the cell walls interacting with the mm-waves. The dimen-
sions of the vapor cell used are 25 mm x 75 mm and
the operating wavelengths are 3.22 mm and 2.88 mm at
93.71 GHz and 104.77 GHz respectively. Therefore, the
vapor cell, in terms of wavelengths is 7.7λx 23.4λ
and 8.7λx 26.3λfor these two cases. Since the laser
beams were not moved between measuring the two mm-
wave frequencies, the observed frequency dependence is
attributed to the difference in mm-wave mode structure
of the vapor cell as a result of the difference in wave-
FIG. 5. Electric field values at 93.71 GHz measured using the
vapor cell and compared to HFSS far-field simulation. Solid
line shows HFSS simulation. Error bars indicate the possible
20% field variation range due to resonant mm-wave scattering
effects of the vapor cell.
FIG. 6. Electric field values at 104.77 GHz measured using the
vapor cell and compared to HFSS far-field simulation. Solid
line shows HFSS simulation. Error bars indicate the possible
%20 field variation range due to resonant mm-wave scattering
effects of the vapor cell.
length between the 93.71 GHz and 104.77 GHz frequen-
cies. As the wavelength changes, the standing wave struc-
ture changes, and thus the electric field amplitude at the
location of the laser beams in the vapor cell will depend
on the mm-wave frequency. We have reported on this
5
effect in detail in9, where we show the ability to image
these standing waves at extreme sub-wavelength resolu-
tion using AT splitting at both microwave (17.04 GHz)
and mm-wave (104.77 GHz) frequencies. From imaging
these standing waves we determined a ±20% variation
about the mean field strength as a function of measure-
ment location in the vapor cell at 104.77 GHz. The error
bars in Figures 5 and 6 indicate this ±20% variation for
the measurements we present here. This perturbing ef-
fect of the electric field by the presence of the dielectric
vapor cell is something we are currently addressing.
In this paper we demonstrate the detection of mil-
limeter waves via Autler-Townes splitting in 85Rb
Rydberg atoms. This method may provide an indepen-
dent, atomic-based, SI-traceable method for measuring
mm-wave electric fields, which addresses a gap in current
calibration techniques in the mm-wave regime. The elec-
tric field amplitude within a rubidium vapor cell in the
WR-10 waveguide band was measured for frequencies of
93.71 GHz, and 104.77 GHz. Experimental results are
presented where we measure the far-field electric field
generated by an open ended waveguide using this tech-
nique. A comparison to far-field electric field values ob-
tained from a finite element simulation is made. The
experimentally observed scaling behavior follows closely
the expected linear behavior of Autler-Townes splitting.
The electric fields measured agree to within ±20% of the
far-field simulations due to standing wave effects.
I. ACKNOWLEDGMENTS
Special thanks to David R. Novotny, and Galen H.
Koepke of the Electromagnetics Division at NIST,
Boulder for assistance with equipment.
1R. Appleby, Proc. IEEE 95, 1683, (2007)
2A. H. Lettington, Optical Engineering 44(9), 093202-1 (2005)
3P.H. Siegel, IEEE Trans.Micr. Ther. Tech, 50, 910, (2002)
4Z. P. and F. Khan, Samsung Electronics, IEEE Comm. Mag. 101
(2011)
5S. Bakhtiari, IEEE Trans. Inst. and Mease, 61, 830, (2012)
6A. Fantom, ”Radio Frequency and Microwave Power Measur-
ment”, Peter Peregrinus, Ltd, London, 1990
7M. H. Francis, R. C. Wittmann, in ”Modern Antenna Hand
Book”, C. A. Balanis, Chp. 19, 929, (2008)
8C. L. Holloway, J. A. Gordon, S. Jefferts, A. Schwarzkopf, D. A.
Anderson, S. A. Miller, N. Thaicharoen,G. Raithel, IEEE Trans.
Ant. Prop. Soc. (2014)
9C. L. Holloway, J. A. Gordon, A. Schwarzkopf, D. A. Anderson,
S. A. Miller, N. Thaicharoen,G. Raithel, Appl. Phys. Lett, (2014)
10D.A. Anderson, S.A. Miller, A. Schwarzkopf, C.L. Holloway, J.A.
Gordon, N. Thaicharoen, and G. Raithelet, Two-photon tran-
sitions and strong-field effects in Rydberg atoms via EIT-AT,
submitted 2014.
11T.F. Gallagher, Rydberg Atoms, Cambridge University Press
(1994)
12D. Steck, Rubidium 87/85 D-Line Data,
http://steck.us/alkalidata
13W. Li, I. Mourachko, M.W. Noel, and T. F. Gallagher, Phys.
Rev. A, 67, 052502, (2003)
14K. J. Boller, A. Imamoglu, S. E. Harris, Phys. Rev. Lett., 66,
2593 (1991)
15S. H. Autler and C. H. Townes, Phys. Rev 100, 703, (1955)
16T. Y. Abi-Salloum, Phys. Rev. A, 81, 053836 (2010)
17A. K. Mohapatra, T. R. Jackson, and C. S. Adams, Phys. Rev.
Lett. 98, 113003 (2007)
18S. N. Sandhya and K. K. Sharma, Phys. Rev. A, 55, 2155 (1997)
19J. A. Sedlacek, A. Schwettmann, H Kubler, R. Low, T. Pfau, J.
P. Shaffer, Nature Phys., 8, 819, (2012)
20IEEE Standard Definitions of Terms for Antennas, IEEE Std
145-1993, pp. 13
21J. A. Sedlacek, A. Schwettmann, H. Kubler, and J. P. Shaffer,
Phys. Rev. Lett., 111, 063001 (2013)
... The Rydberg atomic antennas are devices that measure physical quantities utilizing the quantum properties of microscopic particles [1][2][3][4][5], offering great advantages in high sensitivity, small system size, and concealed anti-damage detection [3,6,7]. Particularly, it has been theoretically demonstrated that the electric field measurement sensitivity limit is -220 dBm/Hz [8], which is far exceeding the classical receiver sensitivity limit of -174 dBm/Hz. ...
... As the resonance-based system is more sensitive to errors, this paper take all the cell walls refractive index and metal reflections into account. The issues of metal reflections, space scattering disturbance can appear multi-reflections inside the cell and stand- Figure 9 Results from the Rydberg-atom superheterodyne, plots of the beat-note intensity of the spectrum analyzer as functions of signal generator, the error bars representing the standard deviation of sensitivity ing waves disturbance(or resonances) [2]. The E-field strength in space can be expressed as [53]: ...
... where c is the light speed in vacuum, ε 0 is the permittivity of free space, R = 0.05 m is the distance from antenna to the laser beam, F 0 is the perturbation factor caused by space scattering and standing wave(or resonances) disturbance in the cell, P SIG G T L is the radiated power of microwave (P SIG G T L = P SIG + G T -L, P SIG represents the output power of signal source, G T = 11 dB represents the gain of antenna, L = -1.5 dB represents the insertion loss of transmission line). The parameter F 0 which can be determined numerically or experimentally [2,54] is estimated F 0 ≈ 0.411 for 2 GHz according to the electric-field. As shown in Fig. 9, the beat-note intensity from the spectrum analyzer is a function of signal generator for the cases with resonator in blue curve and without in black curve. ...
Article
Full-text available
Due to its large electric dipole moment, the Rydberg atom exhibits a strong response to weak electric fields, hence it is regarded as a highly promising atomic antenna. However, to enhance the reception sensitivity, split-ring resonators are needed normally, which will brings sensing blind spots. Thus it is not conducive to the application of full-coverage space communication. Here we propose that an atomic antenna with an asymmetric parallel-plate resonator, can not only enhance the received signal, but also eliminate sensing blind spots (pattern roundness can reach 7.8 dB while the split-ring resonator can be up to 39 dB). We analyze the influence of structural parameters on the field enhancement factor and directionality, and further discuss the limitation of the sensitivity by using thermal resistor noise theory. This work is expected to pave the way for the development of field-enhanced Rydberg atomic antennas that communicate without a blind spot.
... The Rydberg atomic antennas are devices that measure physical quantities utilizing the quantum properties of microscopic particles [1,2,3,4], offering great advantages in high sensitivity, small system size, and concealed anti-damage detection [5,6,3]. Particularly, it has been theoretically demonstrated that the electric field measurement sensitivity limit is −220 dBm/Hz [7], which is far exceeding the classical receiver sensitivity limit of −174 dBm/Hz. ...
... As the resonance-based system is more sensitive to errors, this paper take all the cell walls refractive index and metal reflections into account. The issues of metal reflections, space scattering disturbance can appear multireflections inside the cell and standing waves disturbance(or resonances) [2]. The E-field strength in space can be expressed as [49]: ...
... where c is the light speed in vacuum, ε 0 is the permittivity of free space, R = 0.05 m is the distance from antenna to the laser beam, F 0 is the perturbation factor caused by space scattering and standing wave(or resonances) disturbance in the cell, P SIG G T L is the radiated power of microwave (P SIG G T L = P SIG + G T -L, P SIG represents the output power of signal source, G T = 11 dB represents the gain of antenna, L = -1.5 dB represents the insertion loss of transmission line). The parameter F 0 which can be determined numerically or experimentally [50,2] is estimated F 0 ≈ 0.411 for 2 GHz according to the electric-field. As shown in Fig. 9, the beat-note intensity from the spectrum analyzer is a function of signal generator for the cases with resonator in blue curve and without in black curve. ...
Preprint
Full-text available
Due to its large electric dipole moment, the Rydberg atom exhibits a strong response to weak electric fields, hence it is regarded as a highly promising atomic antenna. However, to enhance the reception sensitivity, split-ring resonators are needed normally, which will brings sensing blind spots. Thus it is not conducive to the application of full-coverage space communication. Here we propose that an atomic antenna with an asymmetric parallel-plate resonator, can not only enhance the received signal, but also eliminate sensing blind spots (pattern roundness can reach 7.8 dB while the split-ring resonator can be up to 39 dB). We analyze the influence of structural parameters on the field enhancement factor and directionality, and further discuss the limitation of the sensitivity by using thermal resistor noise theory.This work is expected to pave the way for the development of field-enhanced Rydberg atomic antennas that communicate without a blind spot.
Article
Определены условия расщепления резонанса электромагнитно-индуцированной прозрачности (ЭИП) полем микроволнового излучения и рассчитаны частоты и амплитуды радиационных переходов между ридберговскими состояниями атомов щелочноземельных элементов группы IIa, необходимые для прецизионных измерений напряженности сверхвысокочастотного (СВЧ) электрического поля. Численные значения частот и матричных элементов аппроксимированы асимптотическими полиномами и табулированы для дипольных переходов между синглетными nS-, nP-, nD- и nF-состояниями с большими значениями главных квантовых чисел n.
Article
Rydberg quantum sensors are sensitive to radio-frequency fields across an ultra-wide frequency range spanning megahertz to terahertz electromagnetic waves resonant with Rydberg atom dipole transitions. Here, we demonstrate an atomic millimeter-wave heterodyne receiver employing continuous-wave lasers stabilized to an optical frequency comb. We characterize the atomic receiver in the W-band at a signal frequency of f = 95.992 512 GHz and demonstrate a sensitivity of 7.9 μV/m/Hz with a linear dynamic range in power greater than 70 dB. We develop frequency selectivity metrics for atomic receivers and demonstrate their use in our millimeter-wave receiver, including signal rejection levels at signal frequency offsets Δf/f = 10−4, 10−5, and 10−6; 3, 6, 9, and 12-dB bandwidths; filter roll-off; and shape factor analysis. Our work represents an important advance toward future studies and applications of atomic receiver science and technology in weak millimeter-wave and high-frequency signal detection.
Article
We report on high-resolution microwave spectroscopy of cesium Rydberg (n+2)D5/2→nFJ transitions in a cold atomic gas. Atoms laser-cooled and trapped in a magnetic-optical trap are prepared in the D Rydberg state using a two-step laser excitation scheme. A microwave field transmitted into the chamber with a microwave horn drives the Rydberg transitions, which are probed via state selective field ionization. Varying duration and power of the microwave pulse, we observe Fourier sideband spectra as well as damped, on-resonant Rabi oscillations with pulse areas up to ≳3π. Furthermore, we investigate the Zeeman effect of the clearly resolved nFJ fine-structure levels in fields up to 120 mG, where the transition into nF7/2 displays a thee-peak Zeeman pattern, while nF5/2 shows a two-peak pattern. Our theoretical models explain all observed spectral characteristics, showing good agreement with the experiment. Our measurements provide a pathway for the study of high-angular-momentum Rydberg states, initialization and coherent manipulation of such states, Rydberg-atom macrodimers, and other Rydberg-atom interactions. Furthermore, the presented methods are suitable for calibration of microwave radiation as well as for nulling and calibration of dc magnetic fields in experimental chambers for cold atoms.
Article
We calculate, via variational techniques, single- and two-photon Rydberg microwave transitions, as well as scalar and tensor polarizabilities of the sodium atom using the parametric one-electron valence potential, including the spin-orbit coupling. The trial function is expanded in a basis set of optimized Slater-type orbitals, resulting in highly accurate and converged eigenenergies up to n=60. We focus our studies on the microwave band 90–150 GHz due to its relevance to laser excitation in the Earth's upper-atmospheric sodium layer for wavelength-dependent radiometry and polarimetry, as precise microwave polarimetry in this band is an important source of systematic uncertainty in searches for signatures of primordial gravitational waves within the anisotropic polarization pattern of photons from the cosmic microwave background. We present the most efficient transition coefficients in this range, as well as the scalar and tensor polarizabilities compared with available experimental and theoretical data.
Article
We demonstrate a method for broadband tunable continuous frequency electric field measurement based on the DC Stark effect in Rydberg atoms. In our experiment, we place a pair of parallel electrode plates inside the atomic vapor cell, utilizing the DC Stark effect to induce splitting and shifting of the Rydberg energy levels, thereby altering the resonance frequency of the Stark subpeaks. By employing the 52D5/2 Rydberg state, we achieve electric field measurements in the frequency range of 5.083–14.470 GHz. At an EDC of 3.45 V/cm and a resonant microwave frequency of 14.470 GHz, using heterodyne technology, the microwave electric field sensitivity is 538.89 μV/cm/√Hz, with a linear dynamic range of 23 dB. In comparison, a Rydberg heterodyne receiver with an EDC of 0 V/cm and a resonant microwave frequency of 5.083 GHz has a sensitivity of 5.43 μV/cm/√Hz and a linear dynamic range of 51 dB. This work will promote the study of atomic microwave receivers in continuous microwave frequency measurement.
Article
Full-text available
A compact millimeter-wave (MMW) sensor has been developed for remote monitoring of human vital signs (heart and respiration rate). The low-power homodyne transceiver operating at 94 GHz was assembled by using solid-state active and passive block-type components and can be battery operated. A description of the MMW system front end and the back-end acquisition hardware and software is presented. Representative test case results on the application of various signal processing and data analysis algorithms developed to extract faint physiological signals of interest in presence of strong background interference are provided. Although the laboratory experiments so far have been limited to standoff distances of up to 15 m, the upper limit of the detection range is expected to be higher. In comparison with its microwave counterparts, the MMW system described here provides higher directivity, increased sensitivity, and longer detection range for measuring subtle mechanical displacements associated with heart and respiration functions. The system may be adapted for use in a wide range of standoff sensing applications including for patient health care, structural health monitoring, nondestructive testing, biometric sensing, and remote vibrometry in general.
Article
Full-text available
It is clearly important to pursue atomic standards for quantities like electromagnetic fields, time, length and gravity. We have recently shown, using Rydberg states, that Rb atoms in a vapor cell can serve as a practical, compact standard for microwave electric field strength. Here, we demonstrate, for the first time, that Rb atoms excited in a vapor cell can also be used for vector microwave electrometry by using Rydberg atom electromagnetically induced transparency. We describe the measurements necessary to obtain an arbitrary microwave electric field polarization at a resolution of $0.5^\circ$. The experiments are compared to theory and found to be in excellent agreement.
Book
In recent years, Rydberg atoms have been the subject of intense study, becoming the testing ground for several quantum mechanical problems. This book provides a comprehensive description of the physics of Rydberg atoms, highlighting their remarkable properties by reference to their behaviour in a wide range of physical situations. Following an overview of the basic properties of Rydberg atoms, their interactions with electric and magnetic fields are analysed in detail. The collisions of Rydberg atoms with neutral and charged species are described, and the use of multichannel quantum defect theory in the study of Rydberg atomic systems is discussed. Experimental and theoretical research in this extensive field is also reviewed, making the book valuable to both graduate students and established researchers in physics and physical chemistry.
Book
Modern Antenna Handbook represents the most current and complete thinking in the field of antennas. The handbook is edited by one of the most recognizable, prominent, and prolific authors, educators, and researchers on antennas and electromagnetics. Each chapter is authored by one or more leading international experts and includes cover-age of current and future antenna-related technology. The information is of a practical nature and is intended to be useful for researchers as well as practicing engineers. From the fundamental parameters of antennas to antennas for mobile wireless communications and medical applications, Modern Antenna Handbook covers everything professional engineers, consultants, researchers, and students need to know about the recent developments and the future direction of this fast-paced field. In addition to antenna topics, the handbook also covers modern technologies such as metamaterials, microelectromechanical systems (MEMS), frequency selective surfaces (FSS), and radar cross sections (RCS) and their applications to antennas, while five chapters are devoted to advanced numerical/computational methods targeted primarily for the analysis and design of antennas.
Article
We present a technique for measuring radio-frequency (RF) electric field strengths with sub-wavelength resolution. We use Rydberg states of rubidium atoms to probe the RF field. The RF field causes an energy splitting of the Rydberg states via the Autler-Townes effect, and we detect the splitting via electromagnetically induced transparency (EIT). We use this technique to measure the electric field distribution inside a glass cylinder with applied RF fields at 17.04 GHz and 104.77 GHz. We achieve a spatial resolution of ≈100 μm, limited by the widths of the laser beams utilized for the EIT spectroscopy. We numerically simulate the fields in the glass cylinder and find good agreement with the measured fields. Our results suggest that this technique could be applied to image fields on a small spatial scale over a large range of frequencies, up into the sub-terahertz regime.
Article
Atom-based standards for length and time as well as other physical quantities such as magnetic fields show clear advantages by enabling stable and uniform measurements. Here we demonstrate a new method for measuring microwave (MW) electric fields based on quantum interference in a rubidium atom. Using a bright resonance prepared within an electromagnetically induced transparency window we could achieve a sensitivity of ~30μVcm-1Hz-1/2 and demonstrate detection of MW electric fields as small as ~8μVcm-1 with a modest set-up. The sensitivity is limited, at present, by the stability of our lasers and can be significantly improved in the future. Our method can serve as a new atom-based traceable standard for MW electrometry, with its reproducibility, accuracy and stability promising advances towards levels comparable with those attained in magnetometry at present.
Article
We report here an effect in a four-level ladderlike system, which is in contrast to the usual quantum interference effects such as electromagnetically induced transperency (EIT) or coherent population trapping: we predict the occurrence of a narrow absorption peak within the EIT window when an EIT atomic system interacts with an additional driving rf field. The Doppler-free-central absorption appears when the three-photon resonance condition is satisfied. In the limit of the rf field strength Ωrf→0, the usual EIT profile is recovered.
Article
Electromagnetically induced transparency (EIT) and Autler-Townes (AT) splitting are two phenomena that could be featured in a variety of three-level atomic systems. The considered phenomena, EIT and AT, are similar “looking” in the sense that they are both characterized by a reduction in absorption of a weak field in the presence of a stronger field. In this paper, we explicitly set the threshold of separation between EIT and AT splitting in a unified study of four different three-level atomic systems. Two resonances are studied and compared in each case. A comparison of the magnitudes of the resonances reveals two coupling-field regimes and two categories of three-level system.
Article
By using a magneto-optical trap we have measured the Rb ns-(n+1)s and ndj-(n+1)dj two-photon millimeter-wave transitions for 32<~n<~37, observing 100-kHz-wide resonances, in spite of the trap’s 10 G/cm magnetic-field gradient, in which one might expect to observe resonances 5 MHz wide. This resolution is possible because of the similarity of the gj factors in the initial and final states. Under the same conditions, the single-photon ns-np resonances are ∼5 MHz wide. To make useful measurements of these intervals, we turned off the trap field and used the 300-K atoms of the background Rb vapor. Together these measurements improve the accuracy of the s, p, and d quantum defects by an order of magnitude.