The rapid progress in quantum information processing leads to a rising demand for devices to control the propagation of electromagnetic wave pulses and to ultimately realize universal and efficient quantum memory. While in recent years, significant progress has been made to realize slow light and quantum memories with atoms at optical frequencies, superconducting circuits in the microwave domain still lack such devices. Here, we demonstrate slowing down electromagnetic waves in a superconducting metamaterial composed of eight qubits coupled to a common waveguide, forming a waveguide quantum electrodynamics system. We analyze two complementary approaches, one relying on dressed states of the Autler–Townes splitting and the other based on a tailored dispersion profile using the qubits tunability. Our time-resolved experiments show reduced group velocities of down to a factor of about 1500 smaller than in vacuum. Depending on the method used, the speed of light can be controlled with an additional microwave tone or an effective qubit detuning. Our findings demonstrate high flexibility of superconducting circuits to realize custom band structures and open the door to microwave dispersion engineering in the quantum regime.

In the light of quantum information processing achieving control over the speed of light in artificially structured media, slowing it down and eventually stopping the light, has recently regained attention. This functionality is of vital importance for the realization of long-living quantum memories1 and for controlling and synchronizing the flow of information.2 One of the most prominent techniques to realize slow light is based on electromagnetically induced transparency (EIT)—a quantum interference effect between different excitation pathways, rendering an otherwise opaque medium transparent and creating a steep dispersion profile.3 Early demonstrations of slow light in EIT media involved vapors of sodium4 and rubidium atoms,5 followed by further experiments with cold atoms coupled to nanofibers,6 and eventually by the demonstration of a highly efficient quantum memory.7 A promising candidate beyond atoms to observe these effects is superconducting qubits in the context of cavity-free waveguide quantum electrodynamics (wQED). The observation of resonance florescence revealed that even single qubits can have strong coupling to propagate electromagnetic waves and show extinction coefficients close to unity.8,9 First demonstrations of the single-qubit Autler–Townes splitting (ATS) with ladder-type three-level systems,9,10 showcased their potential suitability for the EIT-related applications. However, it was subsequently argued that in these experiments, the EIT regime was not unambiguously reached.11 

More recently, focus shifted toward multi-qubit systems with infinite range interactions,12,13 super- and subradiant polaritonic excitations,13–16 collective ATS,14 topological edge states,17,18 creation of non-classical light,19 proposals for superconducting quantum memories,20 and quantum metameterials.21,22 Even though superconducting wQED systems can match the requirements of large optical depth and high coherence,23 slow light has been so far realized only in the context of classical waveguides.24 These devices are passive without the possibility to in situ control the speed of light and to ultimately realize a quantum memory protocol.

Here, we experimentally demonstrate a first realization of slow light in a superconducting wQED system consisting of eight locally tunable transmon qubits coupled to a one-dimensional waveguide. We engineer the required flatband structure by using qubits directly as controllable dispersive elements. First, we consider the standard case of dressed state based slow light and demonstrate that even in the case of imperfect EIT, where the physics is merely governed by the ATS, a moderate retardation of the group velocities by a factor down to 1500 compared to vacuum can be achieved. Second, we engineer a similar band structure based on detuned collective resonances of the participating qubits. This allows for three times larger efficiencies of the slow-light medium. The demonstrated slow light effect can be used for a fixed-delay quantum memory in superconducting wQED and paves way to more general applications such as on demand storage-and-retrieval memory for quantum information processing.

The qubit metamaterial presented in this work consists of eight superconducting transmon qubits,25 capacitively coupled to the mode continuum of a coplanar waveguide, see Fig. 1(a). Each qubit has an integrated SQUID loop and is thereby individually tunable between 3 and 8 GHz by applying local magnetic flux via a flux bias line. We fulfill the metamaterial limit of sub-wavelength dimensions by choosing a dense qubit spacing of d=400μ m, which corresponds to a fraction of the light wavelength λ with the phase delay of φ=2πλd=0.050.16. As shown in former works, in such a setting the qubits obtain an infinite-range photon-mediated effective interaction, which is almost exclusively of collective dissipative nature.12,13 For a detailed study of the mode structure of this metamaterial, its super- and subradiant polaritonic excitations, we guide the reader to Ref. 14. Figure 1(b) shows the first three ladder-type energy levels of the transmon qubits. If the 21 transition is driven with a microwave control tone with the amplitude (Rabi-strength) Ωc, these levels hybridize and split into two dressed states separated by Ωc, forming the ATS.26 For our experiments, we use the ATS of the individual qubits to calibrate the absolute value of the control power Pc=Ωc24Γ10ωc27 (see the supplementary material). With reference to a weak probe tone and assuming a exp(iωt) time dependence, the reflection coefficient of a single qubit is given by10 

r=Γ102[γ10i(ωω10)]+Ωc22γ202i(ωω10+ωcω21).
(1)
FIG. 1.

(a) Microscopic image of the eight qubit metamaterial with local flux bias lines. Both signal and control tone propagate along the central waveguide. (b) Ladder-type level structure of the employed transmon qubits and the relevant radiative relaxation (Γ10,Γ21) and decoherence rates (γ20,γ10). If the control tone is resonant with the 21 transition, the first transmon level hybridizes, creating the Autler–Townes splitting. (c) Calculated band structure of the metamaterial in the lossless limit for different control Rabi-strengths Ωc. Light blue shaded areas indicate bandgaps. The effective group velocity vg in the center band strongly depends on Ωc.

FIG. 1.

(a) Microscopic image of the eight qubit metamaterial with local flux bias lines. Both signal and control tone propagate along the central waveguide. (b) Ladder-type level structure of the employed transmon qubits and the relevant radiative relaxation (Γ10,Γ21) and decoherence rates (γ20,γ10). If the control tone is resonant with the 21 transition, the first transmon level hybridizes, creating the Autler–Townes splitting. (c) Calculated band structure of the metamaterial in the lossless limit for different control Rabi-strengths Ωc. Light blue shaded areas indicate bandgaps. The effective group velocity vg in the center band strongly depends on Ωc.

Close modal

The corresponding transmission coefficient is given by t=1+r, referenced in this article as element of the scattering matrix S21(ω). Here, the qubits are strongly coupled to the waveguide, ensuring a multi-mode Purcell limited average relaxation rate of Γ10/2π=12 MHz. The average decoherence rate of the 10 transition is γ10/2π=(Γ10/2+Γnr)/2π=6.9 MHz with Γnr accounting for pure dephasing and radiative losses to unguided modes. The corresponding average decoherence rate of the 20 transition is γ20/2π=6.9 MHz, which is here by coincidence the same value. We note that this rate is in agreement with other experiments with superconducting qubits,10,28 significantly higher than that of cold atoms.3 The increased rate is due to the ladder-type level structure of the transmon, where it can be shown that γ20>Γ21/2 always holds.29 This necessarily leads to imperfect EIT since the associated dark state obtains a finite lifetime.3 The corresponding implications for the creation of slow light are discussed in the following paragraph. The expected band structure for the metamaterial in the lossless case (Γnr=0,γ20=0) based on Eq. (1) for different values of Ωc is shown in Fig. 1(c). The expected group velocity of light vg=dωdk in the center band strongly depends on the control tone strength Ωc. In the regime of small Ωc, the slope of the band is almost flat, giving rise to slowing down the electromagnetic waves in the waveguide.

In order to create a slow-light medium, the qubits are consecutively tuned to a common resonance frequency ω10/2π=7.812 GHz in the vicinity of the qubits' upper sweet spot. A continuous microwave control tone is applied, driving resonantly the 21 transition at the frequency ωc/2π=7.533 GHz and Rabi-strength Ωc. Figure 2 shows the collective ATS, for N = 7 resonant qubits. For low control powers Pc0, only a single bandgap of strongly suppressed transmission above ω10 is present, which gradually splits for higher Ωc into two separate bandgaps, with a transparency window of finite transmission opening up in between [compare band structure calculation in Fig. 1(c)]. It is noted that due to the finite system size, the observed bandgap Δω is smaller than for the infinite case Δω/2π=Γ10cω10d/2π76MHz. A more detailed discussion on finite size effects is provided in Ref. 14. In this region, the phase of the complex transmission coefficient Arg (S21(ω)) features a steep roll-off [Fig. 2(b)], indicating low group velocities. The observed transmission coefficient is in agreement with a numerical transfer-matrix calculation,14,30 which takes into account a cable resonance caused by impedance mismatches in the cryostat wiring. Due to the aforementioned large 20 decoherence rate γ20γ10, the expected sharp EIT feature at ω10 with near unity transmission for small Ωc is smeared out, and transmission is reduced. The expected effective group velocity vg of a signal sent through the transparency window can be calculated from the phase gradient20,30

vg=[1(N1)ddArg(S21(ω))dω]1|ω=ω10.
(2)
FIG. 2.

(a) Collective ATS of seven resonant qubits in dependence on control tone power Pc. For Pc>125 dBm, a transparency window with finite transmission emerges around the resonance frequency ω10 (black dashed line). The red dashed line is a fit to the expected splitting of Ωc/2π. Purple dotted lines mark a minimal bandwidth of 1/σ=20 MHz of the transparency window for time-domain experiments with σ=50 ns pulses. (b) Absolute value and phase of the complex transmission coefficient S21 for Pc=122 dBm. Black dotted lines are fits to a transfer-matrix model. The effective expected group velocity vg is anti-proportional to the slope of Arg(S21) at ω10.

FIG. 2.

(a) Collective ATS of seven resonant qubits in dependence on control tone power Pc. For Pc>125 dBm, a transparency window with finite transmission emerges around the resonance frequency ω10 (black dashed line). The red dashed line is a fit to the expected splitting of Ωc/2π. Purple dotted lines mark a minimal bandwidth of 1/σ=20 MHz of the transparency window for time-domain experiments with σ=50 ns pulses. (b) Absolute value and phase of the complex transmission coefficient S21 for Pc=122 dBm. Black dotted lines are fits to a transfer-matrix model. The effective expected group velocity vg is anti-proportional to the slope of Arg(S21) at ω10.

Close modal

The traversal time τ=(N1)d/vg of a pulse through the medium with N = 7 qubits inferred from the spectroscopic data of Fig. 2(a) at ω=ω10 is shown in Fig. 3(b). In good agreement with the numerical transfer-matrix model, the expected traversal time, or conversely, the effective group index ng, can be tuned over a large range with the applied control power Pc. In contrast to the textbook case of EIT in the limit of γ200, where τ1/Ωc2 diverges for small Ωc,3 we observe τ approaching a maximum of 15 ns (ng1900) at Pc124 dBm before it decreases again for even lower control tone strengths. A similar behavior was observed in room temperature vapor of 4He.31 An analytical expression for the expected delay τ(ω10) in the case of a resonant control tone can be found from the effective dispersion relation (kd)2φ22χφ with χ=ir1+r and φ=ωcd:32 

τ(ω10,Ωc){(N1)Γ10(2Ωc28γ202)((4γ102Γ10)γ20+Ωc2)2,φχ,(N1)Γ10(2Ωc28γ202)((4γ102Γ10)γ20+Ωc2)3/2φ8Γ10γ20,φχ.
(3)
FIG. 3.

(a) Envelope of the detected Gaussian pulses in the time domain for different control tone powers Pc of the N = 7 qubit ATS. The pulses are generated with a center frequency of ω10=7.812 GHz and width σ=50 ns, corresponding to the center of the transparency window. For better visibility, the pulses are compressed by a factor of 20, where their maximum remains at the original position. (b) Experimentally extracted delays from pulsed and spectroscopic measurements in dependence on control power Pc. Pulse delays of up to 12 ns (15 ns in spectroscopy) were achieved, corresponding to the reduction in the group velocity of 1500 (1900) compared to the speed of light in vacuum.

FIG. 3.

(a) Envelope of the detected Gaussian pulses in the time domain for different control tone powers Pc of the N = 7 qubit ATS. The pulses are generated with a center frequency of ω10=7.812 GHz and width σ=50 ns, corresponding to the center of the transparency window. For better visibility, the pulses are compressed by a factor of 20, where their maximum remains at the original position. (b) Experimentally extracted delays from pulsed and spectroscopic measurements in dependence on control power Pc. Pulse delays of up to 12 ns (15 ns in spectroscopy) were achieved, corresponding to the reduction in the group velocity of 1500 (1900) compared to the speed of light in vacuum.

Close modal

Figure 3(b) shows both asymptotes of Eq. (3), corresponding to the strong and weak Rabi-strength Ωc regime (φχ,φχ). In the limit of γ200, the condition φχ is always fulfilled, and the corresponding τ(ω10,Ωc) is reduced to the formulas of textbook EIT.3 In this case. we find good agreement between the measured delays and Eq. (3). In the opposite limit φχ, only qualitative agreement is observed. The systematic offset to the measured data is rooted in the band structure calculation of Eq. (3), which assumes an infinite metamaterial deviates, in particular, for small Ωc. More details on the derivation of τ(ω10,Ωc) are provided in the supplementary material. Formally, EIT and ATS should be treated as two distinct, but closely related phenomena: EIT is a destructive Fano-type interference effect between two excitation pathways, which can only occur if Ωc<ξγ10, smoothly transitioning to the ATS-regime for Ωc>ξγ10, with two independent resonances of the dressed state doublet.11 The coefficient ξ=(N21)φ/3 is the approximate width of the bandgap for a finite system with N<π/φ20 qubits in units of γ10. In this work, due to large γ20, the accessible control strengths are in the transient regime Ωcξγ10. Here, a quantitative method to distinguish between ATS and EIT based on Akaike's information criterion of Ref. 11 cannot directly be applied to the measured data, since it features an asymmetric line shape due to the microwave background. Using the method for the calculated single-qubit splitting based on Eq. (1) in conjunction with the experimentally extracted qubit decoherence rates yields that the observed line shape is by almost 100% certainty described by the ATS (not shown).

We have verified the spectroscopically inferred time delays τ in pulsed time-domain measurements by using an FPGA-based heterodyne microwave setup (see the supplementary material). Here, we generated Gaussian pulses of width σ=50 ns at resonant center-frequency ω10 with a continuously applied control tone at frequency ω21 and strength Ωc. The pulse amplitude is kept at the single photon level (Pp<ω10Γ10) in order to prevent saturation of the qubits. The digitized transmitted pulses for different control tone strengths are shown in Fig. 3(a). By fitting Gaussians to the measured data, we extracted the temporal position of each pulse center. The extracted pulse arrival times were compared to a reference pulse, measured in the case of far detuned qubits, thus giving access to the pulse delay τ. The delays measured in this way are presented in Fig. 3(b) showing good agreement with the spectroscopically inferred data and numeric. Due to the finite spectral width 1/σ=20 MHz of the Gaussian pulses [indicated by dashed purple lines in Fig. 2(a)] and that of the transparency window, the maximal accessible delay here is limited to τ12 ns (ng1500). In qualitative agreement with the transmission measurement data of Fig. 2(a), the amplitude of the measured pulses becomes smaller when reducing the control tone amplitude. This effect is most prominent in the sector of large delays; however, exactly here a high transmission is desired to realize an efficient quantum memory protocol.33 At maximal delay, the efficiency, i.e., the energy ratio of the transmitted and reference pulse, is about 16%. Due to small average delay-bandwidth products of 0.2, the device under investigation appears suitable for a fixed-delay quantum memory,34 rather than a general purpose on-demand storage procedure, where the pulse has to fit spatially completely into the metamaterial for high storage efficiency.33 Reducing the decoherence rate γ20 would allow for significantly larger group indices and improve the bandwidth-delay product toward values larger than unity. Experimentally, this can be achieved by employing Λ-type three-level systems such as flux- or fluxonium qubits.20 

As pointed out by Shen et al.,35 ATS-like band structures can be realized not only by dressed states but also equivalently by two detuned distinct resonances coupled to the same mode continuum. For this approach, which we call dispersion engineering, neither a third qubit level nor an additional microwave control tone is required. Here, we use the individual qubit frequency control to tune every second qubit to a frequency f1 and every other qubit to f2, creating two collective four-qubit resonances. The frequency f2=7.882 GHz is kept constant throughout the experiment, while the frequency f1 of the other four qubits is varied. The transmission S21 [Figs. 4(a) and 4(b)] resembles that of the dressed state ATS in Fig. 2; the main ATS feature, being a transparency window with a steep phase roll-off between two bandgaps, is intact. Since the large γ20 has no influence on the first two transmon levels, the brightest of the collective four qubit subradiant states are visible as pronounced peaks several MHz below f1 and f214 [compare Fig. 4(b)]. The observed line shape is in good agreement with a numerical transfer matrix calculation. The pulse delays τ and effective group indices ng, calculated from Arg (S21(ω)) with Eq. (2) with respect to the frequency f1, are shown in Fig. 4(c). Since the phase in the transparency window shows a varying slope, an average of the slope in a bandwidth of 10 MHz around the center is used to estimate τ. Analogously to the previous paragraph, we use Gaussian pulses with σ=50 ns to directly measure and validate the expected pulse delay τ. The center frequency is adjusted close to (f1+f2)/2 for each trace. Figure 4(c) compares the delay τ of pulsed and spectroscopic measurements with numerical results from a transfer-matrix model. Here, we find delays up to τ=17 ns, corresponding to group indices of ng=1850, which is comparable to the group indices found in the dressed-state ATS case for a similar splitting between the dressed states. In contrast to the ATS measurement, we achieve an efficiency of 4550%, as the transmittance of the transparency window is solely limited by the non-radiative decoherence Γnr and not influenced by γ20. This also implies potentially possible much higher delays for narrower transparency windows. Even though the dispersion-engineered approach lacks the possibility of fast and simple control via an additional microwave tone in contrast to the ATS, it can be used as a fixed delay quantum memory and showcases the ability to tailor the band structure on demand in superconducting wQED. For further studies, one can think of more complex band structures going beyond the capabilities of the ATS, e.g., engineering several spectral regions of different group indices, referred to as multi-color slow light.36 

FIG. 4.

(a) Absolute transmission |S21| for two collective four-qubit resonances at frequencies f1 and f2. The frequency is alternated between neighboring qubits along the waveguide. The black dashed line marks the center of the transparency window. The purple dashed lines indicate the bandwidth of the 50 ns Gaussian pulses. (b) Transmission for a fixed detuning f2f1=32 MHz. The sharp peaks next to the bandgaps correspond to subradiant four-qubit states. The steep phase roll-off between the resonances indicates a low group velocity. Black dashed lines are fits to a numerical transfer matrix model. (c) Experimentally extracted delays from pulsed and spectroscopic measurements in dependence on frequency f1. Pulse delays up to 17 ns, corresponding to retardation factors of 1850, are achieved.

FIG. 4.

(a) Absolute transmission |S21| for two collective four-qubit resonances at frequencies f1 and f2. The frequency is alternated between neighboring qubits along the waveguide. The black dashed line marks the center of the transparency window. The purple dashed lines indicate the bandwidth of the 50 ns Gaussian pulses. (b) Transmission for a fixed detuning f2f1=32 MHz. The sharp peaks next to the bandgaps correspond to subradiant four-qubit states. The steep phase roll-off between the resonances indicates a low group velocity. Black dashed lines are fits to a numerical transfer matrix model. (c) Experimentally extracted delays from pulsed and spectroscopic measurements in dependence on frequency f1. Pulse delays up to 17 ns, corresponding to retardation factors of 1850, are achieved.

Close modal

In this work, we demonstrated slow light in a superconducting waveguide QED system. A metamaterial consisting of eight densely spaced transmon qubits was used to engineer a band structure with a flat dispersion profile to obtain reduced group velocities of electromagnetic waves. Using the well-known dressed-state ATS, we observed group velocities reduced by a factor of down to 1500, both in spectroscopic and direct time-resolved pulsed measurements. The maximum achievable delay and efficiency is limited by the strong dephasing of the 20 transition, which is inherent to ladder-type three-level systems. Moreover, we demonstrated slow light with a tailored band structure based on distinct detuned collective resonances of the qubits. At comparable retardation, we observed three times larger efficiencies, not limited by large γ20. The demonstrated slow-light metamaterial can be employed as a fixed-delay quantum memory or a tunable pulse retarder to synchronize pulses in quantum information processors. Based on our proof-of principle experiment, further improvements which significantly ameliorate the device performance can be made. A similar metamaterial based on flux or fluxonium qubits, which feature a Λ-type level structure, is expected to reach the EIT regime and may ultimately realize a general purpose and on-demand superconducting quantum memory.

See the supplementary material for additional information on qubit characterization, the experimental setup, power calibration, and band structure calculations.

This work has received funding from the Deutsche Forschungsgemeinschaft (DFG) by Grant No. U.S. 18/15-1, the European Union's Horizon 2020 Research and Innovation Programme under Grant Agreement No. 863313 (SUPERGALAX), BMBF Grant “PtQUBE” via No. 13N15016, and by the Initiative and Networking Fund of the Helmholtz Association, within the Helmholtz Future Project “Scalable solid state quantum computing.” J.D.B. acknowledges financial support from Studienstiftung des Deutschen Volkes, A.S. was supported by Landesgraduiertenförderung-Karlsruhe, and A.N.P. gratefully acknowledges support from Russian Science Foundation via Grant No. 20-12-00194.

The authors have no conflicts to disclose.

Jan David Brehm: Conceptualization (equal); Data curation (lead); Formal analysis (lead); Investigation (lead); Methodology (equal); Software (equal); Visualization (lead); Writing – original draft (lead); Writing – review & editing (lead). Richard Gebauer: Data curation (equal); Methodology (supporting); Software (lead); Validation (supporting); Writing – original draft (supporting); Writing – review & editing (equal). Alexander Stehli: Conceptualization (supporting); Data curation (supporting); Formal analysis (supporting); Investigation (supporting); Validation (supporting); Writing – review & editing (supporting). Alexander Poddubny: Formal analysis (equal); Investigation (equal); Methodology (equal). Oliver Sander: Data curation (supporting); Project administration (supporting); Resources (supporting); Software (supporting); Supervision (supporting); Writing – review & editing (supporting). Hannes Rotzinger: Funding acquisition (supporting); Methodology (supporting); Project administration (equal); Resources (equal); Software (equal); Supervision (supporting); Writing – review & editing (supporting). Alexey V. Ustinov: Conceptualization (equal); Formal analysis (supporting); Funding acquisition (lead); Investigation (supporting); Methodology (equal); Project administration (lead); Resources (lead); Supervision (lead); Validation (supporting); Writing – original draft (supporting); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
A. I.
Lvovsky
,
B. C.
Sanders
, and
W.
Tittel
, “
Optical quantum memory
,”
Nat. Photonics
3
,
706
(
2009
).
2.
T.
Kroh
,
J.
Wolters
,
A.
Ahlrichs
,
A. W.
Schell
,
A.
Thoma
,
S.
Reitzenstein
,
J. S.
Wildmann
,
E.
Zallo
,
R.
Trotta
,
A.
Rastelli
,
O. G.
Schmidt
, and
O.
Benson
, “
Slow and fast single photons from a quantum dot interacting with the excited state hyperfine structure of the Cesium D1-line
,”
Sci. Rep.
9
(
1
),
13728
(
2019
).
3.
M.
Fleischhauer
,
A.
Imamoglu
, and
J. P.
Marangos
, “
Electromagnetically induced transparency: Optics in coherent media
,”
Rev. Mod. Phys.
77
,
633
(
2005
).
4.
L. V.
Hau
,
S. E.
Harris
,
Z.
Dutton
, and
C. H.
Behroozi
, “
Light speed reduction to 17 metres per second in an ultracold atomic gas
,”
Nature
397
,
594
(
1999
).
5.
M. M.
Kash
,
V. A.
Sautenkov
,
A. S.
Zibrov
,
L.
Hollberg
,
G. R.
Welch
,
M. D.
Lukin
,
Y.
Rostovtsev
,
E. S.
Fry
, and
M. O.
Scully
, “
Ultraslow group velocity and enhanced nonlinear optical effects in a coherently driven hot atomic gas
,”
Phys. Rev. Lett.
82
,
5229
(
1999
).
6.
B.
Gouraud
,
D.
Maxein
,
A.
Nicolas
,
O.
Morin
, and
J.
Laurat
, “
Demonstration of a memory for tightly guided light in an optical nanofiber
,”
Phys. Rev. Lett.
114
,
180503
(
2015
).
7.
Y.-F.
Hsiao
,
P.-J.
Tsai
,
H.-S.
Chen
,
S.-X.
Lin
,
C.-C.
Hung
,
C.-H.
Lee
,
Y.-H.
Chen
,
Y.-F.
Chen
,
I. A.
Yu
, and
Y.-C.
Chen
, “
Highly efficient coherent optical memory based on electromagnetically induced transparency
,”
Phys. Rev. Lett.
120
,
183602
(
2018
).
8.
O.
Astafiev
,
A. M.
Zagoskin
,
A. A.
Abdumalikov
,
Y. A.
Pashkin
,
T.
Yamamoto
,
K.
Inomata
,
Y.
Nakamura
, and
J. S.
Tsai
, “
Resonance fluorescence of a single artificial atom
,”
Science
327
,
840
(
2010
).
9.
I.-C.
Hoi
,
C. M.
Wilson
,
G.
Johansson
,
J.
Lindkvist
,
B.
Peropadre
,
T.
Palomaki
, and
P.
Delsing
, “
Microwave quantum optics with an artificial atom in one-dimensional open space
,”
New J. Phys.
15
,
025011
(
2013
).
10.
A. A.
Abdumalikov
,
O.
Astafiev
,
A. M.
Zagoskin
,
Y. A.
Pashkin
,
Y.
Nakamura
, and
J. S.
Tsai
, “
Electromagnetically induced transparency on a single artificial atom
,”
Phys. Rev. Lett.
104
,
193601
(
2010
).
11.
P. M.
Anisimov
,
J. P.
Dowling
, and
B. C.
Sanders
, “
Objectively discerning Autler-Townes splitting from electromagnetically induced transparency
,”
Phys. Rev. Lett.
107
,
163604
(
2011
).
12.
K.
Lalumière
,
B. C.
Sanders
,
A. F.
van Loo
,
A.
Fedorov
,
A.
Wallraff
, and
A.
Blais
, “
Input-output theory for waveguide QED with an ensemble of inhomogeneous atoms
,”
Phys. Rev. A
88
,
043806
(
2013
).
13.
A. F. v
Loo
,
A.
Fedorov
,
K.
Lalumière
,
B. C.
Sanders
,
A.
Blais
, and
A.
Wallraff
, “
Photon-mediated interactions between distant artificial atoms
,”
Science
342
,
1494
(
2013
).
14.
J. D.
Brehm
,
A. N.
Poddubny
,
A.
Stehli
,
T.
Wolz
,
H.
Rotzinger
, and
A. V.
Ustinov
, “
Waveguide bandgap engineering with an array of superconducting qubits
,”
npj Quantum Mater.
6
,
10
(
2021
).
15.
Y.-X.
Zhang
and
K.
Mølmer
, “
Theory of subradiant states of a one-dimensional two-level atom chain
,” arXiv:1812.09784 (
2018
).
16.
J.
Zhong
,
N. A.
Olekhno
,
Y.
Ke
,
A. V.
Poshakinskiy
,
C.
Lee
,
Y. S.
Kivshar
, and
A. N.
Poddubny
, “
Photon-mediated localization in two-level qubit arrays
,”
Phys. Rev. Lett.
124
,
093604
(
2020
).
17.
Y.
Ke
,
J.
Zhong
,
A. V.
Poshakinskiy
,
Y. S.
Kivshar
,
A. N.
Poddubny
, and
C.
Lee
, “
Radiative topological biphoton states in modulated qubit arrays
,” arXiv:2002.10074 (
2020
).
18.
A. V.
Poshakinskiy
,
J.
Zhong
,
Y.
Ke
,
N. A.
Olekhno
,
C.
Lee
,
Y. S.
Kivshar
, and
A. N.
Poddubny
, “
Quantum Hall phase emerging in an array of atoms interacting with photons
,” arXiv:2003.08257 (
2020
).
19.
Y.-L. L.
Fang
,
H.
Zheng
, and
H. U.
Baranger
, “
One-dimensional waveguide coupled to multiple qubits: photon-photon correlations
,”
EPJ Quantum Technol.
1
,
3
(
2014
).
20.
P. M.
Leung
and
B. C.
Sanders
, “
Coherent control of microwave pulse storage in superconducting circuits
,”
Phys. Rev. Lett.
109
,
253603
(
2012
).
21.
P.
Macha
,
G.
Oelsner
,
J.-M.
Reiner
,
M.
Marthaler
,
S.
André
,
G.
Schön
,
U.
Hübner
,
H.-G.
Meyer
,
E.
Il'ichev
, and
A. V.
Ustinov
, “
Implementation of a quantum metamaterial using superconducting qubits
,”
Nat. Commun.
5
,
5146
(
2014
).
22.
A. L.
Rakhmanov
,
A. M.
Zagoskin
,
S.
Savel'ev
, and
F.
Nori
, “
Quantum metamaterials: Electromagnetic waves in a Josephson qubit line
,”
Phys. Rev. B
77
,
144507
(
2008
).
23.
J. L.
Everett
,
D. B.
Higginbottom
,
G. T.
Campbell
,
P. K.
Lam
, and
B. C.
Buchler
, “
Stationary light in atomic media
,”
Adv. Quantum Technol.
2
,
1800100
(
2019
).
24.
M.
Mirhosseini
,
E.
Kim
,
V. S.
Ferreira
,
M.
Kalaee
,
A.
Sipahigil
,
A. J.
Keller
, and
O.
Painter
, “
Superconducting metamaterials for waveguide quantum electrodynamics
,”
Nat. Commun.
9
,
3706
(
2018
).
25.
J.
Koch
,
T. M.
Yu
,
J.
Gambetta
,
A. A.
Houck
,
D. I.
Schuster
,
J.
Majer
,
A.
Blais
,
M. H.
Devoret
,
S. M.
Girvin
, and
R. J.
Schoelkopf
, “
Charge-insensitive qubit design derived from the Cooper pair box
,”
Phys. Rev. A
76
,
042319
(
2007
).
26.
S. H.
Autler
and
C. H.
Townes
, “
Stark effect in rapidly varying fields
,”
Phys. Rev.
100
,
703
(
1955
).
27.
T.
Hönigl-Decrinis
,
R.
Shaikhaidarov
,
S.
de Graaf
,
V.
Antonov
, and
O.
Astafiev
, “
Two-level system as a quantum sensor for absolute calibration of power
,”
Phys. Rev. Appl.
13
,
024066
(
2020
).
28.
I.-C.
Hoi
,
C. M.
Wilson
,
G.
Johansson
,
T.
Palomaki
,
B.
Peropadre
, and
P.
Delsing
, “
Demonstration of a single-photon router in the microwave regime
,”
Phys. Rev. Lett.
107
,
073601
(
2011
).
29.
M.
Yan
,
E. G.
Rickey
, and
Y.
Zhu
, “
Electromagnetically induced transparency in cold rubidium atoms
,”
J. Opt. Soc. Am. B
18
,
1057
(
2001
).
30.
A.
Asenjo-Garcia
,
M.
Moreno-Cardoner
,
A.
Albrecht
,
H.
Kimble
, and
D.
Chang
, “
Exponential improvement in photon storage fidelities using subradiance and ‘selective radiance’ in atomic arrays
,”
Phys. Rev. X
7
,
031024
(
2017
).
31.
F.
Goldfarb
,
T.
Lauprêtre
,
J.
Ruggiero
,
F.
Bretenaker
,
J.
Ghosh
, and
R.
Ghosh
, “
Electromagnetically-induced transparency, slow light, and negative group velocities in a room temperature vapor of 4He
,”
C. R. Phys.
10
,
919
(
2009
).
32.
M. R.
Vladimirova
,
E. L.
Ivchenko
, and
A. V.
Kavokin
, “
Exciton polaritons in long-period quantum-well structures
,”
Semiconductors
32
,
90
(
1998
).
33.
I.
Novikova
,
R. L.
Walsworth
, and
Y.
Xiao
, “
Electromagnetically induced transparency-based slow and stored light in warm atoms
,”
Laser Photonics Rev.
6
,
333
(
2012
).
34.
A.
Rastogi
,
E.
Saglamyurek
,
T.
Hrushevskyi
,
S.
Hubele
, and
L. J.
LeBlanc
, “
Discerning quantum memories based on electromagnetically-induced-transparency and Autler-Townes-splitting protocols
,”
Phys. Rev. A
100
,
012314
(
2019
).
35.
J.-T.
Shen
,
M. L.
Povinelli
,
S.
Sandhu
, and
S.
Fan
, “
Stopping single photons in one-dimensional circuit quantum electrodynamics systems
,”
Phys. Rev. B
75
,
035320
(
2007
).
36.
H.
Wang
,
X.
Gu
,
Y.-x.
Liu
,
A.
Miranowicz
, and
F.
Nori
, “
Optomechanical analog of two-color electromagnetically induced transparency: Photon transmission through an optomechanical device with a two-level system
,”
Phys. Rev. A
90
,
023817
(
2014
).

Supplementary Material