Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Letter https://doi.org/10.

1038/s41586-019-1354-5

Interacting Floquet polaritons


Logan W. Clark1*, Ningyuan Jia1, Nathan Schine1, Claire Baum1, Alexandros Georgakopoulos1 & Jonathan Simon1

Ordinarily, photons do not interact with one another. However, random access architectures24 for superconducting quantum processors
atoms can be used to mediate photonic interactions1,2, raising the via first-order sideband transitions25. In ultracold atoms, many shaken
prospect of forming synthetic materials3 and quantum information lattice experiments16,17 can be modelled as frequency modulations
systems4–7 from photons. One promising approach combines highly (see Supplementary Information section B7), and frequency modula-
excited Rydberg atoms8–12 with the enhanced light–matter coupling tion has also been employed to bind diatomic molecules26.
of an optical cavity to convert photons into strongly interacting In this work, we frequency-modulate an atomic state to customize the
polaritons13–15. However, quantum materials made of optical photons coupling between atoms and photons, demonstrating that the ‘Floquet
have not yet been realized, because the experimental challenge polaritons’ that emerge behave as strongly interacting quasi-particles
of coupling a suitable atomic sample with a degenerate cavity has with spatial wavefunctions inherited from the chosen modes of the
constrained cavity polaritons to a single spatial mode that is resonant optical cavity. After loading a gas of cold Rb atoms at the waist of
with an atomic transition. Here we use Floquet engineering16,17— the cavity, we use an intensity-modulated off-resonance laser to vary the
the periodic modulation of a quantum system—to enable strongly energy of an excited atomic state sinusoidally. This modulation splits
interacting polaritons to access multiple spatial modes of an optical the state into bands separated by multiples of the modulation frequency.
cavity. First, we show that periodically modulating an excited state We choose a frequency that creates bands whose energy difference
of rubidium splits its spectral weight to generate new lines—beyond matches the separation between two transverse modes of the cavity.
those that are ordinarily characteristic of the atom—separated by This choice enables photons in both modes to couple to the atomic
multiples of the modulation frequency. Second, we use this capability ensemble and form polaritons, which we verify by measuring the
to simultaneously generate spectral lines that are resonant with two transmission spectrum. We also demonstrate that Floquet polaritons
chosen spatial modes of a non-degenerate optical cavity, enabling what have strong interactions, which in these particular modes hinders
we name ‘Floquet polaritons’ to exist in both modes. Because both multiple polaritons from entering the cavity simultaneously. We con-
spectral lines correspond to the same Floquet-engineered atomic state, clude by discussing possible uses of Floquet polaritons in many-body
adding a single-frequency field is sufficient to couple both modes to physics and of customized atomic spectra in quantum science more
a Rydberg excitation. We demonstrate that the resulting polaritons broadly.
interact strongly in both cavity modes simultaneously. The production Our experiments begin by controllably loading a sample of 300–1,800
of Floquet polaritons provides a promising new route to the realization cold 87Rb atoms at the waist of a four-mirror optical cavity (Fig. 1a).
of ordered states of strongly correlated photons, including crystals and The cavity has modes that are near-detuned to the atomic transition
topological fluids, as well as quantum information technologies such between the 5S1/2 ground state and the 5P3/2 excited state at 780 nm.
as multimode photon-by-photon switching. To modulate the energy of the excited state, we expose the atoms to a
Photons coupled to an atomic ensemble form polaritons, quasipar- multichromatic optical field tuned near the 5P3/2 → 5D5/2 transition at
ticles that inherit key properties from both their matter and light com- 776 nm. This multichromatic field simultaneously cancels the constant
ponents18. The atomic component of the polaritons provides them with component of the Stark shift and produces a time-varying component
the interactions essential for forming non-trivial ordered phases, and of the Stark shift that oscillates with tunable frequency f (on a scale
their photonic component enables them to propagate through space. of megahertz to gigahertz) and amplitude η (Fig. 1b). We tune the
Adding an optical cavity reshapes the space available for polaritons by modulation amplitude by adjusting the total intensity of the 776-nm
defining a discrete set of modes that can couple to the atomic ensemble. beam; amplitudes are reported in units of the arbitrarily defined
To enable the study of quantum many-body physics with polaritons3,19, η0 = h × 17(1) MHz, where h is the Planck constant, which corresponds
progress has been made in designing cavity structures to create desir- to a beam intensity of approximately 5 W mm−2, determined by the
able spaces for polariton propagation, including the realization of polarizability at a detuning of 1 GHz from the 5P3/2 → 5D5/2 transition.
synthetic Landau levels for photons20–22. However, the challenges of For more details on the experimental setup, see Methods.
engineering these systems have so far resulted in studies of strongly Periodic modulation splits the excited state into bands at energies
interacting photons in only a single transverse mode1,14. Ej = E0 + jhf, relative to its unmodulated energy E0, for integers j
One approach for creating multi-mode polaritons is Floquet (Fig. 1c). For sinusoidal modulation, the collective atom–photon cou-
engineering—the periodic modulation of parameters to generate pling strengths gj at each band should take the form
desirable new properties in quantum mechanical systems. Floquet
engineering has proved to be a powerful tool for studying quantum η 
g j (η , f ) = gJ j   (1)
many-body physics with ultracold atoms, where it has enabled tests of  hf 
quantum phase transitions, the creation of new interaction processes
and the development of synthetic gauge fields for studies of topology17 where g is the unmodulated coupling and Jj is the jth Bessel function
(see Methods and Supplementary Information section B7). of the first kind (see Supplementary Information section B3). We
Frequency modulation, the periodic variation of the energy of note that modulation does not create any new states, but instead redis-
a quantum state, is a widespread form of Floquet engineering16. tributes the spectral weight of a single state between multiple energies;
Periodically modulating a state splits it into multiple bands at different this is reflected in the constraint that the total coupling strength
energies, analogous to frequency modulation in signal processing. 2
g = ∑ j g j (η , f ) is unchanged (see Methods).
Frequency modulation has enabled faster manipulation23 and efficient

1
James Franck Institute and Department of Physics, University of Chicago, Chicago, IL, USA. *e-mail: lwclark@uchicago.edu

N A t U r e | www.nature.com/nature
RESEARCH Letter

a Modulation beam b sufficient modulation, additional avoided crossings become clearly vis-
K ible for the first-order bands at j = ±1 and the second-order bands at

Energy, Ep
0 j = ±2. The frequencies of the observed features deviate slightly from
–K
Ej = E0 + jhf owing to shifts from off-resonance couplings with the
1/f other bands (see Supplementary Information section B4).
Time We extract the coupling strengths gj(η) for each modulation ampli-
tude from the widths of the avoided crossings (Fig. 1e). Although the
c j=2 gj
observed coupling strengths are qualitatively similar to the prediction
j=1 of equation (1), the Bessel functions are distorted by inhomogeneity
f Gc
5P3/2 j=0 of the modulation beam and by slight asymmetry due to higher-order
j = –1 Stark shifts. A theoretical treatment that accounts for these two factors
j = –2 agrees well with the behaviour observed in Fig. 1e (see Supplementary
Modulation amplitude, K Information section A1).
5S1/2
When the atomic transition is split into bands, it can couple res-
Atomic states Cavity onantly to multiple transverse electromagnetic (TEM) modes of the
d K=0 2K0 4K0 cavity simultaneously (Fig. 2a). In our cavity, the fundamental TEM00
mode is conveniently close to the TEM40 mode, which is only 52 MHz
2 away. By modulating the atoms at a frequency near this mode separa-
100
tion, we couple the TEM00 mode with the atoms through the j = 0 band

Probe detuning (MHz)


1
50 and simultaneously couple the TEM40 mode with the atoms through
the j = 1 band. The choice of two nearby modes minimizes the neces-
j
nd
ba

0 sary modulation frequency. We note that each cavity mode couples to


Transmission (%)

50 0
2

a unique collective atomic excitation, in which all of the atoms share


3/
5P

–1 one excitation with relative amplitudes proportional to the electric field


-50
of the mode at each atom’s location. Therefore, the coupling g does not
–2 induce any mixing between different cavity modes (see Supplementary
0 –100 Information section B2).
–100 –50 0 50 100 The next step in providing photons with the strong interactions
Cavity detuning, Gc (MHz) of highly excited Rydberg atoms is to add a 480-nm field to couple
e the 5P3/2 state to a Rydberg state. Because every band is part of the
20 j=2 same 5P3/2 state, this single-frequency field is sufficient to provide the
Atom–cavity coupling, gj (MHz)

j=1
Rydberg coupling. For instance, a cavity mode can couple to a 5P3/2
j=0
15 j = –1 excitation through the j = 1 band, and that 5P3/2 excitation can then
j = –2 still be coupled to a Rydberg excitation via a different band, such as
j = 0. In this work, we use the single-frequency field tuned to create
10 Rydberg coupling via the j = 0 band regardless of the bands used for
atom–cavity coupling.
5
The eigenstates of the atom–cavity system are superpositions of col-
lective modulated atomic excitations with cavity photons, which we
name Floquet polaritons (Fig. 2b). Each cavity mode TEMm0 yields
0 three types of polaritons27: two bright polariton states Bm± and one
0 1 2 3 4 5 6
Modulation amplitude, K (K0) dark polariton state Dm. The bright polaritons are primarily com-
posed of a photon in the corresponding cavity mode and a collective
Fig. 1 | Redistributing the spectral density of atoms coupled to a cavity. 5P3/2 excitation; their 5P3/2 components, with a high decay rate of
a, Photons in a four-mirror optical cavity (red) are strongly coupled to Γ = 2π × 6 MHz, make them short-lived. We are primarily interested
excitations of a cloud of ultracold atoms (dots). An off-resonance laser
in the dark polaritons, which are superpositions of a cavity photon with
(green) is used to customize the atomic spectrum. b, The laser sinusoidally
modulates the energy Ep of the 5P3/2 atomic state with amplitude η a collective Rydberg excitation. Dark polaritons are useful for studying
and frequency f. c, As the modulation amplitude η increases, the 5P3/2 quantum many-body physics because they are long-lived and strongly
state splits into bands at energies Ej = jhf for integers j. Each band has a interacting, as long as the Rydberg blockade radius—the distance over
collective coupling strength of g j to the cavity. δc is the cavity detuning. which one Rydberg atom prevents a second one from being excited—is
d, Transmission spectra measured while scanning the cavity length exhibit about the same as the mode size of the cavity14.
avoided crossings, indicative of atom–cavity coupling, for up to five To detect dark polaritons we measure the transmission spectrum of
different bands, depending on the modulation amplitude. The frequencies the atom–cavity system for a fixed cavity length (Fig. 2c). We choose
of the cavity mode and probe laser are shown as detunings relative to the a length for which the TEM00 mode is resonant with the j = 0 band
unmodulated 5S1/2 → 5P3/2 atomic transition. Transmission is reported and the TEM40 mode is resonant with the j = 1 band. Without mod-
relative to an empty cavity. e, The coupling strengths gj extracted from
the transmission spectra for bands j = 0, ±1, ±2. Error bars are 1σ,
ulation (η = 0), we observe the predicted polariton features in the
smaller than the symbol heights. Curves show a global fit accounting for TEM00 mode, including two bright polaritons widely split as a result
inhomogeneity and slightly asymmetric modulation (see Supplementary of strong light–matter coupling, as well as a dark polariton at a fre-
Information section A1). quency between them. The bright polaritons are broad owing to the
rapid decay of the 5P3/2 component, but the dark polariton’s low decay
rate Γd = 2π × 0.3 MHz makes its transmission feature much narrower.
To test for the redistribution of the excited state into bands, we However, without modulation, there is no coupling through the j = 1
measure the transmission spectrum of a single mode of the cavity band (g1 = 0) and the atoms do not resonantly couple to the TEM40
while scanning the cavity length (Fig. 1d). Whenever the energy of the mode; we therefore observe a transmission feature in that mode that is
mode approaches the energy of an atomic band, we observe an avoided equivalent to an empty cavity.
crossing in the spectrum. Without modulation, only a single avoided Increasing the modulation amplitude couples the atoms with the
crossing is observed, corresponding to the original band j = 0. With TEM40 mode by increasing the sideband strength, g1, at the expense

N A t U r e | www.nature.com/nature
Letter RESEARCH

a b c
Rydberg TEM00 TEM40
B4+ 50

Transmission (%)
D4 40
Ω D4
Γ g1 B4– 30
g0 TEM40
20
5P3/2 TEM00 D0
B0+ 10

D0 0 B4+
0
B0– Mo B0– B0+ B4–
am 2
5S1/2 pli dulat 60 80
tud ion 4 20 40
Atomic Cavity Polaritons e, –20 0
K( Probe frequency (MHz)
states modes K)
0

Fig. 2 | Forming Floquet polaritons in a customized space. a, To form depicted in a; light and dark blue, TEM00 and TEM40, respectively; green,
Rydberg polaritons we add a field at 480 nm to couple the 5P3/2 state to a 5P3/2 state; red, Rydberg state. c, With the cavity length fixed, we probe
Rydberg state with coupling strength Ω. When the modulation frequency the transmission spectrum of the combined atom–cavity system. To avoid
f matches the energy spacing between the TEM00 and TEM40 cavity modes, blockade effects, we use the weakly interacting 39S1/2 Rydberg level, with
photons in both modes are resonantly coupled to excitations of the atoms a small blockade radius of 4 μm compared to the 12-μm cavity waist. The
with strengths determined by the resonant bands. b, The eigenstates of probe laser is in an equal superposition of the TEM00 and TEM40 modes,
this atom–cavity system are sets of three polaritons in each spatial mode and we simultaneously monitor the photons that are output from both
TEMm0, consisting of two bright polaritons Bm± composed primarily of modes on a single detector. Without modulation (η = 0), we observe
a cavity photon and a collective 5P3/2 excitation, and one dark polariton the predicted polariton features in TEM00, but TEM40 exhibits only an
Dm composed of a cavity photon and a collective Rydberg excitation. ordinary cavity transmission line. With sufficient modulation, we detect
The bright polaritons are broad owing to the rapid decay of their 5P3/2 all six of the polariton features predicted in b. Additional frequency shifts
component at a rate of Γ = 2π × 6 MHz. Colours represent the relative and asymmetries in the polariton spectra result from couplings to off-
contributions to each eigenstate from the corresponding bare states resonance bands (see Supplementary Information section B4).

of the original feature strength, g0. When this mode couples to the photon is present, compared to the non-interacting case. A substan-
atoms, the original TEM40 feature is divided into the three polaritons tial suppression indicates that the presence of just one dark polar-
B4± and D4, and the separation between the B0± bright polaritons is iton in any mode of the cavity impedes the entry of a second dark
narrowed owing to the weaker light–matter coupling in that mode. polariton. More specifically, the magnitude of the correlation minima
Although the primary features come from the 5P3/2 bands that are res- G(0)/G(∞) = γ2/(U2 + γ2) reveals the strength of the interactions U
onantly coupled to each cavity mode, the other bands also couple to relative to the polariton lifetime γ; suppression of the zero-time cor-
the cavity modes off-resonantly, inducing small shifts of the polariton relation by half indicates an interaction strength equal to the polariton
energies. These off-resonance couplings cause the observed asymmetry lifetime. The polariton lifetime also determines the correlation time
between the Bm+ and Bm− features (see Supplementary Information over which multiple photon events are suppressed (see Supplementary
section B4). Moreover, although their spectroscopic features remain Information section A2). By contrast, perfect coherent light without
separated, a treatment using Floquet theory reveals that the two dark interactions, such as a laser beam, exhibits a flat correlation function
polariton modes are governed by an effective Hamiltonian in which G(τ) = 1. Classical fluctuations, such as intensity instability, cause the
the modes are nearly degenerate, with their relative energies controlled correlation function to rise above 1. In our system, the background
by the modulation frequency (see Supplementary Information section correlation values result primarily from trial-to-trial fluctuations in
B5). This effect cannot be achieved by the frequency modulation of the the number of atoms.
Rydberg coupling field, which would enable additional modes to host Future applications of this system to quantum information, such as
dark polaritons satisfying two-photon resonance, but which would also in multimode photon-by-photon switching, would benefit from the use
have an unacceptably large mismatch between the cavity photons and of Floquet polaritons with lifetimes and interaction strengths compa-
the 5S1/2 → 5P3/2 transition. A mismatch effectively reduces the optical rable to polaritons optimized for blockade in a single mode (Fig. 3e).
density of the atomic sample, making the dark polaritons very fragile, Simulations using non-Hermitian perturbation theory indicate that
for example, having a reduced lifetime (see Supplementary Information the observed difference in performance between these two cases
section A3). results from surmountable technical limitations (see Supplementary
We next test for interactions between the Floquet dark polari- Information section A2). In particular, modulation splits the atom–
tons (Fig. 3a). Here we couple to the 100S1/2 Rydberg state, which is cavity coupling g and the Rydberg coupling Ω between multiple bands,
expected to make the polaritons interact strongly because its blockade weakening the resonant light–matter coupling for any particular mode.
radius of 25 μm is larger than the cavity mode waist of 12 μm Upgrading the system to achieve the atomic densities typical of free-
(see Methods)14. With sufficient modulation to support dark polar- space experiments8 would increase g by an order of magnitude, and the
itons in both modes, we probe the cavity at the D4 feature in the use of a buildup cavity for the Rydberg coupling beam would similarly
forward direction and simultaneously probe the D 0 feature in enhance Ω. These modifications would enable the multimode perfor-
the backward direction. We then monitor the photons leaking out of mance (with frequency modulation) to reach or surpass the perfor-
the cavity in each mode, and test for photonic interactions via the cor- mance shown in Fig. 3e.
relation function Gmn(τ) between photons from TEMm0 and photons We have realized and characterized interacting Floquet polaritons.
from TEMn0, separated by a time τ (see Supplementary Information These polaritons exist in a completely customizable space in which
section A2). In this system, correlations arise exclusively from strong their modes and energetic structure (see Supplementary Information
interactions between individual polaritons, as a result of the Rydberg section B5) are controlled by the frequency modulation of an atomic
components of the polaritons14. gas in an optical cavity. Floquet polaritons can be used to study many-
Photon antibunching—the suppression of photon coincidence body physics with features that have not yet been realized in any other
events relative to an uncorrelated field—reveals strong interactions platform, including physics in multiply connected ‘polariton worm-
between the Floquet polaritons (Fig. 3b–d). Antibunching appears as holes’ and the non-Abelian topological phases of multilayer fractional
a minimum in the correlation functions Gmn at zero time delay (τ = 0), quantum Hall systems (see Supplementary Information section B6).
where the correlation can be interpreted as the likelihood that a second The customizable structures are also rapidly tunable via adjustments to

N A t U r e | www.nature.com/nature
RESEARCH Letter

a b d 1.7
1.7

Cross-correlation G40
Autocorrelation G00
Probe D4 Modulation beam Probe D0
forward backward 1.3 1.3

0.9 0.9

0.5 0.5
c e
1.7 1.5

Autocorrelation G44

Unmodulated G00
1.3 1.0

TEM00 single-photon TEM40 single-photon 0.9 0.5


counters counters

0.5 0
–6 –3 0 3 6 –6 –3 0 3 6
Time delay, W (μs) Time delay, W (μs)
Fig. 3 | Strong interactions between Floquet dark polaritons. a, We use these modes blockade further transmission in the same mode, as indicated
the highly excited 100S1/2 Rydberg state to make dark polaritons interact by the correlation minima at τ = 0. Strong interactions between polaritons
strongly with each other, so that the presence of a single polariton in the in different modes lead to cross-blockade, shown by the cross-correlation
cavity hinders the entry of others. To test for this blockade effect, we probe function (G40; d). Under conditions optimized for single-mode blockade
the atom–cavity system simultaneously with one laser propagating in the without modulation (η = 0), the correlation minimum for G00 is close to
forward direction tuned to the dark polariton D4 resonance and a second zero (e). Error bars indicate 1σ. The solid curves show fits of the data to a
laser propagating in the backward direction tuned to the D0 resonance. generalization of the optical Bloch equations that allows one polariton to
Single-photon counters monitor the cavity emission in each mode. occupy either of the two modes (see Supplementary Information section
b–e, The autocorrelation functions for the TEM00 (G00; b) and TEM40 A2). Insets show the modulated (b–d) and unmodulated (e) cavity with
(G44; c) modes with frequency modulation η = 2.7η0 show that photons in the corresponding TEM modes.

the modulation, raising the prospect of inducing and studying polari- 13. Guerlin, C., Brion, E., Esslinger, T. & Mølmer, K. Cavity quantum electrodynamics
with a Rydberg-blocked atomic ensemble. Phys. Rev. A 82, 053832 (2010).
ton dynamics28–30. Thus, Floquet polaritons are suitable for studying 14. Jia, N. et al. A strongly interacting polaritonic quantum dot. Nat. Phys. 14,
strongly correlated materials made of photons, including crystals and 550–554 (2018).
Laughlin states29–32 as well as for quantum information science12,24. 15. Georgakopoulos, A., Sommer, A. & Simon, J. Theory of interacting cavity
Rydberg polaritons. Quantum Sci. Technol. 4, 014005 (2018).
More broadly, the Floquet engineering scheme presented here has a 16. Silveri, M. P., Tuorila, J. A., Thuneberg, E. V. & Paraoanu, G. S. Quantum systems
variety of other prospective applications, for example, in matching under frequency modulation. Rep. Prog. Phys. 80, 056002 (2017).
atomic spectra with the spectra of other physical systems for quan- 17. Eckardt, A. Atomic quantum gases in periodically driven optical lattices.
Rev. Mod. Phys. 89, 011004 (2017).
tum information applications7 or in tuning spectra to enable new 18. Fleischhauer, M., Imamoglu, A. & Marangos, J. Electromagnetically induced
laser-cooling schemes33. transparency: optics in coherent media. Rev. Mod. Phys. 77, 633–673 (2005).
19. Douglas, J. S. et al. Quantum many-body models with cold atoms coupled to
Online content photonic crystals. Nat. Photon. 9, 326–331 (2015).
20. Schine, N., Ryou, A., Gromov, A., Sommer, A. & Simon, J. Synthetic Landau levels
Any methods, additional references, Nature Research reporting summaries, source for photons. Nature 534, 671–675 (2016).
data, statements of data availability and associated accession codes are available at 21. Lim, H.-T., Togan, E., Kroner, M., Miguel-Sanchez, J. & Imamoğlu, A. Electrically
https://doi.org/10.1038/s41586-019-1354-5. tunable artificial gauge potential for polaritons. Nat. Commun. 8, 14540
(2017).
Received: 27 June 2018; Accepted: 30 April 2019; 22. Schine, N., Chalupnik, M., Can, T., Gromov, A. & Simon, J. Electromagnetic and
gravitational responses of photonic Landau levels. Nature 565, 173–179
Published online xx xx xxxx. (2019).
23. Strand, J. D. et al. First-order sideband transitions with flux-driven asymmetric
1. Birnbaum, K. M. et al. Photon blockade in an optical cavity with one trapped transmon qubits. Phys. Rev. B 87, 220505 (2013).
atom. Nature 436, 87–90 (2005). 24. Naik, R. et al. Random access quantum information processors using
2. Chang, D. E., Vuletić, V. & Lukin, M. D. Quantum nonlinear optics—photon by multimode circuit quantum electrodynamics. Nat. Commun. 8, 1904
photon. Nat. Photon. 8, 685–694 (2014). (2017).
3. Carusotto, I. & Ciuti, C. Quantum fluids of light. Rev. Mod. Phys. 85, 299–366 25. Beaudoin, F., da Silva, M. P., Dutton, Z. & Blais, A. First-order sidebands in circuit
(2013). QED using qubit frequency modulation. Phys. Rev. A 86, 022305 (2012).
4. Raimond, J. M., Brune, M. & Haroche, S. Manipulating quantum entanglement 26. Beaufils, Q. et al. Radio-frequency association of molecules: an assisted
with atoms and photons in a cavity. Rev. Mod. Phys. 73, 565–582 (2001). Feshbach resonance. Eur. Phys. J. D 56, 99–104 (2010).
5. Duan, L.-M., Lukin, M., Cirac, J. I. & Zoller, P. Long-distance quantum 27. Ningyuan, J. et al. Observation and characterization of cavity Rydberg
communication with atomic ensembles and linear optics. Nature 414, 413–418 polaritons. Phys. Rev. A 93, 041802 (2016).
(2001). 28. Zeuthen, E., Gullans, M. J., Maghrebi, M. F. & Gorshkov, A. V. Correlated photon
6. Kimble, H. J. The quantum internet. Nature 453, 1023–1030 (2008). dynamics in dissipative Rydberg media. Phys. Rev. Lett. 119, 043602 (2017).
7. Saffman, M., Walker, T. G. & Mølmer, K. Quantum information with Rydberg 29. Ivanov, P. A., Letscher, F., Simon, J. & Fleischhauer, M. Adiabatic flux insertion
atoms. Rev. Mod. Phys. 82, 2313–2363 (2010). and growing of Laughlin states of cavity Rydberg polaritons. Phys. Rev. A 98,
8. Peyronel, T. et al. Quantum nonlinear optics with single photons enabled by 013847 (2018).
strongly interacting atoms. Nature 488, 57–60 (2012). 30. Dutta, S. & Mueller, E. Coherent generation of photonic fractional quantum Hall
9. Dudin, Y., Li, L., Bariani, F. & Kuzmich, A. Observation of coherent many-body states in a cavity and the search for anyonic quasiparticles. Phys. Rev. A 97,
Rabi oscillations. Nat. Phys. 8, 790–794 (2012). 033825 (2018).
10. Tiarks, D., Baur, S., Schneider, K., Dürr, S. & Rempe, G. Single-photon transistor 31. Sommer, A., Büchler, H. P. & Simon, J. Quantum crystals and Laughlin droplets
using a Förster resonance. Phys. Rev. Lett. 113, 053602 (2014). of cavity Rydberg polaritons. Preprint at https://arxiv.org/abs/1506.00341
11. Gorniaczyk, H., Tresp, C., Schmidt, J., Fedder, H. & Hofferberth, S. Single-photon (2015).
transistor mediated by interstate Rydberg interactions. Phys. Rev. Lett. 113, 32. Ozawa, T. et al. Topological photonics. Rev. Mod. Phys. 91, 015006 (2019).
053601 (2014). 33. Norcia, M. A., Cline, J. R. K., Bartolotta, J. P., Holland, M. J. & Thompson, J. K.
12. Thompson, J. D. et al. Symmetry-protected collisions between strongly Narrow-line laser cooling by adiabatic transfer. New J. Phys. 20, 023021
interacting photons. Nature 542, 206–209 (2017). (2018).

N A t U r e | www.nature.com/nature
Letter RESEARCH

Acknowledgements We thank L. Feng for comments on the manuscript. This Competing interests The authors declare no competing interests.
work was supported by DOE grant DE-SC0010267 for apparatus construction,
AFOSR grant FA9550-18-1-0317 for modelling and MURI grant FA9550-16- Additional information
1-0323 for data collection and analysis. N.S. acknowledges support from a Extended data is available for this paper at https://doi.org/10.1038/s41586-
University of Chicago Grainger graduate fellowship and C.B. acknowledges 019-1354-5.
support from the NSF GRFP. Supplementary information is available for this paper at https://doi.org/
10.1038/s41586-019-1354-5.
Reviewer information Nature thanks Oliver Morsch, Michael Sentef and the other Reprints and permissions information is available at http://www.nature.com/
anonymous reviewer(s) for their contribution to the peer review of this work. reprints.
Correspondence and requests for materials should be addressed to L.W.C.
Author contributions The experiment was designed and built by all authors. Publisher’s note: Springer Nature remains neutral with regard to jurisdictional
L.W.C., N.J. and N.S. collected the data. L.W.C. and N.J. analysed the data. L.W.C. claims in published maps and institutional affiliations.
and J.S. developed the theory. L.W.C. prepared the manuscript, and all authors
contributed to the final manuscript. © The Author(s), under exclusive licence to Springer Nature Limited 2019

N A t U r e | www.nature.com/nature
RESEARCH Letter

Methods than the original beam. The sidebands that eventually produce the greatest shifts
Experimental setup. The experimental apparatus used for this work is described of the 5P3/2 state are the blue-detuned bands at δ = ∆ and δ = ∆ + f, which have
in the supplementary information of ref. 14. We controllably transport a sample of similar detuning to the original beam and are first order in the power of the radio
300–1,800 87Rb atoms into the region spanned by the TEM40 mode of an optical frequency. These two frequency components are equivalent to a single component
cavity, corresponding to total atom–cavity coupling strengths of g = 8–19 MHz at at δ = ∆ + f/2 that is intensity-modulated at f. Therefore, our setup can be under-
the |5S1/2, F = 2〉 → |5P3/2, F′ = 3〉 atomic transition for each cavity mode, where F stood in essence as producing a red-detuned component with constant intensity
and F′ label the ground and excited hyperfine manifolds, respectively. A blue laser and a blue-detuned component with modulated intensity. Finally, to enable large
beam near 480 nm couples the 5P3/2 → nS1/2 transition for the Rydberg principal shifts of the 5P3/2 state, the entire modulated beam is sent through a tapered ampli-
quantum number n. A diagram of the atomic levels and transitions relevant to this fier to achieve a maximum power of about 1 W before illuminating the atomic
work is shown in Extended Data Fig. 1a. sample. We experimentally adjust the exact detunings and power of the radio fre-
As in our previous work, the atomic sample is ‘sliced’ to a root-mean-square quency in order to achieve the largest possible 5P3/2 modulation for a given total
length of approximately 10 μm, which is sufficiently small for one 100S1/2 Rydberg laser power, as well as ensuring that the average Stark shift of the 5P3/2 state is zero.
excitation to blockade the formation of further Rydberg excitations. In particular, We have designed our setup to avoid a few detrimental side effects of frequency
because we are only weakly driving the cavity, the Rydberg blockade radius7 is modulation. First, we have attempted to make the average Stark shift of the 5P3/2
state as small as possible, such that the j = 0 band always remains at the same
1 /6
C  energy as the unmodulated 5P3/2 state. This choice is important because it ensures
RB =  6  (2) that the 5P3/2 band frequencies do not vary spatially owing to inhomogeneity in the
 γ 
R modulation laser beam intensity profile, nor do they fluctuate over time owing to
total intensity instability. Moreover, cancelling the average Stark shift also ensures
where the effective Rydberg–Rydberg interaction potential V(R) = C6/R6 becomes that we do not have to tune the cavity length to match a new j = 0 band frequency
equal to the Rydberg decay rate γR ≈ h × 200 kHz. For the 39S Rydberg state, the each time we adjust the modulation amplitude. We note that, in this setup, the
coefficient C639 = h × 730 MHz μm6 corresponds to a radius of just RB39 = 4 μm , largest remaining source of temporal drift in the average Stark shift of the 5P3/2
whereas the much more strongly interacting 100S Rydberg state has coefficient state seems to come from the tapered amplifier, in which the relative amplification
C6100 = h × 56 GHz μm6 for a blockade radius of RB100 = 25 μm. This difference in of the various electro-optic modulator sidebands exhibits very small fluctuations
blockade radius corresponds to a blockade volume that is 240 times smaller for the owing to temperature instability. We suspect that this comes from a weak etalon
weakly interacting state. effect in the amplifier chip; we observe that the 5P3/2 energy oscillates as we steadily
The cavity used in this work is the same as that described previously14. The increase the amplifier temperature. We believe that we have somewhat mitigated
cavity is non-degenerate, with a free spectral range of 2,204.6 MHz and a transverse this effect by choosing an amplifier temperature that is at an extremum of the
mode spacing along the vertical axis of 564.05 MHz. As a result, there is a TEM40 oscillation, making the system quadratically insensitive to temperature fluctuation
mode (4 × 564.05 MHz) – 2,204.6 MHz = 51.6 MHz higher than the fundamental around the extremum.
TEM00 mode. Each cavity transverse mode contains a pair of orthogonal polariza- A second detrimental side effect comes from the off-resonance shift of the
tion modes that are nearly linear; however, throughout this work we use only the 5S1/2 state. The multichromatic field is detuned by only 4 nm from the strong
modes that are approximately horizontally polarized. 5S1/2 → 5P3/2 transition at 780 nm. In principle, this shift could be cancelled as
The enormous static electric polarizability of the 100S Rydberg state, well, for example, by using a copropagating beam at approximately 784 nm, but
α ∝ 6 GHz cm2 V−2, makes it very easy for small electric field drifts or inhomoge- cancellation was not practical in this case. We instead minimized the inhomoge-
neities to decohere the collective Rydberg states. A detailed theoretical treatment of neity of this shift—which would otherwise cause broadening of the dark polariton
this effect in the context of our system can be found in a previous study15. We have lines—by ensuring that the beam was large (approximately round with a waist
taken care to design our cavity and its mounting structure to keep all surfaces as far of 70 μm) compared to the horizontal cavity mode waist of 12 μm. Moreover,
away from the atoms as possible (the closest surface is 12 mm away), so that any the 776-nm beam propagates along the vertical axis (the long axis of the TEM40
patch potentials or dipoles that build up from Rb adsorbates have minimal effect mode), ensuring that along that axis its intensity is approximately homogeneous
on the atoms. Moreover, we use active electric field control with eight intracavity across the sample without needing to increase the waist size. With sufficiently small
electrodes in order to zero the electric field at the location of the atoms, maximizing inhomogeneity, we are able to slightly adjust the cavity length to account for the net
the Rydberg coherence time. Our effective Rydberg decay rate, γR ≈ h × 200 kHz shift of the 5S1/2 state without any deterioration of the performance of our system.
at 100S, is not at present limited by the background electric field. For more detail In the future, we intend to upgrade our apparatus to use a modulation field near
on our experimental efforts to control the electric field, see supplement D of our the 5P3/2 → 4D5/2 transition at 1,529 nm instead of the existing 776-nm field. This
previous work14. substitution would have three key advantages. First, the 5P3/2 → 4D5/2 transition
We have reshaped the transverse beam profile to ensure that the blue beam, is stronger than that of 5P3/2 → 5D5/2; switching transitions makes the energy
which propagates approximately along the cavity mode axis, covers the atomic modulation approximately 40 times larger for similar beam intensity. Second, the
sample nearly homogeneously while still achieving sufficiently strong Rydberg new detuning from the 5S1/2 → 5P3/2 transition would be nearly 100 times greater,
coupling. The blue beam then has an elliptical intensity profile with a vertical waist dramatically reducing the off-resonance shift of the 5S1/2 state. Finally, fibre elec-
of 64 μm and a horizontal waist of 20 μm; both which exceed the corresponding tro-optic modulators operating at 1,529 nm can operate with much higher total
vertical (42 μm) and horizontal (12 μm) widths of the TEM40 mode. powers, even exceeding 1 W, enabling us to modulate the beam after it is amplified,
Frequency modulation setup. Floquet engineering with ultracold atoms has and thus prevent the amplifier from causing instability in the intensity ratios of the
enabled tests of quantum phase transitions34–36, the creation of new interaction various frequency components in the field.
processes37–39 and the development of synthetic gauge fields40–45 for studies of Throughout the text, we report modulation amplitudes in units of the arbi-
topology46–49. The frequency modulation setup that we use in this work is directly trarily defined η0, which is defined as the peak energy Ep of the 5P3/2 state in each
analogous to the photon-assisted tunnelling used in many of these experiments modulation cycle for a total 776-nm beam intensity of approximately 5 W mm−2.
with ultracold atoms17 (see Supplementary Information section B7). From the fit result of Fig. 1e, detailed in Supplementary Information section A1,
The goal of our frequency modulation setup is to vary the energy Ep of the we estimate that η0 = h × 17(1) MHz.
5P3/2 state sinusoidally. We achieve this goal using a multichromatic field near Finally, we note that the three sets of experiments reported in this work each
the 5P3/2 → 5D5/2 transition at 776 nm (Extended Data Fig. 1b). The field con- use a slightly different modulation frequency. In Fig. 1 we use a frequency of
tains red-detuned and blue-detuned components (lower and higher frequency f = 53 MHz, for Fig. 2 f = 58 MHz, and for Fig. 3 f = 54 MHz. Each of these fre-
than resonance, respectively), each approximately 1 GHz away from resonance quencies is close to the bare cavity mode spacing of 52 MHz between the TEM00
on opposite sides of the transition, such that there is no average Stark shift of the and TEM40 modes, but differs slightly to compensate for the shifts which come
5P3/2 state. Then, we intensity-modulate the blue-detuned component to cause the from coupling to off-resonance bands (see Supplementary Information section B4).
energy of the 5P3/2 state to oscillate around zero with controllable frequency f and
amplitude η. We tune the modulation amplitude by adjusting the total intensity Data availability
of the 776-nm laser. The experimental data presented in this manuscript is available from the corre-
To generate the multichromatic field, we begin with a single frequency source sponding author upon request.
with a total power of 20 mW that is red-detuned from the transition and locked at
a detuning δ = −∆ (where ∆ = 1 GHz), relative to the resonance frequency. We 34. Ma, R. et al. Photon-assisted tunneling in a biased strongly correlated Bose gas.
then generate the blue-detuned component using a fibre electro-optic modulator, Phys. Rev. Lett. 107, 095301 (2011).
driven with two radio-frequency tones at 2∆ and 2∆ + f. This modulation gener- 35. Parker, C. V., Ha, L.-C. & Chin, C. Direct observation of effective ferromagnetic
ates a variety of sidebands, most of which are further detuned from the resonance domains of cold atoms in a shaken optical lattice. Nat. Phys. 9, 769–774 (2013).
Letter RESEARCH

36. Clark, L. W., Feng, L. & Chin, C. Universal space-time scaling symmetry in the 43. Miyake, H., Siviloglou, G. A., Kennedy, C. J., Burton, W. C. & Ketterle, W. Realizing
dynamics of bosons across a quantum phase transition. Science 354, 606–610 the Harper Hamiltonian with laser-assisted tunneling in optical lattices. Phys.
(2016). Rev. Lett. 111, 185302 (2013).
37. Daley, A. J. & Simon, J. Effective three-body interactions via photon-assisted 44. Tai, M. E. et al. Microscopy of the interacting Harper–Hofstadter model in the
tunneling in an optical lattice. Phys. Rev. A 89, 053619 (2014). two-body limit. Nature 546, 519–523 (2017).
38. Meinert, F., Mark, M. J., Lauber, K., Daley, A. J. & Nägerl, H.-C. Floquet 45. Clark, L. W. et al. Observation of density-dependent gauge fields in a
engineering of correlated tunneling in the Bose–Hubbard model with ultracold Bose–Einstein condensate based on micromotion control in a shaken
atoms. Phys. Rev. Lett. 116, 205301 (2016). two-dimensional lattice. Phys. Rev. Lett. 121, 030402 (2018).
39. Clark, L. W., Gaj, A., Feng, L. & Chin, C. Collective emission of matter-wave jets 46. Jotzu, G. et al. Experimental realization of the topological Haldane model with
from driven Bose–Einstein condensates. Nature 551, 356–359 (2017). ultracold fermions. Nature 515, 237–240 (2014).
40. Lignier, H. et al. Dynamical control of matter-wave tunneling in periodic 47. Aidelsburger, M. et al. Measuring the Chern number of Hofstadter bands with
potentials. Phys. Rev. Lett. 99, 220403 (2007). ultracold bosonic atoms. Nat. Phys. 11, 162–166 (2015).
41. Struck, J. et al. Tunable gauge potential for neutral and spinless particles in 48. Tarnowski, M. et al. Measuring topology from dynamics by obtaining the Chern
driven optical lattices. Phys. Rev. Lett. 108, 225304 (2012). number from a linking number. Nat. Commun. 10, 1728 (2019).
42. Aidelsburger, M. et al. Realization of the Hofstadter Hamiltonian with ultracold 49. Fläschner, N. et al. Observation of dynamical vortices after quenches in a
atoms in optical lattices. Phys. Rev. Lett. 111, 185301 (2013). system with topology. Nat. Phys. 14, 265–268 (2018).
RESEARCH Letter

Extended Data Fig. 1 | Atomic level diagram. a, Three key electronic Third, a multichromatic field near the 5P3/2 → 5D5/2 transition modulates
transitions of 87Rb atoms enable the formation of Floquet polaritons. First, the energy of the 5P3/2 state. b, The multichromatic field has two
cavity photons near 780 nm couple to the 5S1/2 → 5P3/2 atomic transition. components with approximately opposite detunings ±δ: a red-detuned
Second, a beam near 480 nm drives the 5P3/2 → nS1/2 transition to the component with constant intensity Ir, and a blue-detuned component with
Rydberg level with principal quantum number n at coupling strength Ω. sinusoidally modulated intensity Ib[1 + cos(ωt)], where ω = 2πf.

You might also like