This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Flat-band localization in Creutz superradiance lattices

Yanyan He,1,∗ Ruosong Mao,1,∗ Han Cai,1,† Jun-Xiang Zhang,1,†† Yongqiang Li,2 Luqi Yuan,3 Shi-Yao Zhu,1 and Da-Wei Wang1,4 1Interdisciplinary Center for Quantum Information, State Key Laboratory of Modern Optical Instrumentation, and Zhejiang Province Key Laboratory of Quantum Technology and Device, Department of Physics, Zhejiang University, Hangzhou 310027, Zhejiang Province, China;
2Department of Physics, National University of Defense Technology, Changsha 410073, Hunan Province, China;
3State Key Laboratory of Advanced Optical Communication Systems and Networks, School of Physics and Astronomy, Shanghai Jiao Tong University, Shanghai 200240, China;
4CAS Center for Excellence in Topological Quantum Computation, University of Chinese Academy of Sciences, Beijing 100190, China
(January 27, 2025)
Abstract

Flat bands play an important role in diffraction-free photonics and attract fundamental interest in many-body physics. Here we report the engineering of flat-band localization of collective excited states of atoms in Creutz superradiance lattices with tunable synthetic gauge fields. Magnitudes and phases of the lattice hopping coefficients can be independently tuned to control the state components of the flat band and the Aharonov-Bohm phases. We can selectively excite the flat band and control the flat-band localization with the synthetic gauge field. Our study provides a room-temperature platform for flat bands of atoms and holds promising applications in exploring correlated topological materials.

Flat bands are characterized by the zero bandwidth over the whole Brillouin zone. Owing to the destructive interference between the hopping pathways [sutherland1986localization, , bergman2008band, ], the group velocity of excitations vanishes, and hence the diffusion in flat bands is inhibited. The resulting compact localized eigenstates (CLSs) [flach2014detangling, , leykam2018artificial, ] have been experimentally realized in photonics vicencio2015 , mukherjee2015a , mukherjee2015b , mukherjee2017 , mukherjee2018 , kremer2020 , xia2016 , xia2018 , ma2020 and polariton-exciton condensates baboux2016 , harder2020 . Immune to environmental noises, localized states in flat bands are promising candidates for realizing quantum networks rontgen2019 and diffraction-free photonics vicencio2013 , rojas2017 , yu2020 . Flat bands are also of fundamental interest in many-body physics because of their high degeneracy. The density of states is divergent such that even weak interactions lead to strong correlations and exotic topological phases bergholtz2013 , junemann2017exploring , spanton2018 , chen2020 .

Many-body interactions can be engineered to realize correlated topological phases in atoms bloch2008 , li2020 . However, in previous realizations of the flat bands in optical lattices, the underlying lattices jo2011 , taie2015 are gapless and topologically trivial. A feasible model that integrates both band flatness and topology is the two-leg ladder in a uniform magnetic field with cross-linked couplings, i.e., the Creutz lattice Creutz1999End , hugel2014 (see Fig.  1(a)). Despite theoretical proposals in photonic waveguides mukherjee2018 and ultracold atoms [junemann2017exploring, ], flat bands in the Creutz lattice have never been experimentally realized kang2018 , alaeian2019 , kang2020 .

Here we report the synthesis of a Creutz ladder with tunable tight-binding parameters in the form of a momentum-space lattice, i.e., the superradiance lattice [WangPRL2015, , Chen2018, ], in room-temperature cesium atoms. The bipartite ladder consists of timed Dicke states with different momenta [ScullyPRL2006, ]. We find that the corresponding energy band structure exhibits a flat band and a dispersive band, which are distinguished by localized and delocalized excitations, respectively. In the experiment, we excite one site in the ladder with a weak probe field and measure the optical response of the adjacent site. The hopping strengths and the Aharonov-Bohm (AB) phases in the lattice are carefully tuned, which enables us to excite a particular band and control the flat-band localization. We observe that the optical response is significantly suppressed when the flat band is selectively excited. By controlling the AB phases, we reveal the relation between the flat-band localization and gauge fields khomeriki2016 . Our work demonstrates a versatile platform for flat bands of atoms with multiple tunable parameters, which holds promising applications in exploring correlated topological phases.

We first introduce the experimental scheme implemented in the hyperfine levels of the 133Cs D1 line in a bichromatic standing-wave-coupled configuration, as shown in Fig.  1(a). Two standing waves couple

Refer to caption
Figure 1: (a) Schematic configuration of the experiment. The amplitude envelopes of the two standing waves have a ϕ/2\phi/2 phase difference. In the Creutz ladder, the arrows indicate the phase ϕ/2\phi/2 attached to the transitions. Inset is the configuration of atomic levels and laser fields. (b) Typical reflection spectrum. The insets shows the configurations of the fields and the lattice responses when νp\nu_{p} is near resonant with either atomic transitions. The AB flux ϕ=π\phi=\pi. The power of the probe field is 24 μW24\text{ $\mu$W}. The powers of each plane wave component of the two standing waves are P1=29 mWP_{1}=29\text{ mW} and P2=215 mWP_{2}=215\text{ mW} with effective Rabi frequencies Ω1=15 MHz\Omega_{1}=15\text{ MHz} and Ω2=68 MHz\Omega_{2}=68\text{ MHz}. The thick dark (thin light) lines are the experimental data (numerical simulation).

two excited states |a|62P1/2,F=3\left|a\right\rangle\equiv\left|6^{2}P_{1/2},F=3\right\rangle and |b|62P1/2,F=4\left|b\right\rangle\equiv\left|6^{2}P_{1/2},F=4\right\rangle to the same metastable state |c|62S1/2,F=3\left|c\right\rangle\equiv\left|6^{2}S_{1/2},F=3\right\rangle. The frequency of the jjth standing-wave coupling field νj\nu_{j} fulfills the two-photon resonance condition Δc=ν1ωac=ν2ωbc\Delta_{c}=\nu_{1}-\omega_{ac}=\nu_{2}-\omega_{bc}, where ωij\omega_{ij} being the atomic transition frequency between |i\left|i\right\rangle and |j\left|j\right\rangle. The envelopes of the Rabi frequency amplitude of the two standing waves are 2Ω1cos(kcxϕ/4)2\Omega_{1}\cos(k_{c}x-\phi/4) and 2Ω2cos(kcx+ϕ/4)2\Omega_{2}\cos(k_{c}x+\phi/4), where kck_{c} is the xx component of the wave vectors and ϕ/2\phi/2 is the phase difference between the envelopes. The wave-vector difference between the two standing waves is negligible in the length of the atomic vapor cell. We use a weak travelling field with the wave vector kpk_{p} to probe the standing-wave-coupled atomic vapor and measure the backward reflection. The frequency of the probe field νp\nu_{p} is scanned to couple the ground state |g|62S1/2,F=4\left|g\right\rangle\equiv\left|6^{2}S_{1/2},F=4\right\rangle to either the state |a\left|a\right\rangle or |b\left|b\right\rangle. Featured signals can be observed when the probe field is near resonant with each atomic transition. A typical spectrum is shown in Fig.  1(b).

In order to show that our experiment constructs a Creutz ladder and reveal the connection between the reflection signal and the excitation transport in the ladder, we write the Hamiltonian H=Hs+HpH=H_{s}+H_{p} in momentum space supp , with HsH_{s} and HpH_{p} being the parts of the Hamiltonian corresponding to the couplings of standing waves and the probe field, respectively. Here, we set =1\hbar=1 and HsH_{s} reads

Hs\displaystyle H_{s} =n[2t1anan+2t2bnbn\displaystyle=\sum_{n}[2t_{1}a_{n}^{\dagger}a_{n}+2t_{2}b_{n}^{\dagger}b_{n}
+(2t3cosϕ2anbn+t3anbn+1+t3bnan+1\displaystyle+(2t_{3}\cos\frac{\phi}{2}a_{n}^{\dagger}b_{n}+t_{3}a_{n}^{\dagger}b_{n+1}+t_{3}b_{n}^{\dagger}a_{n+1}
+t1eiϕ/2anan+1+t2eiϕ/2bnbn+1+h.c.)],\displaystyle+t_{1}e^{-i\phi/2}a_{n}^{\dagger}a_{n+1}+t_{2}e^{i\phi/2}b_{n}^{\dagger}b_{n+1}+h.c.)], (1)

which gives a tight-binding superradiance lattice composed of the collective atomic excitation operators dj=1/Nm|dmgm|exp[i(kp2jkc)xm]d_{j}^{\dagger}=\sqrt{1/N}\sum_{m}\left|d_{m}\right\rangle\left\langle g_{m}\right|\exp[i(k_{p}-2jk_{c})x_{m}] (d=a,bd=a,b) [WangPRL2015, ], where mm labels the mmth atom at the position xmx_{m}, jj is an integer, and NN is the total number of atoms. t1=Ω12/Δct_{1}=-\Omega_{1}^{2}/\Delta_{c} and t2=Ω22/Δct_{2}=-\Omega_{2}^{2}/\Delta_{c} are the hopping amplitudes along aa-leg and bb-leg, respectively. Here, we can adiabatically eliminate the state |c\left|c\right\rangle, since Δc\Delta_{c} is much larger than all relevant Rabi frequencies (ΔcΩj)(\Delta_{c}\gg\Omega_{j}). The two hoppings acquire a phase ϕ/2\phi/2 in opposite directions. The loop transition along a plaquette accumulates an AB phase ϕ\phi, such that the lattice is effectively in a uniform magnetic field. t3=Ω1Ω2/Δct_{3}=-\Omega_{1}\Omega_{2}/\Delta_{c} and 2t3cosϕ/22t_{3}\cos\phi/2 are the hopping strengths along the diagonals and the rungs of each plaquette in the ladder. The on-site energies of the aa-leg and bb-leg sites are 2t12t_{1} and 2t2t_{2}, respectively.

The probe field coupling Hamiltonian is Hp=NΩpeiΔpta0+NΩpeiΔptb0+h.c.H_{p}=\sqrt{N}\Omega_{p}e^{-i\Delta_{p}^{\prime}t}a_{0}^{\dagger}+\sqrt{N}\Omega_{p}^{\prime}e^{-i\Delta_{p}t}b_{0}^{\dagger}+h.c., where Ωp\Omega_{p} (Ωp\Omega_{p}^{\prime}) and Δp=νpωbg\Delta_{p}=\nu_{p}-\omega_{bg} (Δp=νpωag\Delta_{p}^{\prime}=\nu_{p}-\omega_{ag}) are the Rabi frequency and the frequency detuning of the coupling between the probe field and the atomic transition between |a\left|a\right\rangle (|b\left|b\right\rangle) and |g\left|g\right\rangle. Hence, HpH_{p} shows that the excitation is prepared by the probe field to the site a0a_{0} or b0b_{0} in the ladder. When we probe the site a0a_{0} (b0b_{0}), the phase-matching condition kp2kckpk_{p}-2k_{c}\approx-k_{p} is only satisfied for the excitation on the site a1a_{1} (b1b_{1}), which results in a superradiant backward emission collected by a photodetector. The spectrum in the left (right) of Fig.  1(b) characterizes the excitation transport from a0a_{0} to a1a_{1} (b0b_{0} to b1b_{1}) in the ladder of Eq.  (1). In the experiment, the probe field is weak (Ωpti)(\Omega_{p}\ll t_{i}) such that only a small fraction of the atoms are excited. In this condition, aja_{j}, bjb_{j} are approximately bosonic annihilation operators [WangPRL2015, ].

The Creutz ladder in Eq.  (1) constructed in momentum space is tunable in the experiment. We diagonalize HsH_{s} in real space supp and the band structures are shown in Fig.  2. We define ηt2/t1\eta\equiv t_{2}/t_{1} as the relative hopping strength along the two legs. One can see that all three band structures with different η\eta are composed of a flat band and a dispersive band. In general, the dynamics of the excitation is governed by both bands and cannot be distinguished. Probing only one band by controlling the excitation energy jacqmin2014 , drost2017 , slot2017 is inapplicable since the band gap closes when ϕ\phi approaches zero supp .

Nevertheless, the tunability of the hopping strengths enables us to determine which band to excite by controlling the state component of the bands. In the experiment, we tune η\eta, which is proportional to P2/P1P_{2}/P_{1}, where P1P_{1} and P2P_{2} are the powers of the two standing waves.

Refer to caption
Figure 2: Band structures (left) and the corresponding reflection spectra (right) with (a) η=20.7\eta=20.7, (b) η=4.0\eta=4.0, and (c) η=1/20.7\eta=1/20.7. The reflection spectrum is mainly contributed by the dispersive (flat) band for η1\eta\gg 1 (η1\eta\ll 1). The data in (c) is obtained by measuring the aa-leg response owing to the symmetry of the Hamiltonian (see the text). Dotted gray lines mark the energy of the flat band.

An interesting correlation can be noticed between the parameter η\eta and the band components, where the color represents the polarization σz|aa||bb|\left\langle\sigma_{z}\right\rangle\equiv\left\langle\left|a\right\rangle\left\langle a\right|-\left|b\right\rangle\left\langle b\right|\right\rangle of the eigenstates. In particular, σz=+1\left\langle\sigma_{z}\right\rangle=+1 or 1-1 means the band fully locates on the aa- or bb-leg. In Fig.  2(a), one can see that σz1\langle\sigma_{z}\rangle\approx-1 for almost the whole dispersive band, meaning that the dispersive band supports a large excitation component on the bb-leg for η1\eta\gg 1 (we take η=20.7\eta=20.7 according to the experimental parameters). Therefore, the b0b1b_{0}\rightarrow b_{1} transport dynamics is governed by the dispersive band. On the other hand, for η1\eta\ll 1 (η=1/20.7\eta=1/20.7 as shown in 2(c)), σz1\langle\sigma_{z}\rangle\approx-1 for almost the entire flat band, so the b0b1b_{0}\rightarrow b_{1} transport dynamics is governed by the flat band.

This band selection is manifested in the bandwidth, the central frequency, and the magnitude of the reflection spectrum in Fig.  2. In the experiment, we change η\eta and keep t3P1P2t_{3}\propto\sqrt{P_{1}P_{2}} a constant. In Fig.  2(a) for η1\eta\gg 1, the reflection spectrum of the dispersive band has a larger bandwidth and a lower central frequency. As a comparison, in Fig.  2(c) for η1\eta\ll 1, the reflection spectrum due to the flat band has a much narrower bandwidth and the peak locates near the predicted frequency of the flat band. The localization in the flat band is demonstrated by the decrease of the reflection peak when we decrease η\eta, during which the reflection is more contributed by the flat band.

As a side note, in obtaining Fig.  2(c), we use the symmetry that lattice Hamiltonian HsH_{s} is invariant when we exchange the sublattices aa and bb, inverse η\eta, and flip the flux ϕ\phi supp . In the experiment, the b0b1b_{0}\rightarrow b_{1} transport dynamics with ϕ\phi and η>1\eta>1 is characterized by the reflection spectrum near resonant with level |b\left|b\right\rangle (labelled with Rη(ϕ)R_{\eta}(\phi)), while the one with ϕ-\phi and (1/η)<1(1/\eta)<1 is effectively obtained from the |a\left|a\right\rangle-side reflection spectrum (labelled with R1/η(ϕ)R_{1/\eta}(-\phi)).

Refer to caption
Figure 3: Response of the dispersive and flat bands with different gauge fields. (a) The averaged reflectivity R¯\bar{R} versus ϕ\phi with different η\eta. (b) The normalized probability on the site b1b_{1} versus ϕ\phi. The diamonds (squares) indicate where the curves reach their maxima (minima). Δp(Δp)=2 MHz\Delta_{p}(\Delta_{p}^{\prime})=-2\text{ MHz} and Δc=233.5 MHz\Delta_{c}=233.5\text{ MHz}. The gray dashed lines are plotted to guide the extrema of the curves. (c) The probability distribution of the steady state |ψs\left|\psi_{s}\right\rangle on the bb-leg with ϕ1=0.64π\phi_{1}=0.64\pi, η=20.7\eta=20.7 (blue diamonds), ϕ2=1.62π\phi_{2}=1.62\pi, η=20.7\eta=20.7 (yellow squares), ϕ3=0.9π\phi_{3}=0.9\pi, η=1/20.7\eta=1/20.7 (red diamonds), and ϕ4=1.86π\phi_{4}=1.86\pi, η=1/20.7\eta=1/20.7 (green squares). The powers of each plane wave component of the two standing waves are P1=67 mWP_{1}=67\text{ mW}, P2=90 mWP_{2}=90\text{ mW} for η=3.8\eta=3.8 and 1/3.81/3.8; P1=40 mWP_{1}=40\text{ mW}, P2=153 mWP_{2}=153\text{ mW} for for η=10.6\eta=10.6 and 1/10.61/10.6; and P1=29 mWP_{1}=29\text{ mW}, P2=215 mWP_{2}=215\text{ mW} for η=20.7\eta=20.7 and 1/20.71/20.7. Other experimental parameters are the same as in Fig.  1. The points are simply connected for clarity. Error bars are obtained from four independent data sets.

The reflection spectrum is mostly contributed by the slowly moving atoms that have Doppler shifts smaller than the lattice bandwidth cai2019 . We take the average of reflectivity over the bands, i.e., R¯=R𝑑νp/𝑑νp\bar{R}=\int Rd\nu_{p}/\int d\nu_{p}, to investigate the localization and its dependence on ϕ\phi. In Fig.  3(a), the flat-band localization is demonstrated by the suppression of R¯\bar{R} when η\eta decreases. Furthermore, we notice that the ϕ\phi-dependence of R¯\bar{R} changes with η\eta (see the ϕ\phi calibration in supp ). The sinusoidal curve of averaged reflectivity R¯η(ϕ)\bar{R}_{\eta}(\phi) is shifted from top to bottom in Fig.  3(a).

The ϕ\phi-dependent shift shows the distinct responses to the gauge fields of the two bands. We consider two ideal cases to explain the physics. If the excitation is completely prepared in the dispersive band, the transport dynamics is determined by the flux-dependent unidirectional chiral edge current atala2014 , mancini2015 , stuhl2015 , livi2016 , anisimovas2016 , an2017 , cai2019 , dutt2020 of the dispersive band. The unidirectional chiral current breaks the symmetry between the transition from b0b_{0} to b1b_{1} and the one from b0b_{0} to b1b_{-1}. When the magnetic flux ϕ(0,π)\phi\in(0,\pi), the chiral current enhances the probability in the site b1b_{1}, and hence the reflectivity increases (vice versa). On the other hand, flat-band response to the gauge field can be understood by the CLSs |Fj\left|F_{j}\right\rangle [flach2014detangling, ]. When ϕ=2nπ\phi=2n\pi (nn is an integer), |Fj(ηajbj)|G\left|F_{j}\right\rangle\propto(\eta a_{j}^{\dagger}-b_{j}^{\dagger})\left|G\right\rangle is localized within the jjth unit cell. Therefore, only |F0\left|F_{0}\right\rangle is excited when we probe the site b0b_{0}, leading to the maximum localization. Otherwise, |Fj(ηaj+1eiϕ/2bj+1+ηeiϕ/2ajbj)|G\left|F_{j}\right\rangle\propto(\eta a_{j+1}^{\dagger}-e^{i\phi/2}b_{j+1}^{\dagger}+\eta e^{i\phi/2}a_{j}^{\dagger}-b_{j}^{\dagger})\left|G\right\rangle is localized within two unit cells. Probing site b0b_{0} leads to a coherent superposition between |F0\left|F_{0}\right\rangle and |F1\left|F_{-1}\right\rangle, and hence results in a finite overlap with b±1b_{\pm 1}, which is similar to the “breathing motions” in Refs. mukherjee2018 , kremer2020 .

Refer to caption
Figure 4: Gauge field dependence of the reflectivities from the dispersive band and flat band. (a) The phase difference φ\varphi between the ϕ\phi-dependences of R¯η\bar{R}_{\eta} and R¯1/η\bar{R}_{1/\eta}. (b) The Lissajous curves composed of R¯η(ϕ)\bar{R}_{\eta}(\phi) and R¯1/η(ϕ)\bar{R}_{1/\eta}(-\phi). The phase difference between R¯η\bar{R}_{\eta} and R¯1/η\bar{R}_{1/\eta} increases and approaches π/2\pi/2 when η\eta increases. The yellow dots are experimental data and the green lines are fitted elliptic curves. The powers of each plane wave component of the two standing waves are P1=50 mWP_{1}=50\text{ mW}, P2=124 mWP_{2}=124\text{ mW} for η=6.8\eta=6.8 and 1/6.81/6.8; P1=33 mWP_{1}=33\text{ mW}, P2=183 mWP_{2}=183\text{ mW} for η=15.2\eta=15.2 and 1/15.21/15.2. Other experimental parameters are the same as in Fig.  3. Each plot contains 400 data points. (c) The mean averaged reflectivity R¯η\|\bar{R}\|_{\eta} versus η\eta. R¯η\|\bar{R}\|_{\eta} increases with η\eta monotonically. The dotted lines are the simulated results.

The two responses can be further investigated by the steady state of the collective excited states of atoms, where the wave function |ψs[1+j(αjaj+βjbj)]|G\left|\psi_{s}\right\rangle\approx[1+\sum_{j}(\alpha_{j}a_{j}^{\dagger}+\beta_{j}b_{j}^{\dagger})]\left|G\right\rangle (αj,βj1)(\alpha_{j},\beta_{j}\ll 1) in the weak excitation approximation. In the steady state, we obtain the probability amplitude βj=NΩpG|bj(Δp+iγ^Hs)1b0|G\beta_{j}=\sqrt{N}\Omega_{p}\left\langle G\right|b_{j}(\Delta_{p}+i\hat{\gamma}-H_{s})^{-1}b_{0}^{\dagger}\left|G\right\rangle [ozawa2014anomalous, ], where γ^=j(γaajaj+γbbjbj)\hat{\gamma}=\sum_{j}(\gamma_{a}a_{j}^{\dagger}a_{j}+\gamma_{b}b_{j}^{\dagger}b_{j}) and γa\gamma_{a} (γb\gamma_{b}) is the decoherence rate between the excited state |a\left|a\right\rangle (|b\left|b\right\rangle) and |g\left|g\right\rangle. Especially, we plot |β1/β0|2\left|\beta_{1}/\beta_{0}\right|^{2}, i.e., the normalized probability on the site b1b_{1} in Fig.  3(b). Since the reflectivity is approximately proportional to |β1/β0|2\left|\beta_{1}/\beta_{0}\right|^{2} [Wang2015, ], both features of the R¯\bar{R} in Fig.  3(a) are demonstrated, including the magnitude suppression and the ϕ\phi-dependence. We also plot four typical probability amplitude distributions along the bb-leg in Fig.  3(c) with η1\eta\gg 1 and η1\eta\ll 1. For η=20.7\eta=20.7, the distributions of |βj|2|\beta_{j}|^{2} with ϕ1=0.64π\phi_{1}=0.64\pi and ϕ2=1.62π\phi_{2}=1.62\pi are almost symmetric to each other with respect to the 0th site, implying flux-dependent unidirectional chiral edge current of the dispersive band. However, the distributions with ϕ3=0.9π\phi_{3}=0.9\pi and ϕ4=1.86π\phi_{4}=1.86\pi are symmetrically localized at b0b_{0} with different localization lengths. |β1/β0|2\left|\beta_{1}/\beta_{0}\right|^{2} for η=1/20.7\eta=1/20.7 is minimized at ϕ4\phi_{4}, which means that the eigenstates are maximally localized when the synthetic gauge field almost vanishes. We notice that ϕ4\phi_{4} is slightly different from 2nπ2n\pi, which comes from the suppressed but finite contribution from the dispersive band.

In Fig.  4(b), the different ϕ\phi-dependences of the dispersive and flat bands are illustrated with Lissajous curves. The spatial phase difference between the envelopes of the two standing waves is slowly tuned to cover all values of ϕ\phi. We obtain the data sets {R¯η(ϕ),R¯1/η(ϕ)\bar{R}_{\eta}(\phi),\bar{R}_{1/\eta}(-\phi)} and fit them with ellipses to reconstruct the Lissajous curves. The shape of the ellipse elucidates the phase differences between the two underlying functions. For example, a Lissajous curve composed of two parametric equations with argument uu, e.g., x=sin(u)x=\sin(u) and y=sin(u+φ)y=\sin(-u+\varphi), is a line (circle) when φ=0\varphi=0 (π/2\pi/2). In Fig.  4(a), φ\varphi obtained by fitting the Lissajous curve approaches π/2\pi/2 when η\eta increases, indicating the different types of the excitations on the flat and dispersive bands. In Fig.  4(c), we notice that R¯η\|\bar{R}\|_{\eta} increases with η\eta monotonically, where .\|.\| indicates the mean value of the averaged reflectivity R¯\bar{R} over all ϕ\phi. The numerical simulation agrees with the data.

In conclusion, we experimentally realize Creatz ladders with tunable gauge fields, where the flat band can be selectively excited and the interplay between the flatband localization and the AB phases was investigated. We study the flat-band localization in an open system, where the steady state balanced by pumping, driving, and dissipation exhibits the dynamics in the corresponding closed system [ozawa2014anomalous, ]. We also need to emphasize that our scheme is substantially different from the incoherently pumped polariton-exciton condenstates baboux2016 , harder2020 , masumoto2012 , klembt2017 , whittaker2018 , where coherence is not accessible between multiple CLSs. It is interesting to notice that both bands of the Creutz ladder are topologically non-trivial kremer2020 , kang2020 provided that ϕ2nπ\phi\neq 2n\pi supp . It is a step towards the simulation of the strong correlated quantum phases, including the fractional Chern insulators bergholtz2013 , disorder-free many-body localization kuno2020 , and unusual ferromagnetism tasaki1992 . An interaction term between the sites in momentum space can be introduced by weakly coupling the excited atomic level to a Rydberg state li2020 or by ss-wave interactions an2018 , xie2020 . With a negative Δc\Delta_{c}, the flat band has the lowest energy and can be used to study the many-body ground states of ultracold atoms li2020 , supp .

We acknowledge the support from the National Natural Science Foundation of China (Grants No. 11934011, No. 11874322, No. 91736209 and No. U1330203), the National Key Research and Development Program of China (Grants No. 2019YFA0308100 and No. 2018YFA0307200), and the Fundamental Research Funds for the Central Universities.

  

These authors contributed equally to this work.

††junxiang_\_[email protected]

References

  • [1] B. Sutherland, Localization of electronic wave functions due to local topology, Phys. Rev. B 34, 5208 (1986).
  • [2] D. L. Bergman, C. Wu, and L. Balents, Band touching from real-space topology in frustrated hopping models, Phys. Rev. B 78, 125104 (2008).
  • [3] S. Flach, D. Leykam, J. D. Bodyfelt, P. Matthies, and A. S. Desyatnikov, Detangling flat bands into Fano lattices, Europhys. Lett. 105, 30001 (2014).
  • [4] D. Leykam, A. Andreanov, and S. Flach, Artificial flat band systems: from lattice models to experiments, Advances in Physics: X 3, 1473052 (2018).
  • [5] R. A. Vicencio, C. Cantillano, L. Morales-Inostroza, B.  Real, C. Mejía-Cortés, S. Weimann, A. Szameit, and M. I. Molina, Observation of localized states in Lieb photonic lattices, Phys. Rev. Lett. 114, 245503 (2015).
  • [6] S. Mukherjee, A. Spracklen, D. Choudhury, N. Goldman, P. Öhberg, E. Andersson, and R. R. Thomson, Observation of a localized flat-band state in a photonic Lieb lattice, Phys. Rev. Lett. 114, 245504 (2015).
  • [7] S. Mukherjee and R. R. Thomson, Observation of localized flat-band modes in a quasi-one-dimensional photonic rhombic lattice, Opt. Lett. 40, 5443 (2015).
  • [8] S. Mukherjee and R. R. Thomson, Observation of robust flat-band localization in driven photonic rhombic lattices, Opt. Lett. 42, 2243 (2017).
  • [9] S. Mukherjee, M. D. Liberto, P. Öhberg, R. R.Thomson, and N. Goldman, Experimental Observation of Aharonov-Bohm Cages in Photonic Lattices, Phys. Rev. Lett. 121, 075502 (2018).
  • [10] M. Kremer, I. Petrides, E. Meyer, M. Heinrich, O. Zilberberg, and A. Szameit, A square-root topological insulator with non-quantized indices realized with photonic Aharonov-Bohm cages, Nat. Commum. 11, 1808 (2020).
  • [11] S. Xia, Y. Hu, D. Song, Y. Zong, L. Tang, and Z. Chen, Demonstration of flat-band image transmission in optically induced Lieb photonic lattices, Opt. Lett. 41, 1435 (2016).
  • [12] S. Xia, D. Li, X. Liu, L. Tang, Y. Hu, D. Song, J. Xu, D. Leykam, S. Flach, and Z. Chen, Unconventional flatband line states in photonic Lieb lattices, Phys. Rev. Lett. 121, 263902 (2018).
  • [13] J. Ma, J. W. Rhim, L. Tang, S. Xia, H. Wang, X. Zheng, S. Xia, D. Song, Y. Hu, Y. Li, B. J. Yang, D. Leykam, and Z. Chen, Direct observation of flatband loop states arising from nontrivial real-space topology, Phys. Rev. Lett. 124, 183901 (2020).
  • [14] F. Baboux, L. Ge, T. Jacqmin, M. Biondi, E. Galopin, A. Lemaître, L. Le Gratiet, I. Sagnes, S. Schmidt, H. E. Türeci, A. Amo, and J. Bloch, Bosonic condensation and disorder-induced localization in a flat band, Phys. Rev. Lett. 116, 066402 (2012).
  • [15] T. H. Harder, O. A. Egorov, J. Beierlein, P. Gagel, J. Michl, M. Emmerling, C. Schneider, U. Peschel, S. Höfling, and S. Klembt, Exciton-polaritons in flatland: Controlling flatband properties in a lieb lattice, arXiv: 2002.05405.
  • [16] M. Röntgen, C. Morfonios, I. Brouzos, F. Diakonos, and P. Schmelcher, Quantum network transfer and storage with compact localized states induced by local symmetries, Phys. Rev. Lett. 123, 080504 (2019).
  • [17] S. Rojas-Rojas, L. Morales-Inostroza, R. A. Vicencio, and A. Delgado, Quantum localized states in photonic flat-band lattices, Phys. Rev. A 96, 043803 (2017).
  • [18] R. A. Vicencio and C. Mejía-Cortés, Diffraction-free image transmission in kagome photonic lattices, Journal of Optics 16, 015706 (2013).
  • [19] D. Yu, L. Yuan, and X. Chen, Isolated photonic flatband with the effective magnetic flux in a synthetic space including the frequency dimension, arXiv: 2004.12542.
  • [20] E. J. Bergholtz and Z. Liu, Topological flat band models and fractional Chern insulators, Int. J. Mod. Phys. B 27, 1330017 (2013).
  • [21] J. Jünemann, A. Piga, S. J. Ran, M. Lewenstein, M. Rizzi, and A. Bermúdez, Exploring interacting topological insulators with ultracold atoms: The synthetic Creutz-Hubbard model, Phys. Rev. X 7, 031057 (2017).
  • [22] E. M. Spanton, A. A. Zibrov, H. Zhou, T. Taniguchi, K. Watanabe, M. P. Zaletel, and A. F. Young, Observation of fractional Chern insulators in a van der Waals heterostructure, Science 360, 62 (2018).
  • [23] G. Chen, A. L. Sharpe, E. J. Fox, Y. H. Zhang, S. Wang, L. Jiang, B. Lyu, H. Li, K. Watanabe, T. Taniguchi, Z. Shi, T. Senthil, D. Goldhaber-Gordon, Y. Zhang, and F. Wang, Tunable correlated Chern insulator and ferromagnetism in a moiré superlattice, Nature 579, 56 (2020).
  • [24] I. Bloch, J. Dalibard, and W. Zwerger, Many-body physics with ultracold gases, Rev. Mod. Phys. 80, 885 (2008).
  • [25] Y. Li, H. Cai, D. W. Wang, L. Li, J. Yuan, and W. Li, Many-body chiral edge currents and sliding phases of atomic spin waves in momentum-space lattice, Phys. Rev. Lett. 124, 140401 (2020).
  • [26] G. B. Jo, J. Guzman, C. K. Thomas, P. Hosur, A. Vishwanath, and D. M. Stamper-Kurn, Ultracold atoms in a tunable optical kagome lattice, Phys. Rev. Lett. 108, 045305 (2011).
  • [27] S. Taie, H. Ozawa, T. Ichinose, T. Nishio, S.  Nakajima, and Y. Takahashi, Coherent driving and freezing of bosonic matter wave in an optical Lieb lattice, Sci. Adv. 1, e1500854 (2015).
  • [28] M. Creutz, End states, ladder compounds, and domain-wall fermions, Phys. Rev. Lett. 83, 263617 (1999).
  • [29] D. Hügel and B. Paredes, Chiral ladders and the edges of quantum Hall insulators, Phys. Rev. A 89, 023619 (2014).
  • [30] J. H. Kang, J. H. Han, and Y. Shin, Realization of a cross-linked chiral ladder with neutral fermions in a 1D optical lattice by orbital-momentum coupling, Phys. Rev. Lett. 121, 150403 (2018).
  • [31] J. H. Kang, J. H. Han, and Y. Shin, Creutz ladder in a resonantly shaken 1D optical lattice, New J. Phys. 22, 013023 (2020).
  • [32] H. Alaeian, C. W. S. Chang, M. V. Moghaddam, C.  M. Wilson, E. Solano, and E. Rico, Creating lattice gauge potentials in circuit QED: The bosonic Creutz ladder, Phys. Rev. A 99, 053834 (2019).
  • [33] D. W. Wang, R. B. Liu, S. Y. Zhu, and M. O. Scully, Superradiance lattice, Phys. Rev. Lett. 114, 043602 (2015).
  • [34] L. Chen, P. Wang, Z. Meng, L. Huang, H. Cai, D. W. Wang, S. Y. Zhu, and J. Zhang, Experimental observation of one-dimensional superradiance lattices in ultracold atoms, Phys. Rev. Lett. 120, 193601 (2018).
  • [35] M. O. Scully, E. S. Fry, C. H. Ooi, and K. Wódkiewicz, Directed spontaneous emission from an extended ensemble of N atoms: Timing is everything, Phys. Rev. Lett. 96, 10501 (2006).
  • [36] R. Khomeriki and S. Flach, Landau-Zener Bloch oscillations with perturbed flat bands, Phys. Rev. Lett. 116, 245301 (2016).
  • [37] See Supplemental Material, which includes Refs. wang2013 , li2020 , an2018 , xie2020 .
  • [38] D. W. Wang, H. T. Zhou, M. J. Guo, J. X. Zhang, J. Evers, and S. Y. Zhu, Optical diode made from a moving photonic crystal Phys. Rev. Lett. 110, 093901 (2013).
  • [39] F. A. An, E. J. Meier, J. Ang’ong’a, and B. Gadway, Correlated dynamics in a synthetic lattice of momentum states, Phys. Rev. Lett. 120, 040407 (2018).
  • [40] D. Xie, T. S. Deng, T. Xiao, W. Gou, T. Chen, W. Yi, and B. Yan, Topological quantum walks in momentum space with a Bose-Einstein condensate, Phys. Rev. Lett. 124, 050502 (2020).
  • [41] T. Jacqmin, I. Carusotto, I. Sagnes, M. Abbarchi, D. D. Solnyshkov, G. Malpuech, E. Galopin, A. Lemaître, J. Bloch, and A. Amo, Direct Observation of Dirac cones and a flatband in a honeycomb lattice for polaritons Phys. Rev. Lett. 112, 116402 (2014).
  • [42] R. Drost, T. Ojanen, A. Harju, and P. Liljeroth, Topological states in engineered atomic lattices, Nat. Phys. 13, 668 (2017).
  • [43] M. R. Slot, T. S. Gardenier, P. H. Jacobse, G. C. van Miert, S. N. Kempkes, S. J. Zevenhuizen, C. M. Smith, D. Vanmaekelbergh, and I. Swart, Experimental realization and characterization of an electronic Lieb lattice, Nat. Phys. 13, 672 (2017).
  • [44] H. Cai, J. Liu, J. Wu, Y. He, S. Y. Zhu, J. X. Zhang, and D. W. Wang, Experimental observation of momentum-space chiral edge currents in room-temperature atoms, Phys. Rev. Lett. 122, 023601 (2019).
  • [45] M. Atala, M. Aidelsburger, M. Lohse, J. T. Barreiro, B. Paredes, and I. Bloch, Observation of chiral currents with ultracold atoms in bosonic ladders, Nat. Phys. 10, 588 (2014).
  • [46] M. Mancini, G. Pagano, G. Cappellini, L. Livi, M. Rider, J. Catani, C. Sias, P. Zoller, M. Inguscio, M. Dalmonte, L. Fallani, Observation of chiral edge states with neutral fermions in synthetic Hall ribbons, Science 349, 1510 (2015).
  • [47] B. Stuhl, H. I. Lu, L. Aycock, D. Genkina, and I. Spielman, Visualizing edge states with an atomic Bose gas in the quantum Hall regime, Science 349, 1514 (2015).
  • [48] L. F. Livi, G. Cappellini, M. Diem, L. Franchi, C. Clivati, M. Frittelli, F. Levi, D. Calonico, J. Catani, M. Inguscio, and L. Fallani, Synthetic dimensions and spin-orbit coupling with an optical clock transition, Phys. Rev. Lett. 117, 220401 (2016).
  • [49] E. Anisimovas, M. Račiūnas, C. Sträter, A. Eckardt, I. B.Spielman, and G. Juzeliu¯{\bar{\text{u}}}nas, Semisynthetic zigzag optical lattice for ultracold bosons, Phys. Rev. A 94, 063632 (2016).
  • [50] F. A. An, E. J. Meier, and B. Gadway, Direct observation of chiral currents and magnetic reflection in atomic fluxLattices, Sci. Adv. 3, e1602685 (2017).
  • [51] A. Dutt, Q. Lin, L. Yuan, M. Minkov, M. Xiao, and S. Fan, A single photonic cavity with two independent physical synthetic dimensions, Science 367, 59 (2020).
  • [52] T. Ozawa and I. Carusotto, Anomalous and quantum Hall effects in lossy photonic lattices, Phys. Rev. Lett. 112, 133902 (2014).
  • [53] D. W. Wang, S. Y. Zhu, J. Evers, and M. O. Scully, High-frequency light reflector via low-frequency light control, Phys. Rev. A 91, 011801 (2015).
  • [54] N. Masumoto, N. Y. Kim, T. Byrnes, K. Kusudo, A. Löffler, S. Höfling, A. Forchel, and Y. Yamamoto, Exciton–polariton condensates with flat bands in a two-dimensional kagome lattice, New J. Phys. 14, 065002 (2012).
  • [55] S. Klembt, T. H. Harder, O. A. Egorov, K. Winkler, H. Suchomel, J. Beierlein, M. Emmerling, C. Schneider, and S. Höfling, Polariton condensation in S- and P-flatbands in a two-dimensional Lieb lattice, Appl. Rev. Lett. 111, 231102 (2017).
  • [56] C. Whittaker, E. Cancellieri, P. Walker, D. Gulevich, H. Schomerus, D. Vaitiekus, B. Royall, D. Whittaker, E. Clarke, I. Iorsh, I. A. Shelykh, M. S. Skolnick, and D. N. Krizhanovskii, Exciton polaritons in a two-dimensional lieb lattice with spin-orbit coupling, Appl. Phys. Lett. 120, 097401 (2018).
  • [57] Y. Kuno, T. Orito, and I. Ichinose, Flat-band many-body localization and ergodicity breaking in the Creutz ladder, New J. Phys. 22, 013032 (2020).
  • [58] H. Tasaki, Ferromagnetism in the Hubbard models with degenerate single-electron ground states, Phys. Rev. Lett. 69, 1608 (1992).