Skip to main content
Light, Science & Applications logoLink to Light, Science & Applications
. 2022 Oct 9;11:291. doi: 10.1038/s41377-022-00990-7

Measuring Zak phase in room-temperature atoms

Ruosong Mao 1,#, Xingqi Xu 1,#, Jiefei Wang 1,#, Chenran Xu 1, Gewei Qian 1, Han Cai 1,2,, Shi-Yao Zhu 1,3, Da-Wei Wang 1,3,4,
PMCID: PMC9548506  PMID: 36210366

Abstract

Cold atoms provide a flexible platform for synthesizing and characterizing topological matter, where geometric phases play a central role. However, cold atoms are intrinsically prone to thermal noise, which can overwhelm the topological response and hamper promised applications. On the other hand, geometric phases also determine the energy spectra of particles subjected to a static force, based on the polarization relation between Wannier-Stark ladders and geometric Zak phases. By exploiting this relation, we develop a method to extract geometric phases from energy spectra of room-temperature superradiance lattices, which are momentum-space lattices of timed Dicke states. In such momentum-space lattices the thermal motion of atoms, instead of being a source of noise, provides effective forces which lead to spectroscopic signatures of the Zak phases. We measure Zak phases directly from the anti-crossings between Wannier-Stark ladders in the Doppler-broadened absorption spectra of superradiance lattices. Our approach paves the way of measuring topological invariants and developing their applications in room-temperature atoms.

Subject terms: Quantum optics, Optical techniques


Atoms moving through standing-wave lasers accumulate Zak phases, which have spectroscopic signatures, enabling room-temperature geometric phase reconstruction in momentum-space superradiance lattices.

graphic file with name 41377_2022_990_Figa_HTML.jpg

Introduction

Topological matter has promising applications in noise resilient devices and quantum information processing1,2, thanks to the robust topological response guaranteed by global geometric quantities of the Bloch bands, namely, the topological invariants36. These invariants change stepwisely only when the bulk goes through a topological phase transition, which involves band gap closing and reopening. Characterizing topological invariants is a central task in synthesizing and simulating topological phases of matter. They are usually measured by the response from gapless edge states based on the bulk-edge correspondence. However, edges are not always available in atomic quantum simulators7. On the other hand, the topological invariants are proportional to the geometric phases accumulated across a whole Brillouin zone. We can also measure the geometric phases from the bulk energy bands to obtain the topological invariants. Along this line techniques based on reciprocal-space interference8, quench dynamics912, and Hall transport13,14 have been developed in atomic simulators.

Previous experiments of determining geometric phases from bulk response in ultracold atoms rely on dynamic evolution or adiabatic manipulation916. It has been shown that geometric phases can be obtained from the energy spectra of electrons in a constant force1518, which turns the Bloch energy bands into Wannier-Stark ladders (WSLs) with equidistant discrete energies. In one-dimensional (1D) systems, the displacement of the energies of the WSLs is proportional to the applied force and the positions of the Wannier centers (WCs)19, which reflect the values of one-dimensional geometric phases, i.e., the Zak phases3 of the energy bands2022, as schematically illustrated in Fig. 1a, b. Here we develop and implement such a spectroscopic method to retrieve Zak phases. We show that this method enables the determination of Zak phases in room-temperature atoms, which greatly improves the accessibility of topological matters and facilitates their applications.

Fig. 1. Relation between Zak phases and Wannier-Stark ladders in superradiance lattices.

Fig. 1

a The Zak phase as an accumulated geometric phase of a particle adiabatically driven across a whole Brillouin zone. b The Wannier function in each unit cell. The zeroth unit cell is highlighted with yellow color. The positions of the Wannier centers are 2n+θ/π. c Schematic WSLs shown by the projected density of states with different forces F=δ. The three dotted lines connect the corresponding WSLs in different forces. Quantum transport package Nanoskim54 is used in the calculation. d Schematic configuration of the light fields. Inset: the coupling between the light fields and the atomic levels in the reference frame of the atom with the Rabi frequency Ωi (i=1,2,p), the coupling field detuning in the lab reference frame Δc, and the Doppler shift δ. The shaded area indicates the envelope of the standing wave coupling field and the balls indicate atoms with velocity v. e The momentum-space superradiance lattice with tight-binding Hamiltonian Hs (upper) and a linear potential Hf (lower). d~jd=a,b are the timed Dicke states

We reconstruct the Zak phases of the Rice-Mele (RM) model through the anti-crossing between the WSLs in a 1D tight-binding lattice2325 of timed Dicke states26:

bk=1Nmeikrmg1,g2,,bm,,gN 1

i.e., single-photon collective excitations of an ensemble of N atoms (here rm is the position of the mth atom with the ground state gm and an excited state bm), which can be created by coherent laser fields that transfer momentum ħk to the atoms. We introduce multiple laser fields to couple b to another atomic state a such that timed Dicke states with discrete k values form a momentum-space tight-binding lattice, which is coined the superradiance lattice. When the momentum of a timed Dicke state matches that of light, directional superradiant emission of radiation can be observed27, which provides a convenient way to measure the lattice transport. A substantial difference between the momentum-space superradiance lattice and conventional real-space lattices in solids is that the Brillouin zone (BZ) of the superradiance lattice is in real space, where atoms at different positions can be independently diagonalized by only considering the local field strengths of a spatially periodic coupling field (here a standing wave). The positions of atoms play the same role of the lattice momenta of electrons in solids. A remarkable consequence is that atoms in motion travel through the real-space BZs periodically, following the same dynamics of electrons subjected to a constant electric field. Therefore, atomic motion provides an effective electric field (or more precisely a constant force) for the excitations in superradiance lattices.

The Zak phases of a tight-binding lattice can be obtained from the WSLs when an effective force is introduced in the lattice. Since the motion of atoms provides such an effective force in superradiance lattices, we can take advantage of the thermal motion of the atoms to read out the Zak phases from the energy spectra. In particular, different velocity groups of thermal atoms provide a set of continuous values of the effective force, which results in a set of WSLs with displaced energies proportional the Zak phases15,16 (see Fig. 1c). A key to extract the Zak phases from the absorption spectra of superradiance lattices is that WSLs from the two energy bands have anti-crossings when they approach the same energy, which results in absorption peaks and dips. The Zak phases can be obtained from simple geometric relations between the locations of the anti-crossing points and the band centers. We investigate in detail two celebrated versions of the RM model, the Semenoff insulator and the Su-Schrieffer-Heeger (SSH) model28,29. We also demonstrate the Zak phase reconstruction for general RM models. Our method of measuring Zak phase in 1D systems can be generalized to identify topological invariants in higher dimensions3034.

Results

Experimental setup and model

We perform the experiment with the hyperfine levels of the 87Rb D1 line in a standing-wave-coupled EIT configuration, as shown in Fig. 1d (see the complete setup in Supplementary Note 1). A weak probe field propagating in x direction couples the ground state g52S1/2,F=1 to the excited state b52P1/2,F=2. The excited state is also coupled to a metastable state a52S1/2,F=2 by two strong counter-propagating light fields, forming a 1D bipartite superradiance lattice2325. The absorption spectra of the probe field are used to obtain the Zak phases of the superradiance lattices.

The total Hamiltonian of the superradiance lattice is H=Hs+Hp+Hf, where Hs and Hp are interaction Hamiltonians involving the coupling and probe fields, and Hf is the linear potential induced by atomic motion (see Fig. 1e). Hs is the tight-binding Hamiltonian of the RM superradiance lattices2325 (we set ħ=1 and see Materials and Methods):

Hs=jΔca^2ja^2j+a^2jΩ1b^2j+1+Ω2b^2j1+H.c. 2

where Ω1 and Ω2 are the Rabi frequencies of the co-propagating and counter-propagating coupling fields, and Δc=vcωba with vc being the coupling field frequency and ωba being the transition frequency between states a and b. Here d^j=1/Nmdmgmexpikjxmd=a,b is the creation operator of the timed-Dicke state26 d~jd^jg1,g2,,gN with wave vector kj=kp+j1kcjkc (j is an integer) and kp (kc) being the probe (coupling) field wave vector amplitude, the index m labels the mth atom at the position xm, N is the total number of atoms within the velocity range vΓ/2kc,v+Γ/2kc where v satisfies the Maxwell distribution and Γ is the decay rate of the state b (we neglect the decay of the hyperfine ground state a). The timed Dicke state b~1 in the superradiance lattice can be created from the ground state by Hp=NΩpeiΔptb^1+H.c., where the probe detuning Δp=vpωbg with vp being the probe field frequency and ωbg being the transition frequency between states b and g.

In order to clarify the effect of atomic motion, we show the contribution from atoms with different velocities in x direction. For atoms in each velocity group, the opposite Doppler shifts of the two coupling fields lead to a linear potential (see Fig. 1e) in momentum space23,35:

Hf=δj2ja^2ja^2j+2j+1b^2j+1b^2j+1 3

where the Doppler shift δkcvkpv with v being the velocity of the atoms in x direction.

Wannier-Stark ladders

The energy spectrum of the Hamiltonian Hs+Hf is closely related to the WCs, which are the expected positions of the Wannier functions22 in unit cells. The WCs in the nth unit cell r±n for the upper (+) and lower () energy bands of Hs are related to the geometric Zak phases by (in unit of distance between neighboring lattice sites, see Fig. 1b)20,21:

r±n=2n+θ±/π 4

where the Zak phases θ±i0π/kcdxu±xxu±x with u±x being the periodic Bloch functions of Hs in real space and the integration is over the whole Brillouin zone. Therefore, the Zak phases are the fractional parts of the corresponding WCs20,21.

The extended Bloch energy spectra split into discrete WSLs15,16,35 with energy spacing proportional to the static force δ when Hf is perturbative (see Fig. 1c):

E±nδ=ϵ±+r±nδ 5

where ϵ± denote the energies of the Bloch band centers (bc), defined as the average band energies of Hs (See Supplementary Note 2 for the derivation of Eq. (5) and discussion on its validity). From Eqs. (4) and (5), the Zak phases are obtained by θ±=E±[n]/δ2nπ.

The relation in Eq. (5) can be seen in the upper panels of Fig. 2a, b as functions of δ for two different RM lattices, namely, the Semenoff insulator with θ=π,θ+=0 (Fig. 2a), and the topological SSH model with half-integer Zak phases θ±=0.5π (Fig. 2b). Since the Zak phase is gauge-dependent7, its value depends on the choice of the unit cell in conventional lattices8. However, in our experiments the Zak phase is an observable with a fixed gauge set by the zero-energy site of Hf, which is determined by the Doppler shifts of atoms. This is a significant difference between SLs and conventional lattices3 (see Supplementary Note 3).

Fig. 2. Wannier-Stark ladders and the absorption spectra.

Fig. 2

a, b The upper panels are numerical simulation of the PDOS as functions of the Doppler shift δ and probe detuning Δp. The averaged PDOS in the lower panels are obtained from upper panels by integrating δ (in Maxwell distribution with FWHM 500 MHz). c, d The experimental data of the absorption (1Pt/Pi) and reflection (Pr/Pi) spectra, where Pi, Pt, and Pr are the power of the incident, transmitted, and reflected probe fields, respectively. a, c The Semenoff insulator with Ω1=Ω2=120 MHz and Δc=298 MHz. b, d The SSH model (Δc=0) with Ω1=125 MHz, Ω1/Ω2=5.3. The white dashed lines indicate the uncoupled WSLs in Eq. (5). The highlighted numbers denote the values of r[n] (square) and r+[m] (round) of the corresponding WSLs. Both in the simulated averaged PDODs and the measured absorption spectra, the dips and peaks capture the band centers (denoted by ϵ±) and anti-crossing points (denoted by Δn,m), where the blue dotted lines are used to guide eyes. The arrows point to the local extrema of the corresponding spectral features

The energy spectra of the SLs are shown by the projected density of states (PDOS) on the state b~1, i.e., lδDΔpElψlb~12, where δDΔpEl is the Dirac delta function, ψl and El are the eigenstates and eigenenergies satisfying Hs+Hfψl=Elψl (see Materials and Methods). In the weak force regime where the coupling between the WSLs from different bands are negligible, the spectra follow the linear dependence in Eq. (5) as indicated by the dashed lines. The brightest ladders in the absorption spectra are the ones corresponding to r±[0]=θ±/π in the 0th unit cell, which contains the state b~1.

Anti-crossing of Wannier-Stark ladders

When a pair of WSLs from different bands (E[n] and E+[m]) have the same energy for a δ, the interband coupling removes their degeneracy and results in an anti-crossing36 denoted by acn,m. Their positions in the energy-force diagram can be estimated by the degeneracy points of the uncoupled WSLs37 satisfying Enδ=E+mδ=Δn,m, where Δn,m is the probe detuning of the corresponding anti-crossing point. The values of Δn,m obtained from the experimental absorption spectra are the key to extract the Zak phases.

The optical responses (reflection and absorption) of the superradiance lattice are contributed by all atoms in Maxwell velocity distribution38. We obtain the averaged PDOS in the lower panels of Fig. 2a, b from the corresponding WSLs in the upper panels by integrating δ, which has a Doppler width about 500 MHz and covers all relevant values for the Zak phase reconstruction. We need to emphasize here that our scheme only requires that the velocity distribution shall be large enough to cover all the relevant anti-crossings. The Maxwell distribution of room-temperature atoms satisfies such a requirement (see the experimental spectra at different temperatures in Supplementary Note 4). The method is equally valid for other velocity distributions, as well as for cold atoms whose velocities can be precisely controlled.

Since the absorption coefficient is proportional to the PDOS23, the anti-crossings and band centers modify the PDOS drastically and their signatures can be identified in the absorption spectra. The values of Δn,m and ϵ± are experimentally measured with the corresponding extrema in the absorption spectra (see exemplary data sets in Supplementary Note 5). As shown in Fig. 2c (and more examples in Fig. 4), the spectra of the two-band SLs are generally featured with four dips, of which two are associated with the band centers and the rest two are due to anti-crossings. The band centers are generally characterized by dips in the far left and far right wings of the spectra. Only in a special case with zero Zak phase, a band center is featured by a peak in Fig. 2a (see Supplementary Note 6). Between the two band centers, the anti-crossings of WSLs lead to dips in the spectra, reflecting the energy gaps of the anti-crossings. Since we measure the PDOS of the state b~1 in the 0th unit cell, the major anti-crossings are associated to the Wannier functions localized in the 0th and the neighboring −1st unit cells, i.e., Δ1,0 and Δ0,1.

Fig. 4. Zak phase reconstruction of Rice-Mele superradiance lattices.

Fig. 4

a The Zak phase diagram of the RM model as a function of Δc and Ω1Ω2. b Locations of the anti-crossing points and c values of the Zak phases are measured along the yellow line in a with Ω1=125 MHz and Ω1/Ω2=5.3, compared with their theoretical values (dashed lines). The plots contain 200 data sets. d, f The WSLs and the absorption/reflection spectra with θ=0.3π, θ+=0.7π, Δc=108 MHz. e, g The WSLs and the absorption/reflection spectra with θ=0.4π, θ+=0.6π, Δc=45 MHz

We also show the reflection spectra in Fig. 2c, d, which is the directional emission from the state b~1 along −x direction. The reflection spectra also have features characterizing the anti-crossing of the WSLs (e.g., the peaks of reflection spectra correspond to band centers and anti-crossings). On the other hand, they can also be used to study the lattice transport between sites b~1b~1 for lattices in different topological phases, which is out of the scope of the current paper.

Zak phase measurement

In order to measure the Zak phases, we shall quantify the common features in the absorption spectra of lattices with the same Zak phase. In Fig. 3a, we maintain Ω1=Ω2=120 MHz and decrease Δc from bottom to top. The Zak phase is the same but the coupling between WSLs increases to widen the anti-crossing gap. We locate the anti-crossing points with its normalized energy:

Rn,m=(Δn,mϵ)/(ϵ+ϵ) 6

Fig. 3. Zak phase reconstruction from the absorption spectra.

Fig. 3

The experimental data of the absorption spectra for a the Semenoff insulator superradiance lattices with Ω1,2=120 MHz and Δc=174, 232, and 298 MHz from top to bottom, and b the SSH superradiance lattices with Δc=0, Ω1=118 MHz, and Ω1/Ω2=6.52, 4.25, and 2.15 from top to bottom. We use the marked extrema ϵ± (Δn,m) in absorption peaks and dips to locate the band centers (anti-crossings). The measured Rn,m (points) compared with the normalized ratio of WCs (dashed lines) for c the Semenoff insulator and d the SSH superradiance lattices, from which we reconstruct the Zak phases θ± in e and f. Error bars are obtained from four independent data sets (see WSLs and more absorption spectra in Supplementary Note 7)

On the other hand, according to the geometry of WSLs in the Δpδ diagram, Rn,m is approximately the normalized ratio of WCs between the WSLs:

Rn,mrnrnr+m=2nπ+θ2nπ+θ2mπ+θ+ 7

In Fig. 3c, we obtain R0,21/5, R0,11/3. We solve two equations of θ± from the values of R0,1 and R0,2 and conclude that θπ and θ+0, as shown in Fig. 3e.

For the SSH models, we keep Δc=0 and tune the Rabi frequencies of the two coupling fields from almost dimerization to the topological phase transition point. The Zak phases are maintained the same while the anti-crossing energy gaps increase from top to bottom in Fig. 3b. The measured Rn,m in Fig. 3d agree well with their expected values and the reconstructed Zak phases are obtained, θ±0.5π, as shown in Fig. 3f.

For a general RM Hamiltonian, i.e., Hs with Ω1Ω2 and Δc0, the Zak phases are neither integers nor half-integers. Along the yellow line in the phase diagram in Fig. 4a, we measure the absorption spectra to obtain R0,1 and R1,0 for each coupling field detuning Δc, as shown in Fig. 4b, and accordingly reconstruct the Zak phases in Fig. 4c, in comparison with the theoretical prediction as indicated by the dashed lines. As an example, we show the WSLs with θ=0.3π, θ+=0.7π in Fig. 4d and with θ=0.4π, θ+=0.6π in Fig. 4e. The corresponding absorption and reflection spectra are plotted in Fig. 4f, g, respectively.

Discussions

We realize the spectroscopic reconstruction of the Zak phases of momentum-space superradiance lattices. Without trapping atoms or controlling their velocities810, we take advantage of the atomic thermal motion39,40 to extract geometric phases from the anti-crossings of the WSLs. Therefore, our result pushes forward the room-temperature quantum simulation of topological matter. Meanwhile, it also paves a way for application of topological physics in optical devices that operate at ambient temperature.

The deviation between the reconstructed Zak phases and the theoretical prediction can be attributed to the following two reasons. First, when the band gap is small, the strong coupling between WSLs leads to a wide anti-crossing energy gap, such that the energy dip does not accurately reflect the location of the anti-crossing (see Δc0 for the Semenoff insulator in Fig. 3e). Second, the competition between Hf and Hs induces a systematic error even for small couplings between WSLs. For the SSH model in Fig. 3f, the force required for the two major anti-crossings is δ±(ϵ+ϵ_)/2. The induced potential energy between neighboring sites is comparable to the hopping strengths between lattice sites, such that Hf cannot be treated as a perturbation and the slope of WSLs E±n/δdeviates from r±n (see Supplementary Note 3). The consequence is that the energy of the anti-crossing Δ0,1 (Δ1,0) is always lower (higher) than the degeneracy point predicted by the linear approximation in Eq. (5) (see the difference between spectra extrema Δn,m and crossing points of blue dotted lines in Figs. 2a, b and 4d, e), leading to systematic errors in determining the Zak phases.

To improve the accuracy in extracting the spectroscopic features of Zak phases, we are developing a spectral hole burning technique to map out the two-dimensional velocity-dependent absorption spectra of WSLs, as shown in upper panels of Fig. 2a. By using a narrow linewidth saturation field that couples the ground state to an ancillary state, we can selectively bleach the ground state population of atoms with a certain velocity. By comparing the bleached and unbleached absorption spectra, the contribution from atoms with that velocity is obtained.

Our scheme can be generalized to measure multipole moments of higher-order topological insulators33,34 by detecting the slopes of WSLs18, and to measure Chern numbers by counting the Zak phase winding. In the current framework, two or higher dimensional superradiance lattices16 can be synthesized by introducing more coupling fields4144, following similar methods on dimension extension in photonic lattices45 and synthetic dimensions4648. We can use three coupling fields to form a 2D interference pattern in xy plane (Fig. 5a), illustrating the BZ of a momentum-space honeycomb superradiance lattices41. We can identify the Chern number C± of the upper (+) and lower (−) bands from the winding number of the 1D Zak phase along the perpendicular dimension16,2432:

C±=12π0Ldyθ±(y)y 8

where θ±(y) is the Zak phase along the x-axis cut of the 2D BZ with a fixed y and L is the length of real-space BZ along the y-axis. In Fig. 5b, we schematically show that θ±(y) can be measured from the absorption spectra of a probe field with a beam size much smaller than L. In order to suppress the paraxial diffraction, we need to ensure L is much larger than the wavelength by minimizing the angle between the two copropagating coupling fields. After collecting the θ±(y), we determine the Chern number of the 2D superradiance lattices by counting how many times it winds within L, as shown in Fig. 5c.

Fig. 5. Proposal of measuring Chern numbers from Zak phases.

Fig. 5

a 2D interference pattern in xy plane induced by three coupling fields, illustrating the Brillouin zone of a superradiance lattice Haldane model. b Measurement of the position dependent Zak phases. For each fixed position in y-axis, the θ±(y) along the x direction is obtained from the absorption spectrum of the corresponding probe field. c The winding number of θ±(y) from y=0 to L is related to the Chern number C, where L is the BZ width projected on y-axis

With the ability of measuring geometric phases in SLs, a promising direction in the next stage is to introduce interactions between atoms, e.g., by using Rydberg states49. It is particularly interesting to notice that the short-range interaction in real space has long-range effect in momentum-space SLs, which is difficult to realize in real-space lattices50.

In conclusion, we develop a method of reconstructing Zak phases from the anti-crossings of WSLs by measuring the Doppler-broadened absorption spectra of room-temperature superradiance lattices. This method can be directly generalized to measure Chern numbers by counting the Zak phase windings in 2D lattices16 with more coupling fields4144. Our method can also be implemented in cold atoms24 by controlling the atomic velocity to track the peak shifting, as sketched in Fig. 1c. We can also use a hole burning technique to obtain the WSL of atoms with different velocities. Our results pave the way to detect multipole moments in higher-order topological insulators18,33,34.

Materials and methods

Effective Hamiltonian

Here we derive the effective Hamiltonian H=Hs+Hf+Hp. It is convenient to work with the master equation in the inertial reference frame of the moving atoms. The Rabi frequencies of the two plane wave components of the coupling field are Ω1 (x directional) and Ω2 (x directional), and the corresponding Doppler shifted frequencies are vcδ and vc+δ, respectively. For the probe field propagating along x direction, the Rabi frequency and Doppler shifted frequency are Ωp and vpδ, respectively. The original Hamiltonian of an ensemble of Λ-type three-level atoms in EIT is written as:

H=ωbgbb+ωagaa+Ω1eivcδt+ikcxba+Ω2eivc+δtikcxba+H.c.+Ωpeivpδt+ikpxbg+H.c. 9

In order to eliminate the dynamic phase factors, we transform the Hamiltonian into the interaction picture:

V=U1HUS=Δcaa+ΩpeiΔpδt+ikpxbg+H.c.+Ω1eiδt+ikcxba+Ω2eiδtikcxba+H.c. 10

where U=expiSt and S=ωbgbb+ωbgvcaa.

In the experiments, the powers of the coupling fields are much larger than that of the probe field, and the probe field is far below the saturation strength, i.e., Ω1,2,ΓΩp. Therefore, we only keep the first order of Ωp but keep all orders of Ω1,2. The relevant dynamical equations that govern the evolution of coherence ρag and ρbg are:

ρ°ag=iΩ1eiδtikcx+Ω2eiδt+ikcxρbg+Δcρagγaρag,ρ°bg=iΩ1eiδt+ikcx+Ω2eiδtikcxρag+ΩpeiΔpδt+ikpxγbgρbg 11

where γbg=Γ/2+γb is the decoherence rate of ρbg and γi is the dephasing rate of the level i. The general solutions are assumed as:

ρag=jρag2jeik2jx,ρbg=jρbg2j+1eik2j+1x 12

where kj=kp+j1kcjk_c, x=x+δ/kct=x+vt being the position of atoms in motion. In the weak excitation limit ρag[2j],ρbg[2j+1]1, the wavefunction of a single atom at xm is approximately Ψmρagam+ρbgbm+gm. Therefore, the wavefunction of the whole atomic ensemble reads:

Ψ=mΨmmjρag2jeik2jxmg1g2amgN+mjρbg2j+1eik2j+1xmg1g2bmgN+G=jα2ja^2j+β2j+1b^2j+1+1G 13

where we use the definition of the collective ground state G=g1g2gN and the probability amplitudes of the timed Dicke states α2j=Nρag[2j], β2j+1=Nρbg2j+1. Combining Eqs. (11)–(13), we write the dynamic equation formally:

iddtΨ=Hiγ^Ψ 14

with the effective Hamiltonian:

H=jΔc+2jδa^2ja^2j+δ2j+1b^2j+1b^2j+1+jΩ1a^2jb^2j+1+Ω2a^2jb^2j1+NΩpeiΔptb^1+H.c.

and the dissipation operator:

γ^=jγaa^2ja^2j+γbgb^2j+1b^2j+1

Absorption and PDOS

The induced polarization of the state Ψ in Eq. (13) is defined as:

P=ΨerΨ=jmμρbg2j+1eik2j+1xm+c.c 15

where μ=gerb is the single atom dipole moment. Since the atoms are homogeneously distributed, the polarization density as a function of position reads:

Px=njμρbg2j+1eik2j+1x+c.c 16

where n is the atomic density. Therefore, the susceptibility of the atoms is:

χ=Pxϵ0Epeikpx=jχ[j]eik2jx 17

The optical absorption coefficient A is related to the 0th-order component (linear) of the susceptibility in Eq. (17), and further connected to PDOS of the superradiance lattices:

AImχ0Imβ1Imb~11Δp+iγbgHfvHsb~1=Imlψlvb~12Δp+iγbgElvπlδDΔpElvψlvb~12 18

where we take advantage of the Green’s function approach5153 and the last step is valid when the decoherence rate tends to zero compared with the band width. Here Hf, ψl, and El are functions of v. The power of the transmitted probe field is Pt=PieAs, where s is the length of the Rb vapor cell.

Supplementary information

41377_2022_990_MOESM1_ESM.pdf (7.4MB, pdf)

Supplementary Notes: Measuring Zak phase in room-temperature atoms

Acknowledgements

The authors thank Luqi Yuan for useful discussions. This work was supported by the National Natural Science Foundation of China (Grants Nos. 11874322, 11934011 and U21A20437), the National Key Research and Development Program of China (Grants Nos. 2019YFA0308100, 2018YFA0307200 and 2017YFA0304202), Zhejiang Province Key Research and Development Program (Grant No. 2020C01019), the Strategic Priority Research Program of Chinese Academy of Sciences (Grant No. XDB28000000), the Fundamental Research Funds for the Central Universities and Innovation Program for Quantum Science and Technology (Grant No. 2021ZD0303200). We gratefully acknowledge HZWTECH for providing computation facilities.

Author contributions

H.C. and D.W.W. conceived the project. H.C. and X.X. designed the experiment. J.W., X.X., C.X., and G.Q. built the experimental setup and carried out the measurement. R.M. and H.C. did numerical simulation. R.M., J.W., X.X. and H.C. performed data analysis. D.W.W. and S.Y.Z. supervised the research. H.C. and D.W.W. wrote the manuscript with comments and contributions from all authors.

Conflict of interest

The authors declare no competing interests.

Footnotes

These authors contributed equally: Ruosong Mao, Xingqi Xu, Jiefei Wang

Contributor Information

Han Cai, Email: hancai@zju.edu.cn.

Da-Wei Wang, Email: dwwang@zju.edu.cn.

Supplementary information

The online version contains supplementary material available at 10.1038/s41377-022-00990-7.

References

  • 1.Hasan MZ, Kane CL. Colloquium: topological insulators. Rev. Mod. Phys. 2010;82:263. doi: 10.1103/RevModPhys.82.3045. [DOI] [Google Scholar]
  • 2.Qi XL, Zhang SC. Topological insulators and superconductors. Rev. Mod. Phys. 2011;83:1057. doi: 10.1103/RevModPhys.83.1057. [DOI] [Google Scholar]
  • 3.Zak J. Berry’s phase for energy bands in solids. Phys. Rev. Lett. 1989;62:2747. doi: 10.1103/PhysRevLett.62.2747. [DOI] [PubMed] [Google Scholar]
  • 4.Von Klitzing K, Dorda G, Pepper M. New method for high-accuracy determination of the fine-structure constant based on quantized Hall resistance. Phys. Rev. Lett. 1980;45:494. doi: 10.1103/PhysRevLett.45.494. [DOI] [Google Scholar]
  • 5.Thouless DJ, Kohmoto M, Nightingale MP, den Nijs M. Quantized Hall conductance in a two-dimensional periodic potential. Phys. Rev. Lett. 1982;49:405. doi: 10.1103/PhysRevLett.49.405. [DOI] [Google Scholar]
  • 6.Haldane FDM. Model for a quantum Hall effect without landau levels: condensed-matter realization of the “parity anomaly”. Phys. Rev. Lett. 1988;61:2015. doi: 10.1103/PhysRevLett.61.2015. [DOI] [PubMed] [Google Scholar]
  • 7.Cooper NR, Dalibard J, Spielman IB. Topological bands for ultracold atoms. Rev. Mod. Phys. 2019;91:015005. doi: 10.1103/RevModPhys.91.015005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Atala M, et al. Direct measurement of the Zak phase in topological Bloch bands. Nat. Phys. 2013;9:795. doi: 10.1038/nphys2790. [DOI] [Google Scholar]
  • 9.Meier EJ, et al. Observation of the topological Anderson insulator in disordered atomic wires. Science. 2018;362:929. doi: 10.1126/science.aat3406. [DOI] [PubMed] [Google Scholar]
  • 10.Xie D, Gou W, Xiao T, Gadway B, Yan B. Topological characterizations of an extended Su-Schrieffer-Heeger model. npj Quantum Inf. 2019;5:1. doi: 10.1038/s41534-019-0159-6. [DOI] [Google Scholar]
  • 11.Hauke P, Lewenstein M, Eckardt A. Tomography of band insulators from quench dynamics. Phys. Rev. Lett. 2014;113:045303. doi: 10.1103/PhysRevLett.113.045303. [DOI] [PubMed] [Google Scholar]
  • 12.Zache TV, et al. Dynamical topological transition in the massive Schwinger model with a θ term. Phys. Rev. Lett. 2019;122:050403. doi: 10.1103/PhysRevLett.122.050403. [DOI] [PubMed] [Google Scholar]
  • 13.Aidelsburger M, et al. Measuring the Chern number of Hofstadter bands with ultracold bosonic atoms. Nat. Phys. 2015;11:162. doi: 10.1038/nphys3171. [DOI] [Google Scholar]
  • 14.Chalopin T, et al. Probing chiral edge dynamics and bulk topology of a synthetic Hall system. Nat. Phys. 2020;16:1017. doi: 10.1038/s41567-020-0942-5. [DOI] [Google Scholar]
  • 15.Maksimov DN, Bulgakov EN, Kolovsky AR. Wannier-Stark states in double-periodic lattices. I. One-dimensional lattices. Phys. Rev. A. 2015;91:053631. doi: 10.1103/PhysRevA.91.053631. [DOI] [Google Scholar]
  • 16.Lee WR, Park K. Direct manifestation of topological order in the winding number of the Wannier-Stark ladder. Phys. Rev. B. 2015;92:195144. doi: 10.1103/PhysRevB.92.195144. [DOI] [Google Scholar]
  • 17.Kolovsky AR. Topological phase transitions in tilted optical lattices. Phys. Rev. A. 2018;98:013603. doi: 10.1103/PhysRevA.98.013603. [DOI] [Google Scholar]
  • 18.Poddubny AN. Distinguishing trivial and topological quadrupolar insulators by Wannier-Stark ladders. Phys. Rev. B. 2019;100:075418. doi: 10.1103/PhysRevB.100.075418. [DOI] [Google Scholar]
  • 19.Marzari N, Mostofi AA, Yates JR, Souza I, Vanderbilt D. Maximally localized Wannier functions: theory and applications. Rev. Mod. Phys. 2012;84:1419. doi: 10.1103/RevModPhys.84.1419. [DOI] [Google Scholar]
  • 20.King-Smith RD, Vanderbilt D. Thoery of polarization of crystalline solids. Phys. Rev. B. 1993;47:1651R. doi: 10.1103/PhysRevB.47.1651. [DOI] [PubMed] [Google Scholar]
  • 21.Resta R. Macroscopic polarization in crystalline dielectrics: the geometric phase approach. Rev. Mod. Phys. 1994;66:899. doi: 10.1103/RevModPhys.66.899. [DOI] [Google Scholar]
  • 22.Kivelson S. Wannier functions in one-dimensional disordered systems: application to fractionally charged solitons. Phys. Rev. B. 1982;26:4269. doi: 10.1103/PhysRevB.26.4269. [DOI] [Google Scholar]
  • 23.Wang DW, Liu RB, Zhu SY, Scully MO. Superradiance lattice. Phys. Rev. Lett. 2015;114:043602. doi: 10.1103/PhysRevLett.114.043602. [DOI] [PubMed] [Google Scholar]
  • 24.Chen L, et al. Experimental observation of one-dimensional superradiance lattices in ultracold atoms. Phys. Rev. Lett. 2018;120:193601. doi: 10.1103/PhysRevLett.120.193601. [DOI] [PubMed] [Google Scholar]
  • 25.Mi C, et al. Time-resolved interplay between superradiant and subradiant states in superradiance lattices of Bose-Einstein condensates. Phys. Rev. A. 2021;104:043326. doi: 10.1103/PhysRevA.104.043326. [DOI] [Google Scholar]
  • 26.Scully MO, Fry ES, Ooi CH, Wódkiewicz K. Directed spontaneous emission from an extended ensemble of N atoms: timing is everything. Phys. Rev. Lett. 2006;96:010501. doi: 10.1103/PhysRevLett.96.010501. [DOI] [PubMed] [Google Scholar]
  • 27.He Y, et al. Geometric control of collective spontaneous emission. Phys. Rev. Lett. 2020;125:213602. doi: 10.1103/PhysRevLett.125.213602. [DOI] [PubMed] [Google Scholar]
  • 28.Su WP, Schrieffer JR, Heeger AJ. Solitons in polyacetylene. Phys. Rev. Lett. 1979;42:1698. doi: 10.1103/PhysRevLett.42.1698. [DOI] [Google Scholar]
  • 29.Meier EJ, An FA, Gadway B. Observation of the topological soliton state in the Su–Schrieffer–Heeger model. Nat. Commun. 2016;7:13986. doi: 10.1038/ncomms13986. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Coh S, Vanderbilt D. Electric polarization in a Chern insulator. Phys. Rev. Lett. 2009;102:107603. doi: 10.1103/PhysRevLett.102.107603. [DOI] [PubMed] [Google Scholar]
  • 31.Yu R, Qi XL, Bernevig AB, Fang Z, Dai X. Equivalent expression of Z2 topological invariant for band insulators using the non-Abelian Berry connection. Phys. Rev. B. 2011;84:075119. doi: 10.1103/PhysRevB.84.075119. [DOI] [Google Scholar]
  • 32.Taherinejad M, Garrity KF, Vanderbilt D. Wannier center sheets in topological insulators. Phys. Rev. B. 2014;89:115102. doi: 10.1103/PhysRevB.89.115102. [DOI] [Google Scholar]
  • 33.Benalcazar WA, Bernevig BA, Hughes TL. Quantized electric multipole insulators. Science. 2017;357:61. doi: 10.1126/science.aah6442. [DOI] [PubMed] [Google Scholar]
  • 34.Ezawa M. Higher-order topological insulators and semimetals on the breathing kagome and pyrochlore lattices. Phys. Rev. Lett. 2018;120:026801. doi: 10.1103/PhysRevLett.120.026801. [DOI] [PubMed] [Google Scholar]
  • 35.Wang J, Zhu Y, Jiang KJ, Zhan MS. Bichromatic electromagnetically induced transparency in cold rubidium atoms. Phys. Rev. A. 2003;68:195144. doi: 10.1103/PhysRevA.68.063810. [DOI] [Google Scholar]
  • 36.Glutsch S, Bechstedt F. Interaction of Wannier-Stark ladders and electrical breakdown in superlattices. Phys. Rev. B. 1999;60:16584. doi: 10.1103/PhysRevB.60.16584. [DOI] [Google Scholar]
  • 37.Koochaki Kelardeh H, Apalkov V, Stockman MI. Wannier-Stark states of graphene in strong electric field. Phys. Rev. B. 2014;90:085313. doi: 10.1103/PhysRevB.90.085313. [DOI] [Google Scholar]
  • 38.Kuang S, Wan R, Du P, Jiang Y, Gao J. Transmission and reflection of electromagnetically induced absorption grating in homogeneous atomic media. Opt. Express. 2008;16:15455. doi: 10.1364/OE.16.015455. [DOI] [PubMed] [Google Scholar]
  • 39.Peng P, et al. Anti-parity-time symmetry with flying atoms. Nat. Phys. 2016;12:1139. doi: 10.1038/nphys3842. [DOI] [Google Scholar]
  • 40.Zhang S, et al. Thermal-motion-induced non-reciprocal quantum optical system. Nat. Photon. 2018;12:744. doi: 10.1038/s41566-018-0269-2. [DOI] [Google Scholar]
  • 41.Wang DW, Cai H, Yuan L, Zhu SY, Liu RB. Topological phase transitions in superradiance lattices. Optica. 2015;2:712. doi: 10.1364/OPTICA.2.000712. [DOI] [Google Scholar]
  • 42.Zhang Z, et al. Particlelike behavior of topological defects in linear wave packets in photonic graphene. Phys. Rev. Lett. 2019;122:233905. doi: 10.1103/PhysRevLett.122.233905. [DOI] [PubMed] [Google Scholar]
  • 43.Yuan J, Wu C, Wang L, Chen G, Jia S. Observation of diffraction pattern in two-dimensional optically induced atomic lattice. Opt. Lett. 2019;44:4123. doi: 10.1364/OL.44.004123. [DOI] [PubMed] [Google Scholar]
  • 44.Yuan J, et al. Tunable optical vortex array in a two-dimensional electromagnetically induced atomic lattice. Opt. Lett. 2021;46:4184. doi: 10.1364/OL.432036. [DOI] [PubMed] [Google Scholar]
  • 45.Ozawa T, et al. Topological photonics. Rev. Mod. Phys. 2019;91:015006. doi: 10.1103/RevModPhys.91.015006. [DOI] [Google Scholar]
  • 46.Yuan L, Lin Q, Xiao M, Fan S. Synthetic dimension in photonics. Optica. 2018;5:1396. doi: 10.1364/OPTICA.5.001396. [DOI] [Google Scholar]
  • 47.Yu D, Peng B, Chen X, Liu XJ, Yuan L. Topological holographic quench dynamics in a synthetic frequency dimension. Light Sci. Appl. 2021;10:209. doi: 10.1038/s41377-021-00646-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Dutt A, et al. A single photonic cavity with two independent physical synthetic dimensions. Science. 2020;367:59. doi: 10.1126/science.aaz3071. [DOI] [PubMed] [Google Scholar]
  • 49.Li Y, et al. Many-body chiral edge currents and sliding phases of atomic spin waves in momentum-space lattice. Phys. Rev. Lett. 2020;124:140401. doi: 10.1103/PhysRevLett.124.140401. [DOI] [PubMed] [Google Scholar]
  • 50.An FA, Meier EJ, Ang’ong’a J, Gadway B. Correlated dynamcis in a synthetic lattice of momentum states. Phys. Rev. Lett. 2018;120:040407. doi: 10.1103/PhysRevLett.120.040407. [DOI] [PubMed] [Google Scholar]
  • 51.Ozawa T, Carusotto I. Anomalous and quantum Hall effects in lossy photonic lattices. Phys. Rev. Lett. 2014;112:133902. doi: 10.1103/PhysRevLett.112.133902. [DOI] [PubMed] [Google Scholar]
  • 52.Cai H, et al. Experimental observation of momentum-space chiral edge currents in room-temperature atoms. Phys. Rev. Lett. 2019;122:023601. doi: 10.1103/PhysRevLett.122.023601. [DOI] [PubMed] [Google Scholar]
  • 53.He Y, et al. Flat-band localization in Creutz superradiance lattices. Phys. Rev. Lett. 2021;126:103601. doi: 10.1103/PhysRevLett.126.103601. [DOI] [PubMed] [Google Scholar]
  • 54.Harb M, et al. Quantum transport modelling of silicon nanobeams using heterogeneous computing scheme. J. Appl. Phys. 2016;119:124304. doi: 10.1063/1.4944649. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

41377_2022_990_MOESM1_ESM.pdf (7.4MB, pdf)

Supplementary Notes: Measuring Zak phase in room-temperature atoms


Articles from Light, Science & Applications are provided here courtesy of Nature Publishing Group

RESOURCES