Open Access
21 June 2022 Observation of flat-band and band transition in the synthetic space
Author Affiliations +
Abstract

Constructions of synthetic lattices in modulated ring resonators attract growing attention to interesting physics beyond the geometric dimensionality, where complicated connectivities between resonant frequency modes are explored in many theoretical proposals. We implement experimental demonstration of generating a stub lattice along the frequency axis of light, in two coupled ring resonators of different lengths, with the longer one dynamically modulated. Such a synthetic photonic structure intrinsically exhibits the physics of flat band. We show that the time-resolved band structure read-out from the drop-port output of the excited ring is the intensity projection of the band structure onto a specific resonant mode in the synthetic momentum space, where gapped flat band, mode localization effect, and flat-to-nonflat band transition are observed in experiments and verified by simulations. This work provides evidence for constructing a synthetic stub lattice using two different rings, which, hence, makes a solid step toward experimentally constructing complicated lattices in multiple rings associated with synthetic frequency dimensions.

1.

Introduction

Synthetic dimensions in photonics have attracted broad interest in recent years.14 They promise new ways to study fundamental physical phenomena with exotic artificial connectivities59 and manipulate light in various ways,1018 pointing toward exploration of higher-dimensional physics, beyond three dimensions.1921 Among recent experimental achievements, different degrees of freedom of light, including arrival times of pulses,10,11 frequencies,2224 and modal dimensions,8 have been used to construct synthetic dimensions. Hence, a variety of novel physics have been demonstrated in synthetic dimensions, such as the photonic topological insulator,8 the Hall ladder with effective magnetic flux,25 the trajectory of dynamic band structures,26 and the topological funneling with non-Hermitian physics,27 physical models of which are hard to build in structures with only spatial dimensions.

Among these platforms, the dynamically modulated ring resonator system has manifested as a powerful platform where resonant modes with equally spaced frequencies are coupled by external modulation, and then the synthetic frequency dimension is created.22,23 The modulation applied by external voltages provides the unique advantages of breaking the constraint of fixed geometric structures after fabrication and thus provides an important possibility of achieving complicated functionalities with great experimental flexibility and reconfigurability.4 Experimental implementations have been performed in ring resonator systems including fiber loops or on-chip microrings,2426,2831 and physical phenomena including band structures measurements,24,26 spectral Bloch oscillations,29 and non-Hermitian topology30,32 have been shown, where one ring or two identical rings have been used. On the other hand, theoretical proposals have been explored where formations of photonic lattice with different lengths of rings can enable intriguing studies of rich physics, such as the Haldane model,7 topological quench dynamics,33 two-dimensional Lieb lattice,34 three-dimensional topological insulator,35 and high-order topologies.36,37 However, to realize these theoretical proposals requires constructing complicated lattice structures beyond simple lines or square geometries in synthetic space in two or more rings of different lengths. Therefore, as a crucial step further, one desires to first prove the capability of creating a complex lattice in two coupled rings of different lengths in the experiment.

In this work, we experimentally couple two rings of different lengths, where one ring undergoes the dynamic modulation, and construct a photonic stub lattice (also called one-dimensional Lieb lattice)3842 associated with synthetic frequency dimension. Such a configuration is not straightforward to construct, compared with lattices in one ring24,26 or in two identical rings.32 One intrinsic feature of the stub lattice is the natural existence of the flat (dispersionless) band.4349 In our experiments, the time-resolved energy bands from the drop-port output of the excited ring are obtained, corresponding to the projection of the band structures of the stub lattice, which, however, is on the synthetic dimension. Moreover, by exciting the resonant modes through the selected input port of one ring and recording the output transmission from the same ring, we observe the effective localization of the resonant modes near the flat band. Such a flat band in synthetic space can further be modified by adding the long-range couplings in the modulation, which leads to the transition from the flat to nonflat bands. By combining theoretical analysis, we show that coupling two rings at different lengths leads to experimental observations where bands are projected onto superposition modes, which is very different from previous works on flat-band physics. Such a feature is unique in the platform of synthetic dimensions with modulated ring resonators, so our work, therefore, exhibits a crucial step toward constructing more complicated synthetic lattices in multiple rings of different lengths.

2.

Materials and Methods

We start by illustrating the model of two modulated rings of different lengths, labeled as A and B in Fig. 1(a). In the absence of group velocity dispersion, the ring resonator supports a set of modes with equally spaced frequencies. If we set the central resonant frequency at ω0, the n’th mode in the ring A(B) has the frequency ωA(B),n=ω0+nΩA(B), where ΩA(B)=2πvg/LA(B) is the free spectral range (FSR) of ring A(B), and vg is the group velocity. We consider the length of ring A (LA) twice as long as the length of ring B (LB), i.e., LA=2LB, which gives 2ΩA=ΩBΩ. The electro-optic modulator (EOM) is placed inside ring A with the modulation frequency ΩM=Ω/2, the modulation strength g, and the modulation phase ϕ, which provides the connectivity between adjacent resonant modes in ring A, while there is no modulator in ring B, so resonant modes in ring B remain unconnected. Two resonant modes in two rings at the same frequency can be coupled through a fiber coupler with the coupling strength κ [see Fig. 1(b)]. Therefore, three types of modes exist in the system, defined as An, Bn, and Cn, where An and Cn are the resonant modes at frequencies ωn and ωn+Ω/2 in ring A, and Bn is the resonant mode at frequency ωn in ring B with ωn=ω0+nΩ. In particular, modes An and Bn are coupled for the same n, while An is coupled to Cn1 and Cn through the modulation under the lowest-order approximation, resulting in the synthetic lattice shown in Fig. 1(b).

Fig. 1

Configuration of a synthetic photonic stub lattice. (a) Two coupled ring resonators, where the FSR of ring A is half of the FSR of ring B, i.e., 2ΩA=ΩBΩ. Ring A undergoes dynamic modulation by placing an EOM with the modulation frequency ΩM=Ω/2. Waveguides are connected to rings for input/output signals. (b) The system in (a) can be mapped into a photonic stub lattice along the synthetic frequency dimension (f), with An, Bn, and Cn indicating three types of lattice sites. (c) The corresponding band structures of the synthetic stub lattice in (b) with g=κ and ϕ=0.5π.

AP_4_3_036002_f001.png

The corresponding tight-binding Hamiltonian of the system is34

Eq. (1)

H=n[ωn(anan+bnbn)+(ωn+Ω/2)cncn]+n[κanbn+2gcos(Ωt/2+ϕ)(ancn+ancn1)+h.c.],
where an, bn, and cn (an, bn, and cn) are the creation (annihilation) operators for the modes An, Bn, and Cn, respectively. Equation (1) can be simplified into the interaction picture by taking the rotating-wave approximation,50 which results in

Eq. (2)

Hc=n[κanbn+g(ancneiϕ+ancn1eiϕ)+h.c.].
Equation (2) describes the Hamiltonian of a synthetic lattice structure, which is analog to the spatial stub lattice,3842 but it is along the frequency axis of light.

To understand the underlying physics of the Hamiltonian described in Eq. (2), we can rewrite Eq. (2) into the kf space as follows:

Eq. (3)

Hk=kf[κakfbkf+2gcos(kfΩ/2+ϕ)akfckf+h.c.],
where kf is the wave vector reciprocal to the frequency dimension acting as a time variable.4 The corresponding photonic band structure of the system is then given as

Eq. (4)

εkf,0=0,εkf,±=±[2gcos(kfΩ/2+ϕ)]2+κ2,
where εkf,j(j=0,±) are the eigenvalues from Eq. (3), corresponding to three bands plotted in Fig. 1(c) within the first Brillouin zone with kf[0,2π/Ω]. One can see a flat band εkf,0 in the middle gapped from the upper and lower dispersive bands εkf,±, which indicates that light can be efficiently localized in the flat band without scattering.4447 Let ψkf,j=(ψkf,jA,ψkf,jB,ψkf,jC)T be the eigenstates corresponding to εkf,j, with ψkf,jA, ψkf,jB, and ψkf,jC being the projection of the eigenstates on the three modes (Ak, Bk, and Ck) in the kf space, and then we have

Eq. (5)

ψkf,0=(0,G,κ)T/G2+κ2,ψkf,±=(±G2+κ2,κ,G)T/2(κ2+G2),
with G=2gcos(kfΩ/2+ϕ). One notes that the flat band (j=0) has no projection onto mode Ak due to ψkf,0A=0, while the two dispersive bands (j=±) are antisymmetrically projected onto mode Ak but symmetrically projected onto modes Bk and Ck.

To implement the idea of the synthetic photonic stub lattice described in Eq. (2) for the potential experimental demonstration, we continue with considering a realistic model of two ring resonators coupled with input and output waveguides. In the following, we consider two excitation cases by selectively choosing the input/output ports, which are referred to as BinBout and AinAout, as in Fig. 1(a). First, we inject the field into the system through the input port of ring B and measure the drop-port output of ring B as well (BinBout), in which only frequency mode Bn in ring B is directly excited. The normalized drop-port transmission ToutB can be expressed in the kf space as [see Eqs. (S10)–(S14) in the Supplementary Material]

Eq. (6)

ToutB(t=kf;Δω)=γB2|ψkf,jB|4(Δωεkf,j)2+γ2,
where γB is the coupling strength between ring B and waveguides, γ is the total loss, and Δω is the frequency detuning. εkf,j and ψkf,jB are determined by Eqs. (4) and (5), and j refers to the term having corresponding energy closest to the input frequency. Previous works have demonstrated that the photonic band structure can be measured by time-resolved transmission spectroscopy, where the drop-port output transmission signal is obtained by scanning the frequency of the input laser linearly with time.24,26 Therefore, Eq. (6) indicates that the band structures read out from the drop-port output of ring B exhibit the projection of the band structure on mode Bk in kf space.

On the other hand, for the case of AinAout, by changing the input/output port to ring A, similar input/output coupled amplitude equations can be obtained. The corresponding normalized drop-port transmissions in the kf space are [see Eqs. (S1)–(S9) in the Supplementary Material]

Eq. (7)

ToutA(t=kf;Δω)=γA2|ψkf,jA|2|ψkf,jA+ψkf,jC|2(Δωεkf,j)2+γ2,

Eq. (8)

ToutA(t=kf;Δω+Ω/2)=γA2|ψkf,jC|2|ψkf,jA+ψkf,jC|2(Δωεkf,j)2+γ2,
where γA is the waveguide resonator coupling strength of ring A. Equations (7) and (8) refer to the situation of an input field near resonance with the reference frequencies ω0 and ω0+Ω/2, respectively. This means that the band structure resolved from the drop-port transmission through ring A is the projection of the band structure on the superposition modes of Ak and Ck separated by Ω/2 along the frequency dimension.

In experiments, we use two fiber ring resonators coupled together through a 2×2 fiber coupler with coupler ratio 70:30, as shown in Fig. S1 in the Supplementary Material. The two rings are excited separately by selectively choosing ring A or B as the input port of the laser source (Ain or Bin), while the transmission is recorded from the corresponding drop port (Aout or Bout). After calibration, the lengths of the two rings are LA=20.4  m and LB=10.2  m, corresponding to ΩA=2π·10  MHz and ΩB=2π·20  MHz. To form the synthetic stub lattice described in Fig. 1(b), we drive the EOM in ring A by a sinusoidal radio frequency (RF) signal in the form of VMcos(ΩMt+ϕ) with ΩM=2π·10  MHz, and ϕ=0.5π.

3.

Results

To demonstrate the construction of the synthetic photonic stub lattice in the experiment, we perform the band structure measurements by finely sweeping the frequency of the input laser through multiple free-spectral ranges.26 We first inject the laser source into the input port of ring B and measure the output transmission spectra from the drop port of ring B (BinBout). Figures 2(a3)2(d3) plot the measured output transmission signals, while each transmission spectrum contains multiple sinusoidal signals, as enlarged in Fig. S2(d) in the Supplementary Material. By breaking the transmission signals into time slices with the time window equaling one roundtrip time of ring B (2π/Ω), i.e., the periodicity of the synthetic stub lattice, one gets the time-resolved band structures, as shown in Figs. 2(a1)2(d1), with varied modulation amplitude VM. We calculate the intensity projections of the band structure on mode Bk using Eq. (6) and show the results in Figs. 2(a2)2(d2) with γ=0.07Ω, where the width of the bands results from the loss term γ added in the coupled mode equations [see Eq. (S10) in the Supplementary Material]. The vertical slice of the time-resolved band structure at a fixed time (kf) exhibits a Lorentz function for each of the three bands (j=0,±), with γ characterizing the width of each band in Eq. (6), as plotted in Fig. S2(c) in the Supplementary Material. Without modulation (VM=0), coupled rings result in two Lorentzian resonances of the unmodulated rings, which exhibit two resonances with separation 2κ due to energy splitting between two coupled resonant modes An and Bn [see Fig. 2(a3)]. It leads to two straight energy bands with constant intensity distributions in both experiment [see Fig. 2(a1)] and theory [see Fig. 2(a2)]. The feature of the synthetic stub lattice begins to manifest once the modulation is applied, as shown in Figs. 2(b1)2(d1), where one notices that three bands exist, and the intensity distributions vary with the modulation amplitude. For a small modulation amplitude [see Fig. 2(b1) with VM=1.5  V], the energy of the eigenstate mainly focuses on the upper and lower dispersive bands, which transfers to the middle flat band when the modulation strength becomes larger, as shown in Fig. 2(d1) with VM=2.5  V. The theoretical plots exhibit excellent agreement with experimental measurements, which clearly shows that the energy of the eigenstate flows from dispersive bands to the flat band when increasing g [see Figs. 2(b2)2(d2)]. Moreover, the intensity distributions on the two dispersive bands have a symmetric pattern within the period of kf[0,2π/Ω], which is consistent with the analytical solutions in Eq. (5).

Fig. 2

Band structure measurements for the case of BinBout. (a1)–(d1) Experimentally observed band structures with different modulation amplitudes VM. (a2)–(d2) Simulation results of the projected output intensity distribution of the band structure on mode Bk, based on Eqs. (4)–(6), where g takes different values with fixed κ=0.06Ω and ϕ=0.5π. (a3)–(d3) Measured transmission spectra from the drop port of ring B. The vertical axis represents the frequency detuning of the input laser source normalized to Ω, while the bottom horizontal axis in (a1)–(d2) represents one roundtrip time in ring B with the period of 2π/Ω.

AP_4_3_036002_f002.png

Fig. 3

Band structure measurements for the case of AinAout. (a1)–(c1) Experimentally observed band structures varied with VM. (a2)–(c2) Simulation results of the projected intensity distribution of the band structure on modes Ak and Ck, based on Eqs. (4), (5) and (7), (8), with κ=0.06Ω and ϕ=0.5π. (a3)–(c3) Transmission spectra measured from the drop port of ring A. The bottom horizontal axis in (a1)–(c2) represents one roundtrip time in ring A with the period of 4π/Ω.

AP_4_3_036002_f003.png

We then consider the case of AinAout by switching the input and output fibers to ring A. The output transmissions of modes Ak and Ck separated by Ω/2 are measured simultaneously, as shown in Figs. 3(a3)3(c3), where the scale of the vertical axes is twice as large as the scale in Fig. 2. Since in experiments the time window to break the measured output transmission signals of ring A equals one roundtrip time of ring A (4π/Ω), we measure a combination of intensity projections of the band structure on Ak and Ck, which gives kf[0,4π/Ω], as plotted in Figs. 3(a1)3(c1). Theoretical results from Eqs. (7) and (8) are plotted in Figs. 3(a2)3(c2) with γ=0.07Ω. When there is no modulation, one sees two nearby straight bands near Δω/Ω=0.25 due to energy splitting from coupling between modes An and Bn and one single straight band near Δω/Ω=0.25, referring to the resonance of Cn in both experiment and theory [see Figs. 3(a1) and 3(a2)]. Note that the mode splitting of the top two bands in Figs. 3(a1)3(a3) is the same as that in Figs. 2(a1)2(a3), which is also characterized by 2κ. When the modulation is applied, the band structures near Δω/Ω=±0.25 show different features. For upper bands near Δω/Ω=0.25, one sees two dispersive bands, corresponding to the band structure in Fig. 1(c) projected to modes Ak [see Figs. 3(b1) and 3(c1)], which matches well with the calculated results from Eq. (7) [see Figs. 3(b2) and 3(c2)]. On the other hand, for lower bands near Δω/Ω=0.25, one can clearly see three bands, with the middle one being flat. The intensity projections of two dispersive bands on mode Ck are relatively weak in both experiment and theory. Both intensity distributions of the two dispersive bands on modes Ak and Ck have the antisymmetric patterns within one period, which matches with the theoretical result in Eq. (5). We shall emphasize that the periodicity of the signal with a time window of 4π/Ω can also be noticed from the superposition term |ψkf,jA+ψkf,jC|2, which has unique characteristics from our system where signal amplitudes from modes An and Cn are mixed in the experiment. Furthermore, the roughness of transmission spectra in both Figs. 2 and 3 originates from the small display of the frequency detuning range for containing multiple sinusoidal signal periods [see Fig. S2(d) in the Supplementary Material].26

Next, we measure the frequency mode distributions for the case of BinBout by the heterodyne detection method51,52 to probe the localization effect of the flat band in the synthetic stub lattice. We connect the acousto-optic modulation path (Fig. S1 in the Supplementary Material) for frequency shift, and use it to interfere with the drop-port output of ring B by a 50:50 fiber coupler.53 To show evolutions of frequency modes throughout the whole band structure, we sweep the input laser frequency near the resonance frequency ω0 and process the drop-port output transmission through the fast Fourier transform.54 Figure 4(a) shows the experimentally resolved mode distributions as a function of frequency detuning Δω, where the intensities of modes are well confined near Δω0, which refers to the flat band, but spreads over the dispersive bands at Δω±0.07Ω. We explicitly exhibit the mode intensity distributions for two input frequencies in Fig. 4(b), which are Δω=0 at the flat band and Δω=0.08Ω at the upper dispersive band, respectively. For the input frequency at the flat band [see the left part of Fig. 4(b)], one sees that the intensities of modes Bn mainly locate at the zeroth and ±1st modes with a very small portion diverging to the ±2nd modes. On the other hand, intensities of modes experience spread for the input frequency located at the dispersive band [see the right part of Fig. 4(b)]. Simulations are performed by solving Eq. (6) with sweeping the input frequency and then Fourier transforming the transmitted intensity, in which the loss is chosen as γ=0.03Ω for better fitting with the experimental results. One can see a good agreement between experimental measurement in Figs. 4(a) and 4(b) and simulated results in Figs. 4(c) and 4(d), where the slight discrepancy between experiments and simulations originates from the experimental devices and the disturbance of the environment. The stability of the system can be further improved by utilizing polarization-maintaining fibers and devices or placing the experimental setup in the vacuum chamber.

Fig. 4

Mode distributions for the case of BinBout. (a) Experimentally resolved resonant mode spectra as a function of frequency detuning with VM=3  V. (b) The corresponding mode distributions of two selected input frequencies in (a) located at Δω=0 and Δω=0.08Ω, respectively. (c) Simulated resonant mode spectra with g=κ=0.06Ω, and (d) the corresponding intensity distributions of the two chosen input frequencies at Δω=0 and Δω=0.08Ω, respectively. The horizontal axis represents the mode number n for the frequency ωn.

AP_4_3_036002_f004.png

The existence of the flat band in the constructed synthetic stub lattice is not dependent on the coupling coefficients, i.e., g and κ for the weak modulation condition.39 Further increase of the modulation strength falls on the break of the synthetic stub lattice under the tight-binding limit. In this structure, one can make the band transition between the flat band and nonflat band by simply adding the higher-order modulation to introduce the long-range couplings in the frequency dimension, i.e., with an additional modulation frequency Ω, which makes the modulation 2gcos(Ωt/2+ϕ)+2gcos(Ωt+ϕ). The second term in the modulation brings next-nearest-nearby couplings between two nearby resonant modes An (or Cn). In the experiment, we apply the EOM in ring A with the corresponding form of VMcos(ΩMt+ϕ)+VMcos(2ΩMt+ϕ), where ΩM=2π·10  MHz and ϕ=ϕ=0.5π, and perform the measurements in the case of BinBout, which are shown in Fig. 5. Without higher-order modulation [see Fig. 5(a1) with VM=0], the system exhibits the feature of the flat band, which is the same as Figs. 2(c1) and 2(c2). Once the higher-order modulation term is added into the EOM (VM0), the middle band gradually turns dispersive, while the upper and lower dispersive bands start to show the nonsymmetrical feature, as shown in Figs. 5(b1)5(d1). In addition, the gap throughout the entire kf space gets closed if VM becomes larger [see Fig. 5(d1)]. We, therefore, show the transition from flat to nonflat bands in Figs. 5(a1)5(d1), which are in excellent agreement with the simulation results depicted in Figs. 5(a2)5(d2). The middle band exhibits a dispersive trend once the larger long-range couplings are added, which can lead to the delocalization effect, different from the localization effect observed in Fig. 4. We present this transition between localization and delocalization of light in the frequency dimension with simulation results shown in Fig. S3 in the Supplementary Material. Such an opportunity to dynamically introduce the band transition could be useful for light stopping, which has been proposed in theory.55,56

Fig. 5

Observations of flat-to-nonflat band transition for the case of BinBout. (a1)–(d1) Experimentally measured band structures with different long-range modulation amplitudes VM and fixed VM=2  V. (a2)–(d2) Simulation results of the projected intensity distribution of the band structure on mode Bk varied with g, where g=0.03Ω, γ=0.07Ω, and ϕ=ϕ=0.5π.

AP_4_3_036002_f005.png

4.

Conclusion

We have experimentally demonstrated a synthetic photonic stub lattice along the frequency axis of light, constructed by two coupled fiber ring resonators of different lengths. The flat-band feature is observed under two cases by selectively choosing the input and output ports for excitations and transmission measurements, which shows that measured band structures are intensity projections of the band on the different resonant modes in kf space. We also observed the localization effect near the flat band with distinctive features from dispersive bands and demonstrated the flat-to-nonflat band transition by adding the long-range couplings in modulations, characterizing the intrinsic physics of the synthetic stub lattice. Theoretical simulations performed agree well with experimental results, showing unique features of synthetic frequency dimensions in measuring signals of superposition modes. The construction of the stub lattice in two coupled rings of different lengths proves the experimental feasibility of connecting multiple rings of different types to construct a complicated lattice beyond the line or square geometry in the synthetic space. Our work also highlights potential toward non-Hermitian/topological5762 and quantum photonics6366 in coupled modulated ring resonator systems.

Acknowledgments

We greatly thank Prof. Shanhui Fan for fruitful discussions. The research was supported by National Natural Science Foundation of China (12104297, 12122407, and 11974245), National Key R&D Program of China (2017YFA0303701), Shanghai Municipal Science and Technology Major Project (2019SHZDZX01), Natural Science Foundation of Shanghai (19ZR1475700), and China Postdoctoral Science Foundation (2020M671090). L.Y. acknowledges support from the Program for Professor of Special Appointment (Eastern Scholar) at Shanghai Institutions of Higher Learning. X.C. also acknowledges the support from Shandong Quancheng Scholarship (00242019024).

References

1. 

L. Yuan et al., “Synthetic dimension in photonics,” Optica, 5 (11), 1396 –1405 (2018). https://doi.org/10.1364/OPTICA.5.001396 Google Scholar

2. 

T. Ozawa and H. M. Price, “Topological quantum matter in synthetic dimensions,” Nat. Rev. Phys., 1 (5), 349 –357 (2019). https://doi.org/10.1038/s42254-019-0045-3 Google Scholar

3. 

E. Lustig and M. Segev, “Topological photonics in synthetic dimensions,” Adv. Opt. Photonics, 13 (2), 426 –461 (2021). https://doi.org/10.1364/AOP.418074 AOPAC7 1943-8206 Google Scholar

4. 

L. Yuan, A. Dutt and S. Fan, “Synthetic frequency dimensions in dynamically modulated ring resonators,” APL Photonics, 6 (7), 071102 (2021). https://doi.org/10.1063/5.0056359 Google Scholar

5. 

A. Schwartz and B. Fischer, “Laser mode hyper-combs,” Opt. Express, 21 (5), 6196 –6204 (2013). https://doi.org/10.1364/OE.21.006196 OPEXFF 1094-4087 Google Scholar

6. 

C. Qin et al., “Spectrum control through discrete frequency diffraction in the presence of photonic gauge potentials,” Phys. Rev. Lett., 120 (13), 133901 (2018). https://doi.org/10.1103/PhysRevLett.120.133901 PRLTAO 0031-9007 Google Scholar

7. 

L. Yuan et al., “Synthetic space with arbitrary dimensions in a few rings undergoing dynamic modulation,” Phys. Rev. B, 97 (10), 104105 (2018). https://doi.org/10.1103/PhysRevB.97.104105 Google Scholar

8. 

E. Lustig et al., “Photonic topological insulator in synthetic dimensions,” Nature, 567 (7748), 356 –360 (2019). https://doi.org/10.1038/s41586-019-0943-7 Google Scholar

9. 

K. Wang et al., “Multidimensional synthetic chiral-tube lattices via nonlinear frequency conversion,” Light Sci. Appl., 9 (1), 132 (2020). https://doi.org/10.1038/s41377-020-0299-7 Google Scholar

10. 

A. Regensburger et al., “Photon propagation in a discrete fiber network: an interplay of coherence and losses,” Phys. Rev. Lett., 107 (23), 233902 (2011). https://doi.org/10.1103/PhysRevLett.107.233902 PRLTAO 0031-9007 Google Scholar

11. 

A. Regensburger et al., “Parity-time synthetic photonic lattices,” Nature, 488 (7410), 167 –171 (2012). https://doi.org/10.1038/nature11298 Google Scholar

12. 

X. Luo et al., “Quantum simulation of 2D topological physics in a 1D array of optical cavities,” Nat. Commun., 6 (1), 7704 (2015). https://doi.org/10.1038/ncomms8704 NCAOBW 2041-1723 Google Scholar

13. 

B. A. Bell et al., “Spectral photonic lattices with complex long-range coupling,” Optica, 4 (11), 1433 –1436 (2017). https://doi.org/10.1364/OPTICA.4.001433 Google Scholar

14. 

T. Ozawa and I. Carusotto, “Synthetic dimensions with magnetic fields and local interactions in photonic lattices,” Phys. Rev. Lett., 118 (1), 013601 (2017). https://doi.org/10.1103/PhysRevLett.118.013601 PRLTAO 0031-9007 Google Scholar

15. 

C. W. Peterson et al., “Strong nonreciprocity in modulated resonator chains through synthetic electric and magnetic fields,” Phys. Rev. Lett., 123 (6), 063901 (2019). https://doi.org/10.1103/PhysRevLett.123.063901 PRLTAO 0031-9007 Google Scholar

16. 

L. Yuan et al., “Photonic gauge potential in one cavity with synthetic frequency and orbital angular momentum dimensions,” Phys. Rev. Lett., 122 (8), 083903 (2019). https://doi.org/10.1103/PhysRevLett.122.083903 PRLTAO 0031-9007 Google Scholar

17. 

D. Cheng et al., “Arbitrary synthetic dimensions via multiboson dynamics on a one-dimensional lattice,” Phys. Rev. Res., 3 (3), 033069 (2021). https://doi.org/10.1103/PhysRevResearch.3.033069 PRSTCR 1554-9178 Google Scholar

18. 

G. Li et al., “Single pulse manipulations in synthetic time-frequency space,” Laser Photonics Rev., 16 2100340 (2022). https://doi.org/10.1002/lpor.202100340 Google Scholar

19. 

D. Jukić and H. Buljan, “Four-dimensional photonic lattices and discrete tesseract solitons,” Phys. Rev. A, 87 (1), 013814 (2013). https://doi.org/10.1103/PhysRevA.87.013814 Google Scholar

20. 

I. Petrides, H. M. Price and O. Zilberberg, “Six-dimensional quantum Hall effect and three-dimensional topological pumps,” Phys. Rev. B, 98 (12), 125431 (2018). https://doi.org/10.1103/PhysRevB.98.125431 Google Scholar

21. 

O. Zilberberg et al., “Photonic topological boundary pumping as a probe of 4D quantum Hall physics,” Nature, 553 (7686), 59 –62 (2018). https://doi.org/10.1038/nature25011 Google Scholar

22. 

L. Yuan, Y. Shi and S. Fan, “Photonic gauge potential in a system with a synthetic frequency dimension,” Opt. Lett., 41 (4), 741 –744 (2016). https://doi.org/10.1364/OL.41.000741 OPLEDP 0146-9592 Google Scholar

23. 

T. Ozawa et al., “Synthetic dimensions in integrated photonics: from optical isolation to four-dimensional quantum Hall physics,” Phys. Rev. A, 93 (4), 043827 (2016). https://doi.org/10.1103/PhysRevA.93.043827 Google Scholar

24. 

A. Dutt et al., “Experimental band structure spectroscopy along a synthetic dimension,” Nat. Commun., 10 (1), 3122 (2019). https://doi.org/10.1038/s41467-019-11117-9 NCAOBW 2041-1723 Google Scholar

25. 

A. Dutt et al., “A single photonic cavity with two independent physical synthetic dimensions,” Science, 367 (6473), 59 –64 (2020). https://doi.org/10.1126/science.aaz3071 SCIEAS 0036-8075 Google Scholar

26. 

G. Li et al., “Dynamic band structure measurement in the synthetic space,” Sci. Adv., 7 (2), eabe4335 (2021). https://doi.org/10.1126/sciadv.abe4335 STAMCV 1468-6996 Google Scholar

27. 

S. Weidemann et al., “Topological funneling of light,” Science, 368 (6488), 311 –314 (2020). https://doi.org/10.1126/science.aaz8727 SCIEAS 0036-8075 Google Scholar

28. 

Y. Hu et al., “Realization of high-dimensional frequency crystals in electro-optic microcombs,” Optica, 7 (9), 1189 –1194 (2020). https://doi.org/10.1364/OPTICA.395114 Google Scholar

29. 

H. Chen et al., “Real-time observation of frequency Bloch oscillations with fibre loop modulation,” Light Sci. Appl., 10 (1), 48 (2021). https://doi.org/10.1038/s41377-021-00494-w Google Scholar

30. 

K. Wang et al., “Generating arbitrary topological windings of a non-Hermitian band,” Science, 371 (6535), 1240 –1245 (2021). https://doi.org/10.1126/science.abf6568 SCIEAS 0036-8075 Google Scholar

31. 

A. Balčytis et al., “Synthetic dimension band structures on a Si CMOS photonic platform,” Sci. Adv., 8 (4), eabk0468 (2022). Google Scholar

32. 

K. Wang et al., “Topological complex-energy braiding of non-hermitian bands,” Nature, 598 59 –64 (2021). https://doi.org/10.1038/s41586-021-03848-x Google Scholar

33. 

D. Yu et al., “Topological holographic quench dynamics in a synthetic frequency dimension,” Light Sci. Appl., 10 (1), 209 (2021). https://doi.org/10.1038/s41377-021-00646-y Google Scholar

34. 

D. Yu, L. Yuan and X. Chen, “Isolated photonic flatband with the effective magnetic flux in a synthetic space including the frequency dimension,” Laser Photonics Rev., 14 (11), 2000041 (2020). https://doi.org/10.1002/lpor.202000041 Google Scholar

35. 

Q. Lin et al., “A three-dimensional photonic topological insulator using a two-dimensional ring resonator lattice with a synthetic frequency dimension,” Sci. Adv., 4 (10), eaat2774 (2018). https://doi.org/10.1126/sciadv.aat2774 STAMCV 1468-6996 Google Scholar

36. 

W. Zhang and X. Zhang, “Quadrupole topological phases in the zero-dimensional optical cavity,” Europhys. Lett., 131 (2), 24004 (2020). https://doi.org/10.1209/0295-5075/131/24004 EULEEJ 0295-5075 Google Scholar

37. 

A. Dutt et al., “Higher-order topological insulators in synthetic dimensions,” Light Sci. Appl., 9 (1), 131 (2020). https://doi.org/10.1038/s41377-020-0334-8 Google Scholar

38. 

M. Hyrkäs, V. Apaja and M. Manninen, “Many-particle dynamics of bosons and fermions in quasi-one-dimensional flat-band lattices,” Phys. Rev. A, 87 (2), 023614 (2013). https://doi.org/10.1103/PhysRevA.87.023614 Google Scholar

39. 

F. Baboux et al., “Bosonic condensation and disorder-induced localization in a flat band,” Phys. Rev. Lett., 116 (6), 066402 (2016). https://doi.org/10.1103/PhysRevLett.116.066402 PRLTAO 0031-9007 Google Scholar

40. 

S. Rojas-Rojas et al., “Quantum localized states in photonic flat-band lattices,” Phys. Rev. A, 96 (4), 043803 (2017). https://doi.org/10.1103/PhysRevA.96.043803 Google Scholar

41. 

B. Real et al., “Flat-band light dynamics in stub photonic lattices,” Sci. Rep., 7 (1), 15085 (2017). https://doi.org/10.1038/s41598-017-15441-2 SRCEC3 2045-2322 Google Scholar

42. 

M. N. Huda, S. Kezilebieke and P. Liljeroth, “Designer flat bands in quasi-one-dimensional atomic lattices,” Phys. Rev. Res., 2 (4), 043426 (2020). https://doi.org/10.1103/PhysRevResearch.2.043426 PRSTCR 1554-9178 Google Scholar

43. 

R. A. Vicencio et al., “Observation of localized states in Lieb photonic lattices,” Phys. Rev. Lett., 114 (24), 245503 (2015). https://doi.org/10.1103/PhysRevLett.114.245503 PRLTAO 0031-9007 Google Scholar

44. 

S. Mukherjee et al., “Observation of a localized flat-band state in a photonic Lieb lattice,” Phys. Rev. Lett., 114 (24), 245504 (2015). https://doi.org/10.1103/PhysRevLett.114.245504 PRLTAO 0031-9007 Google Scholar

45. 

D. Leykam, A. Andreanov and S. Flach, “Artificial flat band systems: from lattice models to experiments,” Adv. Phys.: X, 3 (1), 1473052 (2018). https://doi.org/10.1080/23746149.2018.1473052 Google Scholar

46. 

S. Xia et al., “Unconventional flatband line states in photonic Lieb lattices,” Phys. Rev. Lett., 121 (26), 263902 (2018). https://doi.org/10.1103/PhysRevLett.121.263902 PRLTAO 0031-9007 Google Scholar

47. 

J. Ma et al., “Direct observation of flatband loop states arising from nontrivial real-space topology,” Phys. Rev. Lett., 124 (18), 183901 (2020). https://doi.org/10.1103/PhysRevLett.124.183901 PRLTAO 0031-9007 Google Scholar

48. 

P. Karki and J. Paulose, “Stopping and reversing sound via dynamic dispersion tuning in a phononic metamaterial,” Phys. Rev. Appl., 15 (3), 034083 (2021). https://doi.org/10.1103/PhysRevApplied.15.034083 PRAHB2 2331-7019 Google Scholar

49. 

R. A. V. Poblete, “Photonic flat band dynamics,” Adv. Phys. X, 6 (1), 1878057 (2021). https://doi.org/10.1080/23746149.2021.1878057 Google Scholar

50. 

M. O. Scully and M. S. Zubairy, Quantum Optics, Cambridge University Press(1997). Google Scholar

51. 

T. Wildi et al., “Photo-acoustic dual-frequency comb spectroscopy,” Nat. Commun., 11 (1), 4146 (2020). https://doi.org/10.1038/s41467-020-17908-9 NCAOBW 2041-1723 Google Scholar

52. 

T. Tetsumoto et al., “Optically referenced 300 GHz millimetre-wave oscillator,” Nat. Photonics, 15 516 –522 (2021). https://doi.org/10.1038/s41566-021-00790-2 NPAHBY 1749-4885 Google Scholar

53. 

“See Supplemental Material for more details on the theory of band structure measurement in two rings and the experimental setup details,” Google Scholar

54. 

H. J. Nussbaumer, “The fast Fourier transform,” Fast Fourier Transform and Convolution Algorithms, 80 –111 Springer(1981). Google Scholar

55. 

M. F. Yanik and S. Fan, “Stopping light all optically,” Phys. Rev. Lett., 92 (8), 083901 (2004). https://doi.org/10.1103/PhysRevLett.92.083901 PRLTAO 0031-9007 Google Scholar

56. 

S. Sandhu et al., “Dynamically tuned coupled-resonator delay lines can be nearly dispersion free,” Opt. Lett., 31 (13), 1985 –1987 (2006). https://doi.org/10.1364/OL.31.001985 OPLEDP 0146-9592 Google Scholar

57. 

B. Peng et al., “Parity-time-symmetric whispering-gallery microcavities,” Nat. Phys., 10 (5), 394 –398 (2014). https://doi.org/10.1038/nphys2927 NPAHAX 1745-2473 Google Scholar

58. 

M. Wimmer et al., “Observation of optical solitons in PT-symmetric lattices,” Nat. Commun., 6 (1), 7782 (2015). https://doi.org/10.1038/ncomms8782 NCAOBW 2041-1723 Google Scholar

59. 

T. Ozawa et al., “Topological photonics,” Rev. Mod. Phys., 91 (1), 015006 (2019). https://doi.org/10.1103/RevModPhys.91.015006 RMPHAT 0034-6861 Google Scholar

60. 

Y. Song et al., “Two-dimensional non-Hermitian skin effect in a synthetic photonic lattice,” Phys. Rev. Appl., 14 (6), 064076 (2020). https://doi.org/10.1103/PhysRevApplied.14.064076 PRAHB2 2331-7019 Google Scholar

61. 

Q. Guo et al., “Experimental observation of non-Abelian topological charges and edge states,” Nature, 594 (7862), 195 –200 (2021). https://doi.org/10.1038/s41586-021-03521-3 Google Scholar

62. 

Z. Chen and M. Segev, “Highlighting photonics: looking into the next decade,” eLight, 1 (2), 2 (2021). https://doi.org/10.1186/s43593-021-00002-y Google Scholar

63. 

J. Boutari et al., “Large scale quantum walks by means of optical fiber cavities,” J. Opt., 18 (9), 094007 (2016). https://doi.org/10.1088/2040-8978/18/9/094007 Google Scholar

64. 

C. Chen et al., “Observation of topologically protected edge states in a photonic two-dimensional quantum walk,” Phys. Rev. Lett., 121 (10), 100502 (2018). https://doi.org/10.1103/PhysRevLett.121.100502 PRLTAO 0031-9007 Google Scholar

65. 

H. Chalabi et al., “Synthetic gauge field for two-dimensional time-multiplexed quantum random walks,” Phys. Rev. Lett., 123 (15), 150503 (2019). https://doi.org/10.1103/PhysRevLett.123.150503 PRLTAO 0031-9007 Google Scholar

66. 

C. Joshi et al., “Frequency-domain quantum interference with correlated photons from an integrated microresonator,” Phys. Rev. Lett., 124 (14), 143601 (2020). https://doi.org/10.1103/PhysRevLett.124.143601 PRLTAO 0031-9007 Google Scholar

Biography

Guangzhen Li is a assistant researcher in the School of Physics and Astronomy at Shanghai Jiao Tong University (SJTU). She received her PhD in physics from SJTU in 2017 and then worked as a postdoctoral and assistant researcher in the Laboratory of Advanced Photonic Materials and Physics (LAPMP) at SJTU. Her current research interests are synthetic dimensions in photonics, topological photonics, nonlinear optics, and integrated photonics.

Luojia Wang is a assistant researcher in the School of Physics and Astronomy at SJTU. She received her PhD from Peking University in 2013. She worked as a postdoctoral scholar at Texas A&M University, Tongji University, and SJTU, respectively. Her research interest focuses on the theoretical study of quantum optics, nanophotonics, ultrafast nonlinear optics, and topological photonics.

Rui Ye is a PhD student in the School of Physics and Astronomy at SJTU. He received his bachelor’s degree from the School of Physical Science and technology at Northwestern Polytechnical University in 2020. His research focuses on synthetic dimensions in photonics, topological photonics, and integrated photonics.

Shijie Liu is a postdoctoral researcher in the School of Physics and Astronomy at SJTU. He obtained his PhD in physics from SJTU in 2019 and worked as a postdoctoral researcher in LAPMP at SJTU. His research is focused on the fabrication of integrated photonic devices and nonlinear optics.

Yuanlin Zheng is an associate researcher at SJTU’s School of Physics and Astronomy. He obtained his PhD in optics from SJTU in 2013 and subsequently worked as a postdoctoral, assistant, and associate researcher in LAPMP at SJTU. His research focuses on nonlinear optics and integrated photonics for light-matter interaction and light manipulation applications.

Luqi Yuan is currently on the faculty in School of Physics and Astronomy at SJTU. He received his PhD in physics from Texas A&M University in 2014 and was a postdoctoral scholar from 2014 to 2018 at Stanford University. His research interests span broad fields among quantum optics, photonics, AMO physics, and nonlinear optics, including nanophotonics, topological photonics, synthetic dimensions in photonics, hybrid quantum systems, and light-matter interactions.

Xianfeng Chen is a distinguished professor in School of Physics and Astronomy at SJTU. He obtained his PhD in physics at SJTU in 1999. He is the executive director of the research center for optical science and engineering at SJTU. His current research focuses on advanced photonics materials and devices, nonlinear optics, nanophotonics, quantum optics, and ultrafast optics.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Guangzhen Li, Luojia Wang, Rui Ye, Shijie Liu, Yuanlin Zheng, Luqi Yuan, and Xianfeng Chen "Observation of flat-band and band transition in the synthetic space," Advanced Photonics 4(3), 036002 (21 June 2022). https://doi.org/10.1117/1.AP.4.3.036002
Received: 11 February 2022; Accepted: 25 May 2022; Published: 21 June 2022
Lens.org Logo
CITATIONS
Cited by 10 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Modulation

Resonators

Physics

Band structure simulations

Fiber couplers

Superposition

Waveguides

Back to Top