Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Z2 topological edge state in honeycomb lattice of coupled resonant optical waveguides with a flat band

Open Access Open Access

Abstract

Two-dimensional (2D) coupled resonant optical waveguide (CROW), exhibiting topological edge states, provides an efficient platform for designing integrated topological photonic devices. In this paper, we propose an experimentally feasible design of 2D honeycomb CROW photonic structure. The characteristic optical system possesses two-fold and three-fold Dirac points at different positions in the Brillouin zone. The effective gauge fields implemented by the intrinsic pseudo-spin-orbit interaction open up topologically nontrivial bandgaps through the Dirac points. Spatial lattice geometries allow destructive wave interference, leading to a dispersionless, near-flat energy band in the vicinity of the three-fold Dirac point in the telecommunication frequency regime. This nontrivial structure with a near-flat band yields topologically protected edge states. These characteristics underpin the fundamental importance as well as the potential applications in various optical devices. Based on the honeycomb CROW lattice, we design the shape-independent topological cavity and the beam splitter, which demonstrate the relevance for a wide range of photonic applications.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

The exploration of various topological states in quantum systems mainly focuses on condensed-matter and artificial photonic systems [1–3]. Optical topological states may have important applications in optical devices due to its characteristic robustness. By breaking time-reversal symmetry using external magnetic fields, gapless topological edge states can emerge in gyromagnetic photonic crystals [4–6]. Alternatively, topological edge states can be realized by introducing effective magnetic gauge fields originating from periodic time-dependent [7] or spatial z-direction [8] modulations. However, the integration of time-dependent and z-direction modulations into semiconductor platforms is still quite challenging, while magnetic fields are incompatible with integrated miniature optical chips. Recently, 2D coupled resonant optical waveguide (CROW) which exhibits topologically protected edge state has been theoretically proposed [9, 10] and experimentally demonstrated [11, 12]. 2D dielectric CROWs, which can be fabricated using standard nanofabrication processes and are compatible with conventional photonic integrated circuits, pave promising ways toward manipulation of light through topological edge states.

A flat band is a dispersionless band that extents in the whole Brillouin zone. The flat band dispersion leads to a zero group velocity, which has important applications on the dynamic and localization properties, such as slow light devices [13, 14], nonlinear enhancement [15–17], zero index metamaterials [18, 19] and so on. And the nontrivial topological structure with a flat band combines dispersionless and robust edge state in an optical system to make it more tunable and functional. Flat bands or near flat bands has been present in some kinds of photonic structures, for example in a Lieb lattice [20–22], polaritonic systems [23], or waveguide networks [24]. However, the CROW mode is easier to couple with on-chip devices and does not require the magneto-optical materials, which means more compatibility and integration.

In this Letter, we design a 2D honeycomb CROW photonic structure which supports both two-fold and three-fold Dirac points at different high symmetry points in the Brillouin zone. The designed lattice geometry gives rise to destructive wave interference which results in a near-flat photonic band in the telecommunication frequency regime. The effective gauge fields implemented by the pseudo-spin-orbit interaction in the CROW opens up bandgaps with nontrivial topology through the Dirac points. The collective localization in the near-flat band depends on the design of lattice geometry, impurities or structural disorders [20–22, 25, 26]. These modes have a large modal area, similar to the bulk propagating ones, but with a much lower group velocity. Gapless edge states between the near-flat band and the adjacent energy bands are revealed. The field distributions of CROW structure confirm the robustness of topological edge states. A 2D tight-binding model is then set up to analyze both the bulk and edge spectra, which are in conformity with the numerical simulations. Based on these intriguing properties, the proposed honeycomb CROW structure may play a crucial role in the optical devices such as slow light devices [14, 27], topological cavity [28], large-mode-area laser [29, 30] and so on. Based on the honeycomb CROW lattice, we design and simulate the shape-independent topological cavity and the beam splitter, which demonstrate the potential application of honeycomb CROW structure in practical optical devices and new possibilities for photonic engineering. In the viewpoint of general wave theory, the underlying physics and exotic phenomena can be extended to other regimes, such as acoustics [31–34], plasmonics [35] and other systems.

2. Structure design

The designed honeycomb CROW structure is shown in Fig. 1(a). The ring waveguides with refractive index n=3 are periodically arranged in the x-y plane. The topological edge states appear at the boundary of the structure, as shown by the red curves in the Fig. 1(a).The part of the white dashed box in Fig. 1(a) represents the unit cell of the honeycomb lattice, while the detailed structure is shown in Fig. 1(b). In this structure, two identical optical ring waveguides occupy the “site” positions of the graphene-like configuration (denoted as “site” rings with an inner diameter Rs=1.653um). The optical ring waveguide supports both clockwise and counterclockwise modes, which constitute the two pseudo-spins (pseudo-spin up (down) denote the clockwise (anticlockwise) mode) in optical quantum spin Hall effect [10, 11, 36, 37]. The coupling between the clockwise and counterclockwise modes in an isolated ring is forbidden by time-reversal and rotation symmetries. Each site ring waveguide is connected with its nearest-neighbor site-ring waveguides through three coupling waveguides which are denoted as “link” rings with the same inner diameter Rl=1.653um. Five ring waveguides in unit cell are numbered by τ=1,...,5, as shown in Fig. 1(b). By designing directional coupling between two adjacent rings, the optical mode in the link ring always undergoes opposite rotation as compared to when it is in the site ring. As shown in the left inset of Fig. 1(a), the widths (outer radius minus inner radius) of both the site and link rings are the same w=0.2um. A small space gap is assigned between adjacent rings g=0.1um, allowing efficient coupling between them. In the frequency range of 190.0 THz to 196.0 THz, the transmission of the ring resonator is over 20% (the right inset of Fig. 1(a)). Such efficient coupling gives rise to the photonic energy bands in the CROW lattice.

 figure: Fig. 1

Fig. 1 (a) a schematic diagram of honeycomb CROW structures. The topological edge state are shown as the red line. The insets represent the size parameters and the transmission/reflection spectra of ring resonators, respectively. (b) The detailed unit cell diagram of honeycomb CROW structures. The black dashed line represents a single repeating unit, and the arrows denote three nearest neighbor vectors a1=(0,1),a2=(32,12),a3=(32,12).

Download Full Size | PDF

A tight-binding model is employed to describe the honeycomb CROW lattice, which results in the following Hamiltonian [11, 12, 26]:

H0=<ij>tijcicj
where ci and ci are the photon creation and annihilation operators at lattice site i, respectively, t is the tunnel coupling between two adjacent rings. The two pseudospin states are simple duplication [10, 38], so we only consider one of them in the Hamiltonian. The photonic band structure of the CROW system can be obtained from the Hamiltonian in wavevector spaceH=kΨkHkΨk, where Ψk=(c1k,c2k,c3k,c4k,c5k)Tand
Hk=(0eik·a2/20eik·a2/20eik·a2/20eik·a1/20eik·a3/20eik·a1/20eik·a1/20eik·a2/20eik·a1/20eik·a3/20eik·a3/20eik·a3/20)
The three nearest neighbor vectors of the honeycomb structure (site rings) are given by a1=(0,1),a2=(32,12),a3=(32,12). The energy bands of the Hamiltonian are obtained by diagonalizing the above matrix. At the middle of the photonic spectrum, a near-flat band appears across the entire Brillouin zone and the wave functions associated with the flat band have zero amplitude at link rings, demonstrating collective localization phenomenon. Indeed, the Hamiltonian matrix can be block-diagonalized into two subsets: one subset constitutes only site-rings, while the other consists of only link-rings. The decoupling of our system arises from the fact that the honeycomb CROW lattice has a bipartite structure which allows for an effective decoupling between the site rings (τ=2,4) and the link rings (τ=1,3,5). And it is worth noting that while the two site rings form the background honeycomb lattice, the three link ring waveguides constitute a kagome lattice [39]. It turns out that the band structure can be derived from the energy spectrum describing these two sublattices [40]. The single-particle Schrödinger equation describing noninteracting particles on the honeycomb CROW structure can be directly derived from the second-quantized Hamiltonian. Here we consider the first-quantized Hamiltonian and the ϕτk is introduced to represent the wave function corresponding to the cτk, one can find two separate eigensystems given by:
εΓ(k)(ϕ1ϕ3ϕ5)=2(0cosk(a2a1)/2cosk(a2a3)/2cosk(a2a1)/20cosk(a3a1)/2cosk(a2a3)/2cosk(a3a1)/20)(ϕ1ϕ3ϕ5)=HΓ(k)(ϕ1ϕ3ϕ5)
and
εK(k)(ϕ2ϕ4)=2(0v=13exp(ikav)v=13exp(ikav)0)(ϕ2ϕ4)=HK(k)(ϕ2ϕ4)
The two decoupled systems, described by the HamiltoniansHΓ(k) andHK(k), correspond to the link and site rings respectively, with the energy bands given by:
εK(k)=(1±4Ak3),εK(k)=2,εΓ(k)=±|exp(ika1)+exp(ika2)+exp(ika3)|
where Ak=cos2|k(a2a3)/2|+cos2|k(a3a1)/2|+cos2|k(a1a2)/2|. The theoretical result from tight-binding model is shown in Fig. 2(a). As we can see in formula (5), an eigenvalue with near-flat dispersion appears. The bipartite CROW lattice is similar to the edge-centered square (Lieb) lattice [20, 21]. The flat band arises from the destructive wave interference and can be described by Lieb theorem.

 figure: Fig. 2

Fig. 2 The energy bands E(k) corresponding to the Hamiltonian (a) without the Spin-Orbit term, and (b) with the Spin-Orbit coupling term.

Download Full Size | PDF

An intrinsic pseudo-spin-orbit coupling term HSO=iλso<ij>αβ[(dij1×dij2)σ]αβciαcjβ is introduced to account for next-nearest-neighbor hopping in the honeycomb lattice. α and β represent the pseudo spin indexes. Here the λSO corresponds to the next-nearest-neighbor (NNN) coupling strength, dij1 and dij2 are two vectors connecting the NNN site ring, and σ stands for the Pauli matrices for the pseudo-spins. In this system, the time reversal symmetry and the pseudo time reversal symmetry are always maintained. Here the time reversal operator exchanges the wave functions of the two basis without changing the phase, and the pseudo time reversal symmetry operator has the same form as the fermionic time reversal symmetry operator, under which the clockwise and counterclockwise modes becomes a photonic Kramers doublet [3]. And time reversal is a special case for the reciprocity. If there is no absorption and time reversal is kept, then the reciprocity of our CROW system holds [41]. It is worth noting that each spin sector time-reversal symmetry is effectively broken by the pseudo-spin-orbit coupling term, enabling nontrivial energy gaps and topological edge states that are backscattering-immune as long as the spins are decoupled . In CROW lattice, the pseudo spin-orbit coupling term results from the direction-dependent hopping phase, which arises from the difference in the path lengths traveled in the linking rings [38]. In the vicinity of the high symmetry points, the effective Hamiltonians become:

hpK=vK(pxσ1+pyσ2)32αλsoσ3hpΓ=vΓ(pxS1+pyS2)23λsoS3
where p=kK±(Γ0) and α is the spin index. This pseudo-spin-orbit coupling term opens up gaps in the bulk bands associated with the nontrivial Z2 index [26], and the theoretical results from tight-binding model are shown in Fig. 2(b). We also simulate and obtain the photonic bands numerically by full wave simulation, which are shown in Fig. 3. The existence of the flat band coincides with the picture illustrated in the tight-binding model. It is worth noting that topological phase of honeycomb CROW was discussed in [42]. But it was described by the network model and the link sites were simplified as a coupling term in the theoretical model. However, the flat band appears only when the link rings are considered as lattice sites and the Hamiltonian matrix is expanded to 5×5.

 figure: Fig. 3

Fig. 3 (a) Schematic diagram of the energy band. (b) Schematic diagram of the projected band. The gapless edge states are shown in the red and blue dots, which corresponds to the upper and lower borders, respectively. The calculation interval ofkx isΓKΓ (ΓandK are high symmetric points in the first Brillouin zone), and kyis always zero in the calculation.

Download Full Size | PDF

The bulk and edge photonic spectrum of a strip of the honeycomb CROW structure is presented in Fig. 3(b) for zigzag edges. The gapless edge states in the band diagram are for the upper and lower boundaries for the pseudo-spin up mode (shown in Fig. 3(b) as the red and blue dots), respectively.

In full-wave simulation, the pseudo spin-orbit coupling term results from the direction-dependent hopping phase, which arises from the difference in the path lengths traveled in the linking rings. And the effective coupling can be derived by considering two site rings connected by a link ring through full wave calculation, as shown in the inset of Fig. 1(a). The hopping phase can be estimated by the formula: ϕ=2πΔv/FSR [11, 38], where the FSR9e3GHzrepresents the free spectral range of the ring. The resonance frequency of the site rings is detuned from the link rings byΔv600GHz. J15GHz represents the intersite couping strength for antiresonant sites and links. As we can see,J(Δv,FSR), the rings are weakly coupled high quality factor resonators. When we consider the next nearest neighbor hopping between two link rings, the hopping phase can be estimated by Jexp(iϕ/3)csc(ϕ/2), and Jexp(iϕ/2)csc(ϕ/2)for the next nearest neighbor hopping between two site rings in full wave simulation. The pseudo spin orbit coupling term is related to the imaginary part of the hopping phase.

3. Nontrivial Z2 index

In this part, we compute the four Z2 indices associated with the four bulk gaps opened by the Pseudo-spin-orbit coupling term in the honeycomb CROW structure. Since the honeycomb structure possesses inversion symmetry, the Z2 indicesvN associated with the N-th bulk gap can be easily computed through the following formula [43, 44]:

i=03m=1Nξ2m(Γi)=(1)vN

In this formula, ξ2m(Γi)=±1 represents the parity eigenvalue associated with the 2m-th energy bands. All the eigenvalues combine two degenerated pseudo-spin states, corresponding to the clockwise and anti-clockwise states of the site ring. So we only consider the even numbered energy bands. As shown in Fig. 4, we calculate the field distributions of the honeycomb CROW unit cell through full wave simulation at the four high symmetric points: Γ0=(0,0), Γ1=(0,2π3a), Γ2=(3π3a,π3a),Γ3=(3π3a,π3a),where ‘a’ represents the distance between two adjacent site ring center.

 figure: Fig. 4

Fig. 4 The field distributions at the four high symmetric points for the energy bands.

Download Full Size | PDF

Choosing the center site ring inside the unit cell as the center of inversion, the parity eigenvalues of the even numbered energy bands are determined. For example, the field distribution of band 1 calculated by full wave simulation at high symmetric point Γ0=(0,0)is shown at the left lower corner of Fig. 4. As we can see, the phase in the ring waveguides coincides after inversion transformation, so the parity eigenvalue is + 1. On the contrary, the phase of band 1 at Γ1=(0,2π3a) becomes opposite after inversion transformation, which means the parity eigenvalue is −1. All the parity eigenvalues of four bands at different high symmetric point can be obtained, as shown in Table 1. We can see that the formula (7) gives (1)v=1 for every bulk gap, which essentially means that the Z2 indices generated by the Pseudo-spin-orbit term are all nontrivial.

Tables Icon

Table 1. Parity-eigenvalue patterns at the four high symmetric points for the four different energy bands. All the energy gaps are associated with a nontrivial Z2 index .

4. Robust edge state and flat band state

Through full-wave simulation, we obtain the field distribution of the topologically protected edge states in a finite-size lattice of 7 × 4 unit-cells. The excitation port is on the right-lower corner of the structure, as shown in the inset of Fig. 5(a). The photonic edge states are shown to be robust against site defects [Fig. 5(b)], where the inset represents the specific location of the site defect. Robust unidirectional light propagation around the site defect is observed. Besides, the edge state can also be robust against an link ring defect as well as an additional site ring.

 figure: Fig. 5

Fig. 5 The field distributions of topologically protected edge state and the bulk localized mode. (a) The field distribution of topologically protected edge state, white dashed box represents the input port of CROW structure. (b)Robust edge state propagation with the site ring defect.

Download Full Size | PDF

In the honeycomb CROW structure, the near-flat band can be successfully excited in the frequency range of 193.07 THz to 193.10 THz by inputting the mode light into the lower right port of the structure [Fig. 6(a)]. It is worth noting that different field distributions can be excited by various excitation approaches or excitation positions, as the near-flat band is composed of several states. As an example, we notice that using the dipole light source (at the end of the white arrows in Fig. 6(b)) to stimulate the flat band, the excitation of the bulk mode field distribution is significantly different. The dispersionless collective localization modes have a large mode area, similar to the propagation mode, but a lower group velocity. Consequently, the propagation velocity of the light waves is reduced, which may be suggestive of potential applications to slow light devices and large mode area laser.

 figure: Fig. 6

Fig. 6 The field distributions of the collective localization modes. (a) The flat band is activated by inputting the mode light into the lower right port of the structure. (b) The flat band is activated by using the dipole light source (at the end of the white arrows).

Download Full Size | PDF

5. Applications

The above topologically protected edge state for the photonic systems can be leveraged to design counter-intuitive devices that have immediate applications. An example of significance is a topological cavity of arbitrary geometry. Resonant cavities are essential derives governing many photonic phenomena, however their geometry limit the application of optical devices. The geometry-independent topological cavities provides the opportunity to develop complex topological circuitry and pave promising ways toward light manipulation through topological edge states. Our design and the full-wave simulation result are shown in Fig. 7(a). Here, we use the dipole light source to stimulate the cavity mode. In addition, we have also designed an optical beam splitter based on the honeycomb CROW structure. As shown in Fig. 7(b), the dipole light source is at the end of the white arrow. When the light source simultaneously excites the clockwise and counterclockwise modes of the input ring waveguide, the clockwise and counterclockwise edge states can be observed simultaneously in the structure, and the light beam is separated to the upper right and lower left, respectively.

 figure: Fig. 7

Fig. 7 (a) The field distribution of edge state in an arbitrarily shaped topological cavity. The cavity mode is activated by using the dipole light source (at the end of the white arrows). (b) The field distribution of edge state in a topological beam splitter.

Download Full Size | PDF

6. Conclusions

In conclusion, we have designed a topological honeycomb CROW structure containing a near-flat band and topologically protected edge states. The system has been modeled by a 2D tight binding framework which efficiently corroborates the nontrivial energy band gap and the near-flat band. In the honeycomb CROW structure the excitations of topologically protected edge states and pertinent effects have been elucidated. The topologically protected edge states can be used in the asymmetric selective control of light intensity in different boundaries. The collective localized modes formed due to the near-flat band possess large mode area and lower group velocity which may be applied to large mode area laser and slow light devices. These characteristics may be suggestive of potential application prospects and research values in optical devices such as optical couplers, beam splitters and so on. Additionally, the photonic systems in analogy to the quantum or electronic systems can be used for controlling some exotic optical properties, which can envisage an excellent optical platform for quantum simulations.

Funding

National Key Research and Development Program of China (NKRDP) (2017YFA0303702 and 2017YFA0305100); National Natural Science Foundation of China (NNSFC) (Grant No.11625418, No. 11474158, No. 51732006, No. 51721001, No. 51472114 and No. 11675116); Natural Science Foundation of Jiangsu Province (BK20140019)

References and links

1. F. D. Haldane and S. Raghu, “Possible realization of directional optical waveguides in photonic crystals with broken time-reversal symmetry,” Phys. Rev. Lett. 100(1), 013904 (2008). [CrossRef]   [PubMed]  

2. A. B. Khanikaev, S. H. Mousavi, W. K. Tse, M. Kargarian, A. H. MacDonald, and G. Shvets, “Photonic topological insulators,” Nat. Mater. 12(3), 233–239 (2013). [CrossRef]   [PubMed]  

3. C. He, X.-C. Sun, X.-P. Liu, M.-H. Lu, Y. Chen, L. Feng, and Y.-F. Chen, “Photonic topological insulator with broken time-reversal symmetry,” Proc. Natl. Acad. Sci. U.S.A. 113(18), 4924–4928 (2016). [CrossRef]   [PubMed]  

4. Z. Wang, Y. Chong, J. D. Joannopoulos, and M. Soljacić, “Observation of unidirectional backscattering-immune topological electromagnetic states,” Nature 461(7265), 772–775 (2009). [CrossRef]   [PubMed]  

5. Z. Wang, Y. D. Chong, J. D. Joannopoulos, and M. Soljacić, “Reflection-free one-way edge modes in a gyromagnetic photonic crystal,” Phys. Rev. Lett. 100(1), 013905 (2008). [CrossRef]   [PubMed]  

6. S. Raghu and F. D. Haldane, “Analogs of quantum-Hall-effect edge states in photonic crystals,” Phys. Rev. A 78(3), 033834 (2008). [CrossRef]  

7. K. Fang, Z. Yu, and S. Fan, “Realizing effective magnetic field for photons by controlling the phase of dynamic modulation,” Nat. Photonics 6(11), 782–787 (2012). [CrossRef]  

8. M. C. Rechtsman, J. M. Zeuner, Y. Plotnik, Y. Lumer, D. Podolsky, F. Dreisow, S. Nolte, M. Segev, and A. Szameit, “Photonic Floquet topological insulators,” Nature 496(7444), 196–200 (2013). [CrossRef]   [PubMed]  

9. D. Leykam, K. Y. Bliokh, C. Huang, Y. D. Chong, and F. Nori, “Edge modes, degeneracies, and topological numbers in non-hermitian systems,” Phys. Rev. Lett. 118(4), 040401 (2017). [CrossRef]   [PubMed]  

10. G.-Q. Liang and Y. D. Chong, “Optical resonator analog of a two-dimensional topological insulator,” Phys. Rev. Lett. 110(20), 203904 (2013). [CrossRef]   [PubMed]  

11. M. Hafezi, S. Mittal, J. Fan, A. Migdall, and J. M. Taylor, “Imaging topological edge states in silicon photonics,” Nat. Photonics 7(12), 1001–1005 (2013). [CrossRef]  

12. M. Hafezi, E. A. Demler, M. D. Lukin, and J. M. Taylor, “Robust optical delay lines with topological protection,” Nat. Phys. 7(11), 907–912 (2011). [CrossRef]  

13. J. Hou, D. Gao, H. Wu, R. Hao, and Z. Zhou, “Flat band slow light in symmetric line defect photonic crystal waveguides,” IEEE Photonics Technol. Lett. 21(20), 1571–1573 (2009). [CrossRef]  

14. M. D. Settle, R. J. Engelen, M. Salib, A. Michaeli, L. Kuipers, and T. F. Krauss, “Flatband slow light in photonic crystals featuring spatial pulse compression and terahertz bandwidth,” Opt. Express 15(1), 219–226 (2007). [CrossRef]   [PubMed]  

15. J. Li, L. O’Faolain, I. H. Rey, and T. F. Krauss, “Four-wave mixing in photonic crystal waveguides: slow light enhancement and limitations,” Opt. Express 19(5), 4458–4463 (2011). [CrossRef]   [PubMed]  

16. C. Monat, B. Corcoran, M. Ebnali-Heidari, C. Grillet, B. J. Eggleton, T. P. White, L. O’Faolain, and T. F. Krauss, “Slow light enhancement of nonlinear effects in silicon engineered photonic crystal waveguides,” Opt. Express 17(4), 2944–2953 (2009). [CrossRef]   [PubMed]  

17. M. Shinkawa, N. Ishikura, Y. Hama, K. Suzuki, and T. Baba, “Nonlinear enhancement in photonic crystal slow light waveguides fabricated using CMOS-compatible process,” Opt. Express 19(22), 22208–22218 (2011). [CrossRef]   [PubMed]  

18. X. Huang, Y. Lai, Z.-H. Hang, H. Zheng, and C. T. Chan, “Dirac cones induced by accidental degeneracy in photonic crystals and zero-refractive-index materials,” Nat. Mater. 10(8), 582–586 (2011). [CrossRef]   [PubMed]  

19. P. Moitra, Y. Yang, Z. Anderson, I. I. Kravchenko, D. P. Briggs, and J. Valentine, “Realization of an all-dielectric zero-index optical metamaterial,” Nat. Photonics 7(10), 791–795 (2013). [CrossRef]  

20. 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(24), 245504 (2015). [CrossRef]   [PubMed]  

21. 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(24), 245503 (2015). [CrossRef]   [PubMed]  

22. F. Diebel, D. Leykam, S. Kroesen, C. Denz, and A. S. Desyatnikov, “Conical diffraction and composite Lieb bosons in photonic lattices,” Phys. Rev. Lett. 116(18), 183902 (2016). [CrossRef]   [PubMed]  

23. C. E. Whittaker, E. Cancellieri, P. M. Walker, D. R. Gulevich, H. Schomerus, D. Vaitiekus, B. Royall, D. M. Whittaker, E. Clarke, I. V. Iorsh, I. A. Shelykh, M. S. Skolnick, and D. N. Krizhanovskii, “Exciton Polaritons in a Two-Dimensional Lieb Lattice with Spin-Orbit Coupling,” Phys. Rev. Lett. 120(9), 097401 (2018). [CrossRef]   [PubMed]  

24. Q. Lin and S. Fan, “Resonator-free realization of effective magnetic field for photons,” New J. Phys. 17(7), 075008 (2015). [CrossRef]  

25. K. Sun, Z. Gu, H. Katsura, and S. Das Sarma, “Nearly flatbands with nontrivial topology,” Phys. Rev. Lett. 106(23), 236803 (2011). [CrossRef]   [PubMed]  

26. Z. Lan, N. Goldman, and P. Öhberg, “Coexistence of spin-1/2 and spin-1 Dirac-Weyl fermions in the edge-centered honeycomb lattice,” Phys. Rev. B 85(15), 155451 (2012). [CrossRef]  

27. J. Li, T. P. White, L. O’Faolain, A. Gomez-Iglesias, and T. F. Krauss, “Systematic design of flat band slow light in photonic crystal waveguides,” Opt. Express 16(9), 6227–6232 (2008). [CrossRef]   [PubMed]  

28. B. Bahari, A. Ndao, F. Vallini, A. El Amili, Y. Fainman, and B. Kanté, “Nonreciprocal lasing in topological cavities of arbitrary geometries,” Science 358(6363), 636–640 (2017). [CrossRef]   [PubMed]  

29. G. Harari, M. A. Bandres, Y. Lumer, M. C. Rechtsman, Y. D. Chong, M. Khajavikhan, D. N. Christodoulides, and M. Segev, “Topological insulator laser: Theory,” Science 359(6381), eaar4003 (2018). [CrossRef]   [PubMed]  

30. M. A. Bandres, S. Wittek, G. Harari, M. Parto, J. Ren, M. Segev, D. N. Christodoulides, and M. Khajavikhan, “Topological insulator laser: Experiments,” Science 359(6381), eaar4005 (2018). [CrossRef]   [PubMed]  

31. Y. G. Peng, C. Z. Qin, D. G. Zhao, Y. X. Shen, X. Y. Xu, M. Bao, H. Jia, and X. F. Zhu, “Experimental demonstration of anomalous Floquet topological insulator for sound,” Nat. Commun. 7, 13368 (2016). [CrossRef]   [PubMed]  

32. C. He, X. Ni, H. Ge, X.-C. Sun, Y.-B. Chen, M.-H. Lu, X.-P. Liu, and Y.-F. Chen, “Acoustic topological insulator and robust one-way sound transport,” Nat. Phys. 12(12), 1124–1129 (2016). [CrossRef]  

33. X. Ni, C. He, X.-C. Sun, X.-P. Liu, M.-H. Lu, L. Feng, and Y.-F. Chen, “Topologically protected one-way edge mode in networks of acoustic resonators with circulating air flow,” New J. Phys. 17(5), 053016 (2015). [CrossRef]  

34. C. He, Z. Li, X. Ni, X.-C. Sun, S.-Y. Yu, M.-H. Lu, X.-P. Liu, and Y.-F. Chen, “Topological phononic states of underwater sound based on coupled ring resonators,” Appl. Phys. Lett. 108(3), 031904 (2016). [CrossRef]  

35. F. Gao, Z. Gao, X. Shi, Z. Yang, X. Lin, H. Xu, J. D. Joannopoulos, M. Soljačić, H. Chen, L. Lu, Y. Chong, and B. Zhang, “Probing topological protection using a designer surface plasmon structure,” Nat. Commun. 7, 11619 (2016). [CrossRef]   [PubMed]  

36. C. L. Kane and E. J. Mele, “Quantum spin Hall effect in graphene,” Phys. Rev. Lett. 95(22), 226801 (2005). [CrossRef]   [PubMed]  

37. C. L. Kane and E. J. Mele, “Z2 topological order and the quantum spin Hall effect,” Phys. Rev. Lett. 95(14), 146802 (2005). [CrossRef]   [PubMed]  

38. D. Leykam, S. Mittal, M. Hafezi, and Y. D. Chong, “Reconfigurable topological phases in next-nearest-neighbor coupled resonator lattices,” Phys. Rev. Lett. 121(2), 023901 (2018). [CrossRef]   [PubMed]  

39. H. M. Guo and M. Franz, “Topological insulator on the kagome lattice,” Phys. Rev. B 80(11), 113102 (2009). [CrossRef]  

40. B. Sutherland, “Localization of electronic wave functions due to local topology,” Phys. Rev. B Condens. Matter 34(8), 5208–5211 (1986). [CrossRef]   [PubMed]  

41. L. Deák and T. Fülöp, “Reciprocity in quantum, electromagnetic and other wave scattering,” Ann. Phys. 327(4), 1050–1077 (2012). [CrossRef]  

42. M. Pasek and Y. D. Chong, “Network models of photonic Floquet topological insulators,” Phys. Rev. B 89(7), 075113 (2014). [CrossRef]  

43. T. L. Hughes, E. Prodan, and B. A. Bernevig, “Inversion-symmetric topological insulators,” Phys. Rev. B 83(24), 245132 (2011). [CrossRef]  

44. L. Fu and C. L. Kane, “Topological insulators with inversion symmetry,” Phys. Rev. B 76(4), 045302 (2007). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (7)

Fig. 1
Fig. 1 (a) a schematic diagram of honeycomb CROW structures. The topological edge state are shown as the red line. The insets represent the size parameters and the transmission/reflection spectra of ring resonators, respectively. (b) The detailed unit cell diagram of honeycomb CROW structures. The black dashed line represents a single repeating unit, and the arrows denote three nearest neighbor vectors a 1 = ( 0 , 1 ) , a 2 = ( 3 2 , 1 2 ) , a 3 = ( 3 2 , 1 2 ) .
Fig. 2
Fig. 2 The energy bands E ( k ) corresponding to the Hamiltonian (a) without the Spin-Orbit term, and (b) with the Spin-Orbit coupling term.
Fig. 3
Fig. 3 (a) Schematic diagram of the energy band. (b) Schematic diagram of the projected band. The gapless edge states are shown in the red and blue dots, which corresponds to the upper and lower borders, respectively. The calculation interval of k x is Γ K Γ ( Γ and K are high symmetric points in the first Brillouin zone), and k y is always zero in the calculation.
Fig. 4
Fig. 4 The field distributions at the four high symmetric points for the energy bands.
Fig. 5
Fig. 5 The field distributions of topologically protected edge state and the bulk localized mode. (a) The field distribution of topologically protected edge state, white dashed box represents the input port of CROW structure. (b)Robust edge state propagation with the site ring defect.
Fig. 6
Fig. 6 The field distributions of the collective localization modes. (a) The flat band is activated by inputting the mode light into the lower right port of the structure. (b) The flat band is activated by using the dipole light source (at the end of the white arrows).
Fig. 7
Fig. 7 (a) The field distribution of edge state in an arbitrarily shaped topological cavity. The cavity mode is activated by using the dipole light source (at the end of the white arrows). (b) The field distribution of edge state in a topological beam splitter.

Tables (1)

Tables Icon

Table 1 Parity-eigenvalue patterns at the four high symmetric points for the four different energy bands. All the energy gaps are associated with a nontrivial Z2 index .

Equations (7)

Equations on this page are rendered with MathJax. Learn more.

H 0 = < i j > t i j c i c j
H k = ( 0 e i k · a 2 / 2 0 e i k · a 2 / 2 0 e i k · a 2 / 2 0 e i k · a 1 / 2 0 e i k · a 3 / 2 0 e i k · a 1 / 2 0 e i k · a 1 / 2 0 e i k · a 2 / 2 0 e i k · a 1 / 2 0 e i k · a 3 / 2 0 e i k · a 3 / 2 0 e i k · a 3 / 2 0 )
ε Γ ( k ) ( ϕ 1 ϕ 3 ϕ 5 ) = 2 ( 0 cos k ( a 2 a 1 ) / 2 cos k ( a 2 a 3 ) / 2 cos k ( a 2 a 1 ) / 2 0 cos k ( a 3 a 1 ) / 2 cos k ( a 2 a 3 ) / 2 cos k ( a 3 a 1 ) / 2 0 ) ( ϕ 1 ϕ 3 ϕ 5 ) = H Γ ( k ) ( ϕ 1 ϕ 3 ϕ 5 )
ε K ( k ) ( ϕ 2 ϕ 4 ) = 2 ( 0 v = 1 3 exp ( i k a v ) v = 1 3 exp ( i k a v ) 0 ) ( ϕ 2 ϕ 4 ) = H K ( k ) ( ϕ 2 ϕ 4 )
ε K ( k ) = ( 1 ± 4 A k 3 ) , ε K ( k ) = 2 , ε Γ ( k ) = ± | e x p ( i k a 1 ) + e x p ( i k a 2 ) + e x p ( i k a 3 ) |
h p K = v K ( p x σ 1 + p y σ 2 ) 3 2 α λ s o σ 3 h p Γ = v Γ ( p x S 1 + p y S 2 ) 2 3 λ s o S 3
i = 0 3 m = 1 N ξ 2 m ( Γ i ) = ( 1 ) v N
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.