Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Topological polaritons and photonic magic angles in twisted α-MoO3 bilayers

This article has been updated

Abstract

Twisted two-dimensional bilayer materials exhibit many exotic electronic phenomena. Manipulating the ‘twist angle’ between the two layers enables fine control of the electronic band structure, resulting in magic-angle flat-band superconductivity1,2, the formation of moiré excitons3,4,5,6,7,8 and interlayer magnetism9. However, there are limited demonstrations of such concepts for photons. Here we show how analogous principles, combined with extreme anisotropy, enable control and manipulation of the photonic dispersion of phonon polaritons in van der Waals bilayers. We experimentally observe tunable topological transitions from open (hyperbolic) to closed (elliptical) dispersion contours in bilayers of α-phase molybdenum trioxide (α-MoO3), arising when the rotation between the layers is at a photonic magic twist angle. These transitions are induced by polariton hybridization and are controlled by a topological quantity. At the transitions the bilayer dispersion flattens, exhibiting low-loss tunable polariton canalization and diffractionless propagation with a resolution of less than λ0/40, where λ0 is the free-space wavelength. Our findings extend twistronics10 and moiré physics to nanophotonics and polaritonics, with potential applications in nanoimaging, nanoscale light propagation, energy transfer and quantum physics.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Rotation-induced topological transition of phonon polaritons.
Fig. 2: Experimental observation of topological polaritons.
Fig. 3: Tunable topological transition magic angles.
Fig. 4: Polariton canalization near the topological transition magic angle.

Similar content being viewed by others

Data availability

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Change history

  • 08 July 2020

    The online publication date in the printed version of this article was listed incorrectly as 10 June 2020; the date was correct online.

References

  1. Cao, Y. et al. Unconventional superconductivity in magic-angle graphene superlattices. Nature 556, 43–50 (2018).

    Article  ADS  CAS  PubMed  Google Scholar 

  2. Cao, Y. et al. Strange metal in magic-angle graphene with near Planckian dissipation. Phys. Rev. Lett. 124, 076801 (2020).

    Article  ADS  CAS  PubMed  Google Scholar 

  3. Wu, F., Lovorn, T. & MacDonald, A. H. Topological exciton bands in moiré heterojunctions. Phys. Rev. Lett. 118, 147401 (2017).

    Article  ADS  PubMed  Google Scholar 

  4. Yu, H., Liu, G.-B., Tang, J., Xu, X. & Yao, W. Moiré excitons: from programmable quantum emitter arrays to spin–orbit-coupled artificial lattices. Sci. Adv. 3, e1701696 (2017).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  5. Seyler, K. L. et al. Signatures of moiré-trapped valley excitons in MoSe2/WSe2 heterobilayers. Nature 567, 66–70 (2019).

    Article  ADS  CAS  PubMed  Google Scholar 

  6. Tran, K. et al. Evidence for moiré excitons in van der Waals heterostructures. Nature 567, 71–75 (2019).

    Article  ADS  CAS  PubMed  Google Scholar 

  7. Jin, C. et al. Observation of moiré excitons in WSe2/WS2 heterostructure superlattices. Nature 567, 76–80 (2019).

    Article  ADS  CAS  PubMed  Google Scholar 

  8. Alexeev, E. M. et al. Resonantly hybridized excitons in moiré superlattices in van der Waals heterostructures. Nature 567, 81–86 (2019).

    Article  ADS  CAS  PubMed  Google Scholar 

  9. Chen, W. et al. Direct observation of van der Waals stacking-dependent interlayer magnetism. Science 366, 983–987 (2019).

    Article  ADS  CAS  PubMed  Google Scholar 

  10. Carr, S. et al. Twistronics: manipulating the electronic properties of two-dimensional layered structures through their twist angle. Phys. Rev. B 95, 075420 (2017).

    Article  ADS  Google Scholar 

  11. Jiang, L. et al. Soliton-dependent plasmon reflection at bilayer graphene domain walls. Nat. Mater. 15, 840–844 (2016).

    Article  ADS  CAS  PubMed  Google Scholar 

  12. Sunku, S. S. et al. Photonic crystals for nano-light in moiré graphene superlattices. Science 362, 1153–1156 (2018).

    Article  ADS  MathSciNet  CAS  PubMed  Google Scholar 

  13. Ni, G. X. et al. Soliton superlattices in twisted hexagonal boron nitride. Nat. Commun. 10, 4360 (2019).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  14. Gomez-Diaz, J. S., Tymchenko, M. & Alù, A. Hyperbolic plasmons and topological transitions over uniaxial metasurfaces. Phys. Rev. Lett. 114, 233901 (2015).

    Article  ADS  CAS  PubMed  Google Scholar 

  15. Hu, G., Krasnok, A., Mazor, Y., Qiu, C.-W. & Alù, A. Moiré hyperbolic metasurfaces. Nano Lett. 20, 3217–3224 (2020).

    Article  ADS  CAS  PubMed  Google Scholar 

  16. High, A. A. et al. Visible-frequency hyperbolic metasurface. Nature 522, 192–196 (2015).

    Article  ADS  CAS  PubMed  Google Scholar 

  17. Li, P. et al. Infrared hyperbolic metasurface based on nanostructured van der Waals materials. Science 359, 892–896 (2018).

    Article  ADS  CAS  PubMed  Google Scholar 

  18. Kildishev, A. V., Boltasseva, A. & Shalaev, V. M. Planar photonics with metasurfaces. Science 339, 1232009 (2013).

    Article  CAS  PubMed  Google Scholar 

  19. Liu, Y. M. & Zhang, X. Metasurfaces for manipulating surface plasmons. Appl. Phys. Lett. 103, 141101 (2013).

    Article  ADS  CAS  Google Scholar 

  20. Dai, S. et al. Tunable phonon polaritons in atomically thin van der Waals crystals of boron nitride. Science 343, 1125–1129 (2014).

    Article  ADS  CAS  PubMed  Google Scholar 

  21. Caldwell, J. D. et al. Sub-diffractional volume-confined polaritons in the natural hyperbolic material hexagonal boron nitride. Nat. Commun. 5, 5221 (2014).

    Article  ADS  CAS  PubMed  Google Scholar 

  22. Yoxall, E. et al. Direct observation of ultraslow hyperbolic polariton propagation with negative phase velocity. Nat. Photon. 9, 674–678 (2015).

    Article  ADS  CAS  Google Scholar 

  23. Hu, G., Shen, J., Qiu, C.-W., Alù, A. & Dai, S. Phonon polaritons and hyperbolic response in van der Waals materials. Adv. Opt. Mater. 8, 1901393 (2020).

    Article  CAS  Google Scholar 

  24. Ma, W. et al. In-plane anisotropic and ultra-low-loss polaritons in a natural van der Waals crystal. Nature 562, 557–562 (2018).

    Article  ADS  CAS  PubMed  Google Scholar 

  25. Zheng, Z. et al. A mid-infrared biaxial hyperbolic van der Waals crystal. Sci. Adv. 5, eaav8690 (2019).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  26. Krishnamoorthy, H. N. S., Jacob, Z., Narimanov, E., Kretzschmar, I. & Menon, V. M. Topological transitions in metamaterials. Science 336, 205–209 (2012).

    Article  ADS  MathSciNet  CAS  PubMed  MATH  Google Scholar 

  27. Lifshitz, I. M. Anomalies of electron characteristics of a metal in the high-pressure region. Sov. Phys. JETP 11, 1130 (1960).

    Google Scholar 

  28. Keilmann, F. & Hillenbrand, R. Near-field microscopy by elastic light scattering from a tip. Phil. Trans. R. Soc. A 362, 787–805 (2004).

    Article  ADS  CAS  PubMed  Google Scholar 

  29. Belov, P. A., Simovski, C. R. & Ikonen, P. Canalization of subwavelength images by electromagnetic crystals. Phys. Rev. B 71, 193105 (2005).

    Article  ADS  CAS  Google Scholar 

  30. de Castro, I. A. et al. Molybdenum oxides – from fundamentals to functionality. Adv. Mater. 29, 1701619 (2017).

    Article  CAS  Google Scholar 

  31. Shvets, G., Trendafilov, S., Pendry, J. B. & Sarychev, A. Guiding, focusing, and sensing on the subwavelength scale using metallic wire arrays. Phys. Rev. Lett. 99, 053903 (2007).

    Article  ADS  CAS  PubMed  Google Scholar 

  32. Li, Z.-Y. & Lin, L.-L. Evaluation of lensing in photonic crystal slabs exhibiting negative refraction. Phys. Rev. B 68, 245110 (2003).

    Article  ADS  CAS  Google Scholar 

  33. Woessner, A. et al. Highly confined low-loss plasmons in graphene–boron nitride heterostructures. Nat. Mater. 14, 421–425 (2015).

    Article  ADS  CAS  PubMed  Google Scholar 

  34. Basov, D. N., Fogler, M. M. & García de Abajo, F. J. Polaritons in van der Waals materials. Science 354, aag1992 (2016).

    Article  CAS  PubMed  Google Scholar 

  35. Low, T. et al. Polaritons in layered two-dimensional materials. Nat. Mater. 16, 182–194 (2017).

    Article  ADS  CAS  PubMed  Google Scholar 

  36. Basov, D. N., Averitt, R. D. & Hsieh, D. Towards properties on demand in quantum materials. Nat. Mater. 16, 1077–1088 (2017).

    Article  ADS  CAS  PubMed  Google Scholar 

  37. Correas-Serrano, D., Alù, A. & Sebastian Gomez-Diaz, J. Plasmon canalization and tunneling over anisotropic metasurfaces. Phys. Rev. B 96, 075436 (2017).

    Article  ADS  Google Scholar 

  38. Tilley, D. R. Surface Polaritons: Electromagnetic Waves at Surfaces and Interfaces (Elsevier, 1982).

  39. Tretyakov, S. Analytical Modeling in Applied Electromagnetics (Artech House, 2003).

  40. Zheng, Z. et al. Highly confined and tunable hyperbolic phonon polaritons in van der Waals semiconducting transition metal oxides. Adv. Mater. 30, e1705318 (2018).

    Article  CAS  PubMed  Google Scholar 

  41. Palik, E. D., (ed.) Handbook of Optical Constants of Solids Vol. 3 (Academic, 1998).

Download references

Acknowledgements

We thank Z. Q. Xu for discussions and help with the sample fabrication. This work was performed at the CUNY Advanced Science Research Center (ASRC), National University of Singapore (NUS) and the Melbourne Centre for Nanofabrication (MCN) in the Victorian Node of the Australian National Fabrication Facility (ANFF). G.H., A.K., Y.M. and A.A. acknowledge support from the Air Force Office of Scientific Research (MURI grant no. FA9550-18-1-0379), the Department of Defense under the Vannevar Bush Fellowship program, the Office of Naval Research (grant no. N00014-19-1-2011), the Simons Foundation and the National Science Foundation. Q.O. acknowledges support from ARC Centre of Excellence in Future Low-Energy Electronics Technologies (FLEET). J.W. acknowledges the A*STAR Career Development Award Funding from the Science and Engineering Research Council (grant no. A1820g0087). Z.D. acknowledges the National Natural Science Foundation of China (grant no. 51601131). Q.B. acknowledges support from the Australian Research Council (ARC, FT150100450, CE170100039 and IH150100006). C.-W.Q. acknowledges financial support from A*STAR Pharos Program (grant number 15270 00014, with project number R-263-000-B91-305).

Author information

Authors and Affiliations

Authors

Contributions

G.H. and A.A. conceived the idea. G.H., A.A., Y.M., A.K. and C.-W.Q. developed the theory. G.H. performed the simulations and advised in the experimental designs. Q.O. and Q.B. led the experiments. Q.O. designed the structures, and performed the optical measurements. G.S. and Q.O. fabricated the defects and edges on the sample. Y.W. contributed to material synthesis. G.H., Q.O., J.W., Z.D., Q.Z., Q.B., C.-W.Q. and A.A. analysed the data and all authors discussed the results. G.H., A.A., Q.O., Q.B. and C.-W.Q. wrote the manuscript, with input and comments from all authors. A.A., Q.B. and C.-W.Q. supervised the project.

Corresponding authors

Correspondence to Qiaoliang Bao, Cheng-Wei Qiu or Andrea Alù.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature thanks Zhe Fei, Vincenzo Giannini and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data figures and tables

Extended Data Fig. 1 Properties of a single α-MoO3 layer.

a, Permittivity of α-MoO3. ε11 and ε22 are two permittivity tensor components along the [100] and [001] crystal directions, respectively. The permittivity is obtained by fitting the obtained data (refs. 24,25,40) with the Lorentzian model. b, In-plane open angle (β) of a single layer hyperbolic α-MoO3 flake. c, Isofrequency dispersion of a single α-MoO3 layer (with a thickness of 100 nm) in free space at 925.9 cm−1. The red line is the theoretical dispersion, the dashed black lines are the asymptotic line of the hyperbolic dispersion and the green arrow shows the group velocity direction at large wavevector. The open angle therefore illustrates the behaviour of the asymptotic dispersion line as well as the group velocity directions.

Extended Data Fig. 2 Rotation-induced dispersion engineering of tBL α-MoO3.

a, b, The crossing of two dispersion lines of the top and bottom layer upon anticlockwise (a) and clockwise (b) rotation. The black line is the dispersion of the bottom layer; the yellow, purple, green, blue and red lines correspond to the dispersion of the top layer at twist angles of ±30°, ±45°, ±70°, ±80° and ±90°, respectively. cf, The dispersion of a tBL α-MoO3 sample (d1 = d2 = 100 nm, \({\varepsilon }_{d}=1\)) at twist angles of ±30° (c), ±70° (d), ±80° (e) and ±90° (f). The plotted dispersion lines correspond to the sample of the bottom layer without rotation (solid black lines), the dispersion of the top layer (solid purple lines), the analytical dispersion for tBL α-MoO3 with ds = 1 nm (solid red lines), and the analytical dispersion for tBL α-MoO3 with ds = 0.1 nm (dashed blue lines), ds = 10 nm (dashed green lines) and ds = 0 nm (dashed cyan lines). gj, The dispersion of a tBL α-MoO3 sample (d1 = d2 = 100 nm, ds = 1 nm). The plotted dispersion lines correspond to the sample of the bottom layer without rotation (solid black lines), the dispersion of the top layer (solid purple lines) and the analytical dispersion for tBL α-MoO3 with εd = 1 (solid red lines), εd = 2 (dashed green lines), εd = 3 (dashed blue lines), εd = 4 (dashed cyan lines), εd = 5 (dashed dark yellow lines) and εd = 10 (dashed orange lines).

Extended Data Fig. 3 Numerical field distributions and dispersions for a single layer α-MoO3 at ω = 925.93 cm−1.

aj, The numerical field distribution (Re(Ez); top), the numerically obtained dispersions (FFT(Re(Ez)); contours, bottom) and the analytical dispersions (red dashed lines, bottom), at thicknesses (H) of 40 nm (a), 50 nm (b), 60 nm (c), 80 nm (d), 100 nm (e), 200 nm (f), 300 nm (g), 400 nm (h), 500 nm (i) and 600 nm (j).

Extended Data Fig. 4 Numerical field distributions and dispersions of twisted bilayer α-MoO3 at a fixed thickness of d2 = 200 nm.

ae, The numerical field distribution (Re(Ez); top), the numerically obtained dispersions (FFT(Re(Ez)); contours, bottom) and the analytical dispersions (red dashed lines, bottom) for top-layer thicknesses (d1) of 40 nm (a), 100 nm (b), 200 nm (c), 300 nm (d) and 400 nm (e) at ω = 925.93 cm−1. The dashed light magenta lines, dashed cyan lines and solid red lines are the analytical dispersions of the individual bottom layer, individual top layers and twisted bilayers, respectively. The twist angle is 30° in all cases.

Extended Data Fig. 5 Numerical result for a top layer of fixed thickness of d1 = 200 nm.

ae, The numerical field distribution (Re(Ez); top), the numerically obtained dispersions (FFT(Re(Ez)); contours, bottom) and the analytical dispersions (red dashed lines, bottom) for bottom-layer thicknesses (d2) of 40 nm (a), 100 nm (b), 200 nm (c), 300 nm (d) and 400 nm (e) at ω = 925.93 cm−1. The dashed light magenta lines, dashed cyan lines and solid red lines are the analytical dispersions of individual bottom layers, individual top layers and twisted bilayers, respectively. The twist angle is 30° in all cases.

Extended Data Fig. 6 Mechanism of defect-enabled observation of PhPs.

The metallic AFM tip in the s-SNOM setup could scatter the incident field and launch the highly confined PhPs (referred to as tip-launched PhPs and shown as red/blue solid lines) via near-field light–matter interactions. For case A, the AFM tip is close to the point defect in the sense that the tip-launched PhPs could mostly reach and then be scattered by the point defect (referred to as reflected PhPs and shown as red and blue dashed lines). Thus the wave could be reflected, and strong signals of total PhPs would be expected. For case B, the AFM tip is far from the point defect in the sense that the tip-launched PhPs cannot reach the point defect or the reflected signals cannot be collected by the AFM tip before it is mostly decayed. Therefore, no signal or only weak signals are expected from the reflected PhPs. Note that the optical path should be twice the distance between the tip and point defect, leading to the measured field \({{\rm{e}}}^{-{\rm{i}}{\vec{k}}_{{\rm{P}}{\rm{h}}{\rm{P}}}\cdot 2(\vec{r}-{\vec{r}}_{{\rm{P}}{\rm{D}}})}\). Therefore, when performing the spatial Fourier transform, the obtained spectrum should be of \(2{\vec{k}}_{{\rm{PhP}}}\), that is, the value of the dispersion lines should be multiplied by 2.

Extended Data Fig. 7 Comparison of experimental and numerical results of the field distribution and the dispersion at ω = 903.8 cm−1.

a, Experimentally measured field distribution (left) and dispersion (right) of a single α-MoO3 layer of thickness 125 nm (the same as in Fig. 2e, i). b, Numerical field distribution (left) launched by the dipole and the obtained dispersion (right) of a single α-MoO3 layer of thickness 125 nm. c, Experimentally measured field distribution (left) and the dispersion (right) of the tBL α-MoO3 flake with Δθ = −44°, d1 = d2 = 128 nm (the same as in Fig. 2f, j). d, Numerical field distribution (left) launched by the dipole and the obtained dispersion (right) of the tBL α-MoO3 flake with Δθ = −44°, d1 = d2 = 128 nm. e, Experimentally measured field distribution (left) and dispersion (right) of the tBL α-MoO3 flake with Δθ = 65°, d1 = 120 nm and d2 = 235 nm (the same as in Fig. 2g, k). f, Numerical field distribution (left) launched by the dipole and the obtained dispersion (right) of the tBL α-MoO3 flake with Δθ = 65°, d1 = 120 nm and d2 = 235 nm. g, Experimentally measured field distribution (left) and dispersion (right) of the tBL α-MoO3 flake with Δθ = −77°, d1 = 125 nm, and d2 = 210 nm (the same as in Fig. 2h, l). h, Numerical field distribution (left) launched by the dipole and the obtained dispersion (right) of the tBL α-MoO3 flake with Δθ = −77°, d1 = 125 nm and d2 = 210 nm. In the dispersion plot, the red lines correspond to the analytical dispersions (kx, ky) and the white lines correspond to the analytical dispersion with a factor 2—that is, (2kx, 2ky). The experimentally obtained dispersion fits better for the white-coloured dispersion (2kx, 2ky) owing to doubled optical path via the collection of field by the tip, whereas the numerical one fits better for the dashed red dispersion line (kx, ky) as it is propagating wave.

Extended Data Fig. 8 Observation of the topological transition at ω = 925.9 cm−1.

ai, Optical images (a, d, g), experimentally measured field distributions (b, e, h) and the obtained dispersions (c, f, i) of tBL α-MoO3 flakes. For ac, Δθ = −33°, d1 = 224 nm and d2 = 280 nm. For df, Δθ = 65°, d1 = 120 nm and d2 = 235 nm. For gi, Δθ = −77°, d1 = 125 nm and d2 = 210 nm. These results show that for small twist angles (|Δθ| = 33°, 65°; that is, <72°), the obtained dispersions are hyperbolic, and for large twist angles (|Δθ| = 77° > 72°), the dispersion is elliptical.

Extended Data Fig. 9 The canalization angle at ω = 903.8 cm−1.

a, The near-field images near the point defect in Fig. 4a. b, The cut-line plots of the PhP signal at different propagation distances along the group velocity \({\vec{v}}_{{\rm{g}}}\) from the point defect. c, The extracted FWHM from different line profiles at different propagation lengths, along the group velocity \({\vec{v}}_{{\rm{g}}}\). The reference of zero distance is the black line.

Extended Data Fig. 10 Field canalization and extreme anisotropy at ω = 903.8 cm−1.

a, The field distribution (|Ez|) of a single α-MoO3 flake with a thickness of 100 nm. The permittivity is artificially set as ε11 = −100 + 100i, ε22 = 1.1 and εzz = 8.9. The energy mostly propagates along the x axis, which is caused by a flattened band at the extreme anisotropy. b, The line plot of the field along the dashed blacked line in a. The red dots are the numerical value and the blue line is the fitting. This plot gives the exponential decay \({k}_{{\rm{i}}}^{-1}=0.35\,{\rm{\mu }}{\rm{m}}\). c, The decay length with respect to εi (the imaginary part of ε11), as ε11 = −100 + εi × i, while the other permittivity components do not change.

Supplementary information

Supplementary Information

This file contains Supplementary Discussions 1–4.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hu, G., Ou, Q., Si, G. et al. Topological polaritons and photonic magic angles in twisted α-MoO3 bilayers. Nature 582, 209–213 (2020). https://doi.org/10.1038/s41586-020-2359-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-020-2359-9

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing