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

Optical nanoimaging of highly-confined phonon polaritons in atomically-thin nanoribbons of α-MoO3

Open Access Open Access

Abstract

Phonon polaritons (PhPs), collective modes hybridizing photons with lattice vibrations in polar insulators, enable nanoscale control of light. In recent years, the exploration of in-plane anisotropic PhPs has yielded new levels of confinement and directional manipulation of nano-light. However, the investigation of in-plane anisotropic PhPs at the atomic layer limit is still elusive. Here, we report the optical nanoimaging of highly-confined phonon polaritons in atomically-thin nanoribbons of α-MoO3 (5 atomic layers). We show that narrow α-MoO3 nanoribbons as thin as a few atomic layers can support anisotropic PhPs modes with a high confinement ratio (∼133 times smaller wavelength than that of light). The anisotropic PhPs interference fringe patterns in atomic layers are tunable depending on the PhP wavelength via changing the illumination frequency. Moreover, spatial control over the PhPs interference patterns is also achieved by varying the nanostructures’ shape or nanoribbon width of atomically-thin α-MoO3. Our work may serve as an empirical reference point for other anisotropic PhPs that approach the thickness limit and pave the way for applications such as atomically integrated nano-photonics and sensing.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Recent research has revealed the excellent abilities of PhPs including the high optic field confinement (large momentum) [1,2], ultraslow group velocities [3,4], long lifetimes (low propagation loss) [5,6] and extreme polariton anisotropy (hyperbolic dispersion) [7,8]. These characteristics make PhPs ideal for super-resolution imaging [912], field-enhanced molecular sensing [13,14], sub-diffraction waveguiding [10,15], infrared metasurfaces [7,16,17], thermal energy transfer [18] and many others [19,20]. In recent years, α-MoO3 has been proven to be a natural biaxial van der Waals (vdW) crystal supporting hyperbolic PhPs in the mid-infrared region [2128], which can achieve a stronger confinement than isotropic PhPs and show great potential to directional control of light. At the same time, by precisely adjusting the thickness of polar vdW layers, e.g., via mechanical exfoliation technique, the PhPs modes are controlled by the number of vdW layers and the field could be further confined when scaling down the thickness to atomic scale. The highly confined in-plane isotropic PhPs in atomically thin layers or even monolayer [5,29] is demonstrated and their real-space propagation are presented. Additionally, the PhPs achieved in atomically thin slabs also lays the foundation for highly-confined acoustic modes, as has been theoretically demonstrated [30,31]. However, the picture of in-plane anisotropic PhPs when approaching thickness limit is still unrevealed.

Here we present the optical nanoimaging of highly-confined phonon polaritons in α-MoO3 nanoribbons with atomically thin thicknesses (6.5–7 nm: 5 atomic layers). The high-quality nanoribbon structure of α-MoO3 strongly squeezes the electromagnetic field of PhPs into a nanoscale cavity between the two long edges of nanoribbon. Consequently, we are able to detect enhanced intensity of the electromagnetic field on the α-MoO3 nanoribbon and observe significant PhP dispersion, particularly in the elliptical band.

2. Result and discussion

To explore the polaritonic response of ultrathin α-MoO3 nanoribbons, we performed nanoimaging measurement using scattering-type scanning near-field optical microscopy (s-SNOM, Fig. 1(a)). In such a system, the metallic atomic force microscope (AFM) tip of s-SNOM in the vicinity of the sample is illuminated by a p-polarized infrared light of frequency ω. Acting as an infrared antenna, the tip focuses the incident field precisely at its apex, creating a nanoscale infrared spot locally probing material properties and exciting polaritons. Simultaneously, the electromagnetic radiation scattered by the tip is captured along with the topography of α-MoO3 nanoribbons, resulting in near-field images with nanoscale resolution.

 figure: Fig. 1.

Fig. 1. Ultra-confined PhPs in ultrathin α-MoO3 nanoribbons on SiO2 (300 nm)/Si substrate. (a) Schematic of real-space nanoimaging of PhPs launched in α-MoO3 nanoribbons with s-SNOM. (b) Calculated momentum dispersion of PhPs in the elliptic (top panel) and hyperbolic (bottom panel) regimes along [100] (x-direction) with different thicknesses (grey: 50 nm and red: 6.5 nm) of α-MoO3. Yellow triangles show the experimental data points extracted from s-SNOM measurements. (c) Calculated thickness-dependence of kx/k0 (kx: PhP complex wavevector, k0: free-space wavevector) in the elliptic (red line, ω = 1000 cm−1) and hyperbolic (black line, ω = 870 cm−1) regimes. (d) Near-field amplitude image S3 of α-MoO3 nanoribbons measured at ω = 870 cm−1 (top panel) and AFM image (bottom panel). The black line in AFM image shows the height profile of an ultrathin α-MoO3 nanoribbon which is about 6.5 nm, corresponding to five atomic layers.

Download Full Size | PDF

α-MoO3 has a layered structure in which covalent-bonded ac planes (in-plane) are arranged along the b-axis (out-of-plane) direction through vdW forces. Note that the b lattice constant of an α-MoO3 unit cell is 1.4 nm [32], therefore, our AFM result indicates that the 6.5-nm α-MoO3 has 5 layers. The rectangular shape of α-MoO3 crystals is predominantly determined by their anisotropic crystal structure, with the long axis corresponding to the [001] direction [21]. Previous works have indicated that ribbon-type α-MoO3 flakes oriented along the [001] direction as favored upon exfoliation because the energy release along the [001] direction is much greater than that along the [100] direction [33,34].

It is noted that the confinement of PhPs in α-MoO3 increases with decreased thickness, allowing for significant modifications to the dispersion properties via the thickness tailoring. We calculated the momentum dispersion relation in the [100] direction for 50-nm and 6.5-nm thick α-MoO3 (Fig. 1(b)). The dark stripes represent the allowed fundamental modes in the upper (elliptic regime) and the lower (hyperbolic regime) Reststrahlen bands. The dispersions show that the PhP wavelength can be easily varied by the excitation frequency and scales significantly with the thickness of α-MoO3. It is noted that when the thickness of α-MoO3 is at atomic scale (5 layers), the confinement factor could reach beyond 133, the experimental results are indicated as triangles which will be discussed later. Figure 1(c) shows the thickness dependence of the confinement factor. When the thickness is larger than 10 nm, the confinement factor changes slowly with thickness. However, when the thickness drops below 10 nm, the confinement factor increases sharply as the thickness decreases. Figure 1(d) presents an overall view of α-MoO3 nanoribbons’ near-field amplitude images measured at 870 cm−1 in hyperbolic bands of α-MoO3. There are only two bright fringes occur at the boundaries of ribbons parallel to the [100] direction and some slight fringes in the interior of the ribbons (more clearly seen by the line profiles in Fig. 3). This suggests an in-plane anisotropic interference pattern of PhPs in the atomically-thin α-MoO3, demonstrating the unique in-plane hyperbolic polaritons that propagate in one direction while being cut off in the perpendicular direction near the thickness limit.

To analyze the anisotropic PhPs in α-MoO3 nanoribbons in more detail, we further conducted frequency-dependent near-field nanoimaging experiments on the sample. When the free-space light field is initially coupled onto the α-MoO3 via a metallic tip, the excited PhP waves propagate in the ultra-thin nanostructure until reaching a physical boundary where strong reflection occurs. The reflected and incident waves then interfere with each other, resulting in the formation of characteristic interference fringes (Fig. 2(a)). More complex interference patterns can be generated by increasing the number of boundaries, which enables the back-and-forth reflections of the PhP waves in all different directions (Fig. 2(b)). Thus, the interference patterns can be modulated by the wavelength of PhP which is related to the frequency of excitation light, or by precisely modifying the geometry of nanostructures. The PhP waves within the α-MoO3 nanostructures were the sum of the tip-launched PhP waves and those reflected from the boundaries [35]:

$$\begin{array}{{c}} {\psi = {{\tilde{\psi }}_{\textrm{PhP},0}} + \mathop \sum \limits_j {{\tilde{\psi }}_{\textrm{PhP},j}}} \end{array}$$
$$\begin{array}{{c}} {{{\tilde{\psi }}_{\textrm{PhP},j}} = {R_j}{{\tilde{\psi }}_{\textrm{PhP},0}}\exp\left\{ { - \left( {\frac{{4\pi }}{{{\lambda_{\textrm{PhP}}}}}} \right){L_j}({{\gamma_{\textrm{PhP}}} + i} )} \right\}} \end{array}$$
where ${\tilde{\psi }_{\textrm{PhP},0}}$ and ${\tilde{\psi }_{\textrm{PhP},j}}$ are the tip-launched PhP wave and the waves reflected by the boundaries. Parameters ${R_j} = {R_0}\textrm{exp}({i\mathrm{\Delta }\varphi } )$, ${\lambda _{\textrm{PhP}}}$, ${\gamma _{\textrm{PhP}}}$ and ${L_j}$ describe the reflection coefficient ($\varphi $ denotes the phase shift), the wavelength of PhP, the damping rate and the distance between α-MoO3 edge and AFM tip, respectively. The interference patterns are associated with $|\psi |$.

 figure: Fig. 2.

Fig. 2. Sketches of the reflection and interference of PhP waves in α-MoO3 with one (a) or two boundaries (b). (c, d) Elliptical-band PhPs in ultrathin α-MoO3 nanoribbon. (c) Near-field amplitude images S3 of α-MoO3 nanoribbon measured at ω = 997 cm−1, 1000 cm−1, 1002 cm−1, 1004 cm−1 and 1005 cm−1, respectively, normalized to the amplitude of the surrounding SiO2. (d) Extracted experimental amplitude profiles (black lines) compared with calculation (red dashed lines) along the blue dashed line marked in image (c).

Download Full Size | PDF

Figure 2(c, d) displays the optical near-field images of an ultrathin α-MoO3 nanoribbon in the elliptical band using s-SNOM. The frequency-dependent variation of interference fringes in the cavity of the nanoribbon is clearly observed. Note that the α-MoO3 nanoribbon is nearly transparent at 1005 cm−1 since the third-order demodulated signal S3 was almost equal inside and outside the cavity, which indicates the fading of PhPs. Simultaneously, owing to the high confinement of PhPs inside the cavity from 997 cm−1 to 1004 cm−1, we observed an intracavity signal intensity S3 far above the SiO2 background. To extract the polariton wavevector kp from measured interference fringes, we use Eq. (1) to fit the extracted line profiles, which have been used in previous investigations [35]. The subtle differences between the fitting and experiment data come from the oblique illumination [36]. Our fitting results (see enlarged image at 1000 cm−1 in Fig. S3) indicate that the wavelength of PhPs in the 5-layer α-MoO3 is as small as 75 nm at 10 µm (1000 cm−1), implying a high confinement factor 133 that is consistent with the calculated dispersion relation (Fig. 1(b)). In addition, the positions of peaks and valleys near the two boundaries can be seen shifting from 1002 cm−1 to 997 cm−1 in Fig. 2 due to the decreasing λp, pointing at a higher confinement at the smaller frequency. A small deviation is observed between the theoretical dispersion and the experiment data (obtained from Fig. 2) for 6.5-nm thick α-MoO3 nanoribbon. The reason might be that the optical and electronic properties gradually deviate from the bulk response when the thickness of material is reduced below 4–6 vdW layers which are commonly found in MoS2 [37,38].

The experimental near-field images in the elliptical band provide an indication that the ultrathin α-MoO3 nanoribbon support a high confinement of PhPs, consistent with the calculated dispersion. We further investigated the PhPs of ultrathin α-MoO3 nanoribbons in the hyperbolic band (Fig. 3). In this frequency range (860 cm−1 to 890 cm−1), the in-plane polariton wavelength λp becomes much smaller (see also Fig. 1(b)) and one observes an expansion of low-intensity near-field signals in the center of nanoribbon cavity, similar to the darker regions found in near-field images from 1002 cm−1 to 997 cm−1 in Fig. 2(c, d). In addition, a rather faint bright line in the center of the cavity can be observed at 860 cm−1 and 870 cm−1, corresponding to the small bump at 250 nm in the amplitude profile, see Fig. 3(b). The phenomenon in the hyperbolic band is far less pronounced than that in the elliptical band. This can be understood by considering that the fast group velocity of hyperbolic band leads to less drastic dispersion changes compared to elliptic band with slow group velocities, and the PhP modes of α-MoO3 supported in the hyperbolic band was reported to have a relatively weak Ez component than that in the elliptical band [27].

 figure: Fig. 3.

Fig. 3. Hyperbolic-band PhPs in ultrathin α-MoO3 nanoribbon. (a) Near-field amplitude images S3 of the same α-MoO3 nanoribbon as in Fig. 2, measured at ω = 890 cm−1, 882 cm−1, 870 cm−1, 860 cm−1. (b) Extracted experimental amplitude profiles along the blue dashed line marked in image (a).

Download Full Size | PDF

We next proceed to study the cavity-width tailored PhPs in two wedge-shaped α-MoO3 nanoribbons (Fig. 4(a)). We investigated the orientation-dependent dispersion of PhPs, considering that α-MoO3 is a typical anisotropic vdW material and PhPs propagate anisotropically along the surface. Under the approximation of high momentum where kp/k0 ≫ 1, the in-plane dispersion relation of the PhP propagating as a function of frequency ω can be resolved [23]:

$$\begin{array}{{c}} {{k_\textrm{p}}d = \psi \left( {\textrm{arctan}\left( {\frac{{\psi {\varepsilon_\textrm{a}}}}{{{\varepsilon_z}}}} \right) + \textrm{arctan}\left( {\frac{{\psi {\varepsilon_\textrm{s}}}}{{{\varepsilon_z}}}} \right) + \pi l} \right),\; \; l = 0,\; 1,\; 2 \ldots } \end{array}$$

 figure: Fig. 4.

Fig. 4. Elliptical-band PhPs of wedge-shaped α-MoO3 nanoribbons (thickness: 7 nm). (a) Near-field amplitude images S3 of two wedge-shaped nanoribbons at measured at 1002 cm−1. (b) Calculated momentum dispersion. The isofrequency contour at 1002 cm−1 is highlighted with red line. (c) Projection of the calculated momentum dispersion on the kx-ky plane. Gray dashed lines: isofrequency contours at different frequencies from 1000 cm−1 (outside) to 1005 cm−1 (inside) with 1 cm−1 step size. Red line: isofrequency contour at 1002 cm−1. Red dotted line: a circle for comparison. (d) Extracted amplitude profiles (solid lines) along the corresponding dashed lines marked in (a), in comparison with calculation (red dashed lines).

Download Full Size | PDF

Here, $\psi = i\sqrt {{\varepsilon _z}/({{\varepsilon_x}\textrm{co}{\textrm{s}^2}\theta + {\varepsilon_y}\textrm{si}{\textrm{n}^2}\theta } )} $, εx,y,z is the dielectric permittivity of α-MoO3 along the corresponding crystallographic axis, d is the α-MoO3 flake thickness, εa and εs are the dielectric permittivities of the surrounding air and the substrate, θ is the angle between the x axis ([100]) and the in-plane component vector, respectively.

In Fig. 4(b, c), we plot the isofrequency curves for the fundamental modes in 7-nm thick α-MoO3. When k0 increases, for instance at 1002 cm−1, the isofrequency contour (red line) becomes more circular (indicated by the red dotted line) instead of being elliptical. Under this condition, the angle between the optical axis and the edge has less impact on the anisotropy of PhP propagation. We extracted three amplitude profiles at three different positions from the near-field optical image in Fig. 4(a) and the results are shown in Fig. 4(d). Here, when the cavity width expands (the nanoribbon widths for L1, L2 and L3 positions extracted from AFM images are 297 nm, 205 nm and 128 nm, respectively), an increase in the number of peaks can be determined, in good agreement with the calculated results (red dashed lines). In addition, there was no indication of a strong decay in the intensity of interference fringes when the cavity width, i.e., the propagation distance, is varied.

3. Conclusions

In summary, we have successfully demonstrated that the atomically-thin layers of α-MoO3 nanoribbons enable a high electric field confinement and the active control over the polariton propagation properties by changing the illumination frequency. Especially, the interference fringe patterns can be significantly changed in the elliptical band. The nanoribbons naturally provide a physical boundary for the strong interference of phonon polaritons and the atomically thin thickness of α-MoO3 offers a desired field confinement for strong light−matter interactions. The observation and analysis of in-pane anisotropic PhPs propagation in atomically thin layers may provide an empirical reference point for lattice dynamics models of vdW few-layers and lay the foundation for the atomically integrated design of different vdW materials.

Funding

Fundamental Research Funds for the Central Universities (2022JYCXJJ009); National Natural Science Foundation of China (62075070); National Key Research and Development Program of China (2021YFA1201500); Natural Science Foundation of Hubei Province (2022CFA053); China Postdoctoral Science Foundation (2021M701298); Innovation Fund of WNLO.

Acknowledgments

We thank the Analytical and Testing Center of HUST for help with the measurements

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Supplemental document

See Supplement 1 for supporting content.

References

1. P. Li, X. Yang, T. W. Mass, J. Hanss, M. Lewin, A. K. Michel, M. Wuttig, and T. Taubner, “Reversible optical switching of highly confined phonon-polaritons with an ultrathin phase-change material,” Nat. Mater. 15(8), 870–875 (2016). [CrossRef]  

2. J. D. Caldwell, A. V. Kretinin, Y. Chen, V. Giannini, M. M. Fogler, Y. Francescato, C. T. Ellis, J. G. Tischler, C. R. Woods, A. J. Giles, M. Hong, K. Watanabe, T. Taniguchi, S. A. Maier, and K. S. Novoselov, “Sub-diffractional volume-confined polaritons in the natural hyperbolic material hexagonal boron nitride,” Nat. Commun. 5(1), 5221 (2014). [CrossRef]  

3. E. Yoxall, M. Schnell, A. Y. Nikitin, O. Txoperena, A. Woessner, M. B. Lundeberg, F. Casanova, L. E. Hueso, F. H. L. Koppens, and R. Hillenbrand, “Direct observation of ultraslow hyperbolic polariton propagation with negative phase velocity,” Nat. Photonics 9(10), 674–678 (2015). [CrossRef]  

4. K. L. Tsakmakidis, O. Hess, R. W. Boyd, and X. Zhang, “Ultraslow waves on the nanoscale,” Science 358(6361), eaan5196 (2017). [CrossRef]  

5. S. Dai, Z. Fei, Q. Ma, A. S. Rodin, M. Wagner, A. S. McLeod, M. K. Liu, W. Gannett, W. Regan, K. Watanabe, T. Taniguchi, M. Thiemens, G. Dominguez, A. H. Castro Neto, A. Zettl, F. Keilmann, P. Jarillo-Herrero, M. M. Fogler, and D. N. Basov, “Tunable phonon polaritons in atomically thin van der Waals crystals of boron nitride,” Science 343(6175), 1125–1129 (2014). [CrossRef]  

6. A. J. Giles, S. Dai, I. Vurgaftman, T. Hoffman, S. Liu, L. Lindsay, C. T. Ellis, N. Assefa, I. Chatzakis, T. L. Reinecke, J. G. Tischler, M. M. Fogler, J. H. Edgar, D. N. Basov, and J. D. Caldwell, “Ultralow-loss polaritons in isotopically pure boron nitride,” Nat. Mater. 17(2), 134–139 (2018). [CrossRef]  

7. P. Li, I. Dolado, F. J. Alfaro-Mozaz, F. Casanova, L. E. Hueso, S. Liu, J. H. Edgar, A. Y. Nikitin, S. Velez, and R. Hillenbrand, “Infrared hyperbolic metasurface based on nanostructured van der Waals materials,” Science 359(6378), 892–896 (2018). [CrossRef]  

8. Q. Zhang, G. Hu, W. Ma, P. Li, A. Krasnok, R. Hillenbrand, A. Alu, and C. W. Qiu, “Interface nano-optics with van der Waals polaritons,” Nature 597(7875), 187–195 (2021). [CrossRef]  

9. P. Li, I. Dolado, F. J. Alfaro-Mozaz, A. Y. Nikitin, F. Casanova, L. E. Hueso, S. Velez, and R. Hillenbrand, “Optical Nanoimaging of Hyperbolic Surface Polaritons at the Edges of van der Waals Materials,” Nano Lett. 17(1), 228–235 (2017). [CrossRef]  

10. P. Li, M. Lewin, A. V. Kretinin, J. D. Caldwell, K. S. Novoselov, T. Taniguchi, K. Watanabe, F. Gaussmann, and T. Taubner, “Hyperbolic phonon-polaritons in boron nitride for near-field optical imaging and focusing,” Nat. Commun. 6(1), 7507 (2015). [CrossRef]  

11. S. Dai, Q. Ma, T. Andersen, A. S. McLeod, Z. Fei, M. K. Liu, M. Wagner, K. Watanabe, T. Taniguchi, M. Thiemens, F. Keilmann, P. Jarillo-Herrero, M. M. Fogler, and D. N. Basov, “Subdiffractional focusing and guiding of polaritonic rays in a natural hyperbolic material,” Nat. Commun. 6(1), 6963 (2015). [CrossRef]  

12. A. J. Sternbach, S. H. Chae, S. Latini, A. A. Rikhter, Y. Shao, B. Li, D. Rhodes, B. Kim, P. J. Schuck, X. Xu, X. Y. Zhu, R. D. Averitt, J. Hone, M. M. Fogler, A. Rubio, and D. N. Basov, “Programmable hyperbolic polaritons in van der Waals semiconductors,” Science 371(6529), 617–620 (2021). [CrossRef]  

13. M. Autore, P. Li, I. Dolado, F. J. Alfaro-Mozaz, R. Esteban, A. Atxabal, F. Casanova, L. E. Hueso, P. Alonso-Gonzalez, J. Aizpurua, A. Y. Nikitin, S. Velez, and R. Hillenbrand, “Boron nitride nanoresonators for phonon-enhanced molecular vibrational spectroscopy at the strong coupling limit,” Light: Sci. Appl. 7(4), 17172 (2017). [CrossRef]  

14. S. Foteinopoulou, G. C. R. Devarapu, G. S. Subramania, S. Krishna, and D. Wasserman, “Phonon-polaritonics: enabling powerful capabilities for infrared photonics,” Nanophotonics 8(12), 2129–2175 (2019). [CrossRef]  

15. I. Dolado, F. J. Alfaro-Mozaz, P. Li, E. Nikulina, A. Bylinkin, S. Liu, J. H. Edgar, F. Casanova, L. E. Hueso, P. Alonso-Gonzalez, S. Velez, A. Y. Nikitin, and R. Hillenbrand, “Nanoscale Guiding of Infrared Light with Hyperbolic Volume and Surface Polaritons in van der Waals Material Ribbons,” Adv. Mater. 32(9), 1906530 (2020). [CrossRef]  

16. G. Hu, Q. Ou, G. Si, Y. Wu, J. Wu, Z. Dai, A. Krasnok, Y. Mazor, Q. Zhang, Q. Bao, C. W. Qiu, and A. Alu, “Topological polaritons and photonic magic angles in twisted alpha-MoO3 bilayers,” Nature 582(7811), 209–213 (2020). [CrossRef]  

17. M. Chen, X. Lin, T. H. Dinh, Z. Zheng, J. Shen, Q. Ma, H. Chen, P. Jarillo-Herrero, and S. Dai, “Configurable phonon polaritons in twisted alpha-MoO3,” Nat. Mater. 19(12), 1307–1311 (2020). [CrossRef]  

18. K. J. Tielrooij, N. C. H. Hesp, A. Principi, et al., “Out-of-plane heat transfer in van der Waals stacks through electron-hyperbolic phonon coupling,” Nat. Nanotechnol. 13(1), 41–46 (2018). [CrossRef]  

19. L. Shen, X. Lin, M. Y. Shalaginov, T. Low, X. Zhang, B. Zhang, and H. Chen, “Broadband enhancement of on-chip single-photon extraction via tilted hyperbolic metamaterials,” Appl. Phys. Rev. 7(2), 021403 (2020). [CrossRef]  

20. Z. Jacob, “Nanophotonics: Hyperbolic phonon-polaritons,” Nat. Mater. 13(12), 1081–1083 (2014). [CrossRef]  

21. W. Ma, P. Alonso-Gonzalez, S. Li, A. Y. Nikitin, J. Yuan, J. Martin-Sanchez, J. Taboada-Gutierrez, I. Amenabar, P. Li, S. Velez, C. Tollan, Z. Dai, Y. Zhang, S. Sriram, K. Kalantar-Zadeh, S. T. Lee, R. Hillenbrand, and Q. Bao, “In-plane anisotropic and ultra-low-loss polaritons in a natural van der Waals crystal,” Nature 562(7728), 557–562 (2018). [CrossRef]  

22. Y. Wu, Q. Ou, Y. Yin, Y. Li, W. Ma, W. Yu, G. Liu, X. Cui, X. Bao, J. Duan, G. Alvarez-Perez, Z. Dai, B. Shabbir, N. Medhekar, X. Li, C. M. Li, P. Alonso-Gonzalez, and Q. Bao, “Chemical switching of low-loss phonon polaritons in alpha-MoO3 by hydrogen intercalation,” Nat. Commun. 11(1), 2646 (2020). [CrossRef]  

23. J. Duan, G. Alvarez-Perez, A. I. F. Tresguerres-Mata, J. Taboada-Gutierrez, K. V. Voronin, A. Bylinkin, B. Chang, S. Xiao, S. Liu, J. H. Edgar, J. I. Martin, V. S. Volkov, R. Hillenbrand, J. Martin-Sanchez, A. Y. Nikitin, and P. Alonso-Gonzalez, “Planar refraction and lensing of highly confined polaritons in anisotropic media,” Nat. Commun. 12(1), 4325 (2021). [CrossRef]  

24. Z. Zheng, F. Sun, W. Huang, J. Jiang, R. Zhan, Y. Ke, H. Chen, and S. Deng, “Phonon Polaritons in Twisted Double-Layers of Hyperbolic van der Waals Crystals,” Nano Lett. 20(7), 5301–5308 (2020). [CrossRef]  

25. G. Alvarez-Perez, T. G. Folland, I. Errea, J. Taboada-Gutierrez, J. Duan, J. Martin-Sanchez, A. I. F. Tresguerres-Mata, J. R. Matson, A. Bylinkin, M. He, W. Ma, Q. Bao, J. I. Martin, J. D. Caldwell, A. Y. Nikitin, and P. Alonso-Gonzalez, “Infrared Permittivity of the Biaxial van der Waals Semiconductor alpha-MoO3 from Near- and Far-Field Correlative Studies,” Adv. Mater. 32(29), 1908176 (2020). [CrossRef]  

26. Z. Zheng, J. Chen, Y. Wang, X. Wang, X. Chen, P. Liu, J. Xu, W. Xie, H. Chen, S. Deng, and N. Xu, “Highly Confined and Tunable Hyperbolic Phonon Polaritons in Van Der Waals Semiconducting Transition Metal Oxides,” Adv. Mater. 30(13), 1705318 (2018). [CrossRef]  

27. S. J. Yu, Y. Jiang, J. A. Roberts, M. A. Huber, H. Yao, X. Shi, H. A. Bechtel, S. N. Gilbert Corder, T. F. Heinz, X. Zheng, and J. A. Fan, “Ultrahigh-Quality Infrared Polaritonic Resonators Based on Bottom-Up-Synthesized van der Waals Nanoribbons,” ACS Nano 16(2), 3027–3035 (2022). [CrossRef]  

28. Y. Zhao, J. Chen, M. Xue, R. Chen, S. Jia, J. Chen, L. Bao, H. J. Gao, and J. Chen, “Ultralow-Loss Phonon Polaritons in the Isotope-Enriched alpha-MoO3,” Nano Lett. 22(24), 10208–10215 (2022). [CrossRef]  

29. S. Dai, W. Fang, N. Rivera, Y. Stehle, B. Y. Jiang, J. Shen, R. Y. Tay, C. J. Ciccarino, Q. Ma, D. Rodan-Legrain, P. Jarillo-Herrero, E. H. T. Teo, M. M. Fogler, P. Narang, J. Kong, and D. N. Basov, “Phonon Polaritons in Monolayers of Hexagonal Boron Nitride,” Adv. Mater. 31(37), 1806603 (2019). [CrossRef]  

30. Z. Yuan, R. Chen, P. Li, A. Y. Nikitin, R. Hillenbrand, and X. Zhang, “Extremely Confined Acoustic Phonon Polaritons in Monolayer-hBN/Metal Heterostructures for Strong Light–Matter Interactions,” ACS Photonics 7(9), 2610–2617 (2020). [CrossRef]  

31. W. Lyu, H. Teng, C. Wu, X. Zhang, X. Guo, X. Yang, and Q. Dai, “Anisotropic acoustic phonon polariton-enhanced infrared spectroscopy for single molecule detection,” Nanoscale 13(29), 12720–12726 (2021). [CrossRef]  

32. L. Cai, C. J. McClellan, A. L. Koh, H. Li, E. Yalon, E. Pop, and X. Zheng, “Rapid Flame Synthesis of Atomically Thin MoO3 down to Monolayer Thickness for Effective Hole Doping of WSe2,” Nano Lett. 17(6), 3854–3861 (2017). [CrossRef]  

33. X.-L. Li, J.-F. Liu, and Y.-D. Li, “Low-temperature synthesis of large-scale single-crystal molybdenum trioxide (MoO3) nanobelts,” Appl. Phys. Lett. 81(25), 4832–4834 (2002). [CrossRef]  

34. L. Cheng, M. Shao, X. Wang, and H. J. C. A. E. J. Hu, “Single-crystalline molybdenum trioxide nanoribbons: photocatalytic, photoconductive, and electrochemical properties,” Chem. - Eur. J. 15(10), 2310–2316 (2009). [CrossRef]  

35. Z. B. Zheng, J. T. Li, T. Ma, H. L. Fang, W. C. Ren, J. Chen, J. C. She, Y. Zhang, F. Liu, H. J. Chen, S. Z. Deng, and N. S. Xu, “Tailoring of electromagnetic field localizations by two-dimensional graphene nanostructures,” Light: Sci. Appl. 6(10), e17057 (2017). [CrossRef]  

36. W. Ma, G. Hu, D. Hu, R. Chen, T. Sun, X. Zhang, Q. Dai, Y. Zeng, A. Alu, C. W. Qiu, and P. Li, “Ghost hyperbolic surface polaritons in bulk anisotropic crystals,” Nature 596(7872), 362–366 (2021). [CrossRef]  

37. Y. L. Huang, Y. Chen, W. Zhang, S. Y. Quek, C. H. Chen, L. J. Li, W. T. Hsu, W. H. Chang, Y. J. Zheng, W. Chen, and A. T. Wee, “Bandgap tunability at single-layer molybdenum disulphide grain boundaries,” Nat. Commun. 6(1), 6298 (2015). [CrossRef]  

38. R. Ganatra and Q. Zhang, “Few-Layer MoS2: A Promising Layered Semiconductor,” ACS Nano 8(5), 4074–4099 (2014). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Numerical simulation and supplemental experiment results

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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 (4)

Fig. 1.
Fig. 1. Ultra-confined PhPs in ultrathin α-MoO3 nanoribbons on SiO2 (300 nm)/Si substrate. (a) Schematic of real-space nanoimaging of PhPs launched in α-MoO3 nanoribbons with s-SNOM. (b) Calculated momentum dispersion of PhPs in the elliptic (top panel) and hyperbolic (bottom panel) regimes along [100] (x-direction) with different thicknesses (grey: 50 nm and red: 6.5 nm) of α-MoO3. Yellow triangles show the experimental data points extracted from s-SNOM measurements. (c) Calculated thickness-dependence of kx/k0 (kx: PhP complex wavevector, k0: free-space wavevector) in the elliptic (red line, ω = 1000 cm−1) and hyperbolic (black line, ω = 870 cm−1) regimes. (d) Near-field amplitude image S3 of α-MoO3 nanoribbons measured at ω = 870 cm−1 (top panel) and AFM image (bottom panel). The black line in AFM image shows the height profile of an ultrathin α-MoO3 nanoribbon which is about 6.5 nm, corresponding to five atomic layers.
Fig. 2.
Fig. 2. Sketches of the reflection and interference of PhP waves in α-MoO3 with one (a) or two boundaries (b). (c, d) Elliptical-band PhPs in ultrathin α-MoO3 nanoribbon. (c) Near-field amplitude images S3 of α-MoO3 nanoribbon measured at ω = 997 cm−1, 1000 cm−1, 1002 cm−1, 1004 cm−1 and 1005 cm−1, respectively, normalized to the amplitude of the surrounding SiO2. (d) Extracted experimental amplitude profiles (black lines) compared with calculation (red dashed lines) along the blue dashed line marked in image (c).
Fig. 3.
Fig. 3. Hyperbolic-band PhPs in ultrathin α-MoO3 nanoribbon. (a) Near-field amplitude images S3 of the same α-MoO3 nanoribbon as in Fig. 2, measured at ω = 890 cm−1, 882 cm−1, 870 cm−1, 860 cm−1. (b) Extracted experimental amplitude profiles along the blue dashed line marked in image (a).
Fig. 4.
Fig. 4. Elliptical-band PhPs of wedge-shaped α-MoO3 nanoribbons (thickness: 7 nm). (a) Near-field amplitude images S3 of two wedge-shaped nanoribbons at measured at 1002 cm−1. (b) Calculated momentum dispersion. The isofrequency contour at 1002 cm−1 is highlighted with red line. (c) Projection of the calculated momentum dispersion on the kx-ky plane. Gray dashed lines: isofrequency contours at different frequencies from 1000 cm−1 (outside) to 1005 cm−1 (inside) with 1 cm−1 step size. Red line: isofrequency contour at 1002 cm−1. Red dotted line: a circle for comparison. (d) Extracted amplitude profiles (solid lines) along the corresponding dashed lines marked in (a), in comparison with calculation (red dashed lines).

Equations (3)

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

ψ = ψ ~ PhP , 0 + j ψ ~ PhP , j
ψ ~ PhP , j = R j ψ ~ PhP , 0 exp { ( 4 π λ PhP ) L j ( γ PhP + i ) }
k p d = ψ ( arctan ( ψ ε a ε z ) + arctan ( ψ ε s ε z ) + π l ) , l = 0 , 1 , 2
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.