Next Article in Journal
Wear Characteristics of WC-Co Cutting Tools Obtained by the U-FAST Method During Particleboard Milling
Previous Article in Journal
Molecular Dynamics Simulation on Orientation-Dependent Mechanical Behaviors of ZnO Single Crystals Under Nanoindentation
Previous Article in Special Issue
Growths of SiC Single Crystals Using the Physical Vapor Transport Method with Crushed CVD-SiC Blocks Under High Vertical Temperature Gradients
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Localized Surface Phonon Polaritons and Infrared Optical Absorption of ScAlN Nanoresonators

1
School of Nuclear Science, Energy and Power Engineering, Shandong University, Jinan 250061, China
2
Optics & Thermal Radiation Research Center, Institute of Frontier and Interdisciplinary, Shandong University, Qingdao 266237, China
*
Authors to whom correspondence should be addressed.
Materials 2025, 18(16), 3906; https://doi.org/10.3390/ma18163906
Submission received: 26 July 2025 / Revised: 12 August 2025 / Accepted: 14 August 2025 / Published: 21 August 2025
(This article belongs to the Special Issue Research Progress of Advanced Crystals: Growth and Doping)

Abstract

Alloying AlN with ScN provides a robust strategy for engineering its intrinsic bandgap, phonons and dielectric functions, and ScAlN alloys have demonstrated great promise in applications including the 5G mobile network, surface acoustic wave devices and nanophotonics. Sc doping has been shown to greatly influence the phonons and infrared dielectric functions of AlN, yet few studies have focused on its influence on surface phonon polaritons, which are crucial to modulating the radiative properties of ScAlN metasurfaces. Herein, we combined first-principles and finite element method (FEM) simulations to fully investigate the effects of Sc incorporation on the phonon dispersion relation, propagation and localization of SPhPs and the modulated radiative properties of ScAlN nanoresonators. As the Sc doping concentration increases, the highest optical phonon frequencies are reduced and are largely directly related to enlarged lattice parameters. Consequently, the coupling strength among incident photons and phonons decreases, which leads to a reduced absorption peak in the infrared dielectric functions. Moreover, the propagation length of the SPhPs in ScAlN is largely reduced, and localized resonance modes gradually disappear at a higher Sc doping concentration. This work provides physical insights into the spectra tuning mechanisms of ScAlN nanoresonators via Sc doping and facilitates their applications in nanophotonic devices.

1. Introduction

Scandium aluminum nitride (ScAlN) exhibits an extraordinary piezoelectric response, spontaneous polarizability and ferroelectricity, which advocate for it as a promising candidate in fields such as high-electron-mobility transistors and ferroelectric and surface acoustic wave devices [1,2,3,4,5]. Owing to its large bandgap, robust bandgap tunability via alloying and good compatibility with complementary metal oxide semiconductor (CMOS) fabrication technology, ScAlN also emerges as an active material for integrated photonics operating from to the UV to the mid-IR spectral range [6,7,8,9]. An in-depth exploitation of the light–matter interactions in ScAlN is relevant to accelerating photonic device design.
Dielectric function, as one of the fundamental thermophysical properties of matter, reflects the strength of light–matter interactions. Exploring the infrared dielectric function of ScAlN can help us understand the infrared phonon modes and their interactions with incident photons, which determine the optical absorption. Deng et al. [10] prepared a ScAlN thin film deposited on sapphire (0001) and measured the Raman and infrared reflectance spectra to extract the infrared phonon modes and investigate the effects of various Sc doping concentrations further. Mock and co-workers [11] grew wurtzite ScxAl1-xN (0 ≤ x ≤ 0.2) thin films using the molecular beam epitaxy method and employed spectroscopic ellipsometry to measure their infrared dielectric functions. The measurements showed that the infrared dielectric function is largely modified and the phonons shift to lower frequencies with increasing incorporation of scandium. With substantial modulation of the infrared dielectric functions through Sc doping, the light–matter interactions, especially in ScAlN nanophotonics, would also differ largely. For pure AlN, the surface phonon polaritons (SPhPs) lying between the Reststrahlen band contribute significantly to engineering the absorption spectra by introducing nanoresonator structures [12]. In GaN/AlN heterostructures, the optical phonon modes are greatly modulated and a new Reststrahlen band emerges, which provides the opportunity to tune the infrared properties [13,14]. Also, tunable spectral behavior can be achieved through hybridization of the doping-dependent surface phonon–plasmon resonances in AlN/SiC heterostructures integrated with graphene [15]. Recently, SPhPs and infrared spectral modulation in polar crystals, including GaN [16], SiC [17] and SrTiO3 [18], have been fully explored. Though alloying AlN with Sc at a low concentration retains its polar nature, the behaviors of SPhPs in ScAlN are subtly different from those in AlN, mainly due to modification of the infrared dielectric function and phonon modes. Thus, it is of great necessity to explore the effects of Sc doping on the light–matter interactions in ScAlN-based nanophotonics.
In this work, we systematically investigate the influence of Sc doping on the infrared phonon modes, dielectric function and SPhPs of ScAlN and the localized resonances and absorption spectra of corresponding nanoresonators using first-principles and finite element method (FEM) simulations. The first-principles calculations indicate that the highest optical phonon frequencies are reduced by increasing Sc doping. As a consequence, the interactions among the incident photons and phonons become weakened, and the maximum value of the infrared dielectric function decreases, which reduces the propagation length of the SPhPs further. FEM simulations show that the resonance modes gradually disappear at a high Sc doping concentration, and the radiative properties of ScAlN nanoresonators can be robustly modulated by the structural geometry.

2. The Computational Methodology

2.1. The Lattice Structure and Phonons

Under ambient conditions, AlN possesses a non-centrosymmetric wurtzite (space group, P63mc) structure, while ScN exhibits a meta-stable layered hexagonal structure [19]. When doping AlN with a low Sc concentration (typically up to 35%), ScAlN retains the wurtzite lattice structure but is a disordered solid solution, as shown in Figure 1a. The special quasi-random structure (SQS) method [20] was adopted to simulate a disordered ScAlN supercell containing 16 atoms, and density-functional-theory-based first-principles simulations were performed using the Vienna ab initio simulation package (VASP) [21,22]. The Perdew–Burke–Ernzerhof (PBE) functional [23,24] was used to describe the exchange-correlation potential for Al, Sc and N atoms, and the plane-wave cutoff energy was set as 800 eV [23,24]. First-principles calculations demonstrate that the lattice constant a of ScAlN gradually increases with an increasing Sc concentration, while c varies slightly for a Sc doping concentration up to 40% [25]. For instance, the optimized lattice constants a and c of Sc0.05Al0.95N are 3.152 Å and 5.024 Å, while they are 3.221 Å and 5.047 Å for Sc0.2Al0.8N, which are in good agreement with previous experimental data [26,27]. With the relaxed lattice structure, the infrared phonon modes of ScAlN were calculated using the Phonopy package [28]. The finite displacement method was used to compute the interatomic force constants, and a large supercell containing 128 atoms was chosen.

2.2. Dielectric Function and Finite Element Simulations

With varying Sc doping concentrations, the intrinsic infrared photon–phonon interactions in ScAlN change significantly and thus influence the electromagnetic interactions of light with ScAlN-based nanoresonators. The infrared dielectric functions of ScAlN thin films grown through molecular beam epitaxy were measured using spectroscopic ellipsometry [11] and chosen as the thermophysical input for finite element method (FEM) simulations of ScAlN-based nanoresonators. In Figure 1b, a schematic diagram of a representative ScAlN nanoresonator is presented, and its geometry is defined by the period (P), diameter (D) and height (H). A transverse magnetic (TM) wave was launched within the xz plane at an incident angle of 0° (normal incidence kz ) in the simulation. Floquet periodic boundary conditions were imposed in the X and Y directions for the FEM simulation to simulate the nanoresonator array structure, while perfectly matched layers with a thickness of a 1/3 of a wavelength were applied in the Z direction. The global convergence grid was set to 1/10 of a wavelength, and the grid of the nanoresonator was refined to 0.1 μm.

3. Results and Discussion

3.1. The Phonon Dispersion Relation of ScAlN

Sc incorporation alters the lattice structure of AlN and thus affect its lattice dynamic properties. Figure 2 compares the calculated phonon dispersion relations of AlN and Sc0.125Al0.875N. The non-existence of negative phonon modes in Sc0.125Al0.875N indicates its dynamical stability. With doped Sc, it is found that the highest frequency for the optical phonon decreases accordingly. For instance, the highest phonon frequency for AlN is 882.65 cm−1 but is reduced to 853.14 cm−1 for Sc0.125Al0.875N. Such a decrease in the phonon frequency is directly related to an enlarged lattice parameter. Equally, along the high-symmetric Γ-A path in the first Brillouin zone, the low-lying longitudinal acoustic phonons are softened by the doped Sc atoms.

3.2. The Dielectric Function and Surface Phonon Polaritons

The infrared dielectric function arises from intrinsic photon–phonon interactions, and phonon modes altered by Sc doping influence the dielectric function further. In Figure 3, the anisotropic complex dielectric functions of ScAlN with varying Sc concentrations under ambient conditions are presented and extracted from Ref. [11]. Owing to its wurtzite lattice structure, the ordinary (⊥c) dielectric function of ScAlN is largely different from that in the extraordinary (∥c) direction. As the doped Sc concentration increases, it is observed that the complex dielectric function is significantly altered. For instance, the maximum of the ordinary ε2 (see Figure 3a) is greatly reduced at a Sc doping concentration up to 20%; the width is largely broadened; and the location of the absorption peak shifts to a lower photon energy.
To quantify the effect of doping on the anisotropic dielectric function of ScAlN, the classical LO-TO oscillator model is used to fit the data from [29,30]:
ε ( ω ) = ε ω 2 i Γ L O ω ω L O 2 ω 2 i Γ T O ω ω T O 2 ,
where ωTO and ωLO are the frequencies of the longitudinal and transverse optical phonons, respectively; ε is the high-frequency dielectric constant; and Γ is the broadened width. The fitted ωTO and ωLO values for the ordinary and extraordinary dielectric functions of ScAlN are presented in Table 1. For the ordinary dielectric function, both ωTO and ωLO decrease as the Sc doping concentration increases, while ωLO is largely reduced, but ωTO remains constant for the extraordinary direction.
Moreover, for incident photons with frequencies lying between ωTO and ωLO , it is found that the real part ε1 of the dielectric function is negative. These incident photons will propagate along the surface of ScAlN instead of being transmitted into it and couple with the lattice vibrations, thus forming surface phonon polaritons (SPhPs) [31,32,33,34]. The spectral range lying between ωTO and ωLO is defined as the Reststrahlen band (RB), where the incident photons are highly reflected by the ScAlN’s surface. With an increasing Sc doping concentration, the RB’s width is reduced, indicating that fewer incident photons can excite SPhPs. The intrinsic propagation length of the SPhPs propagating along the ScAlN’s surface is defined as Λ = (2Im ksp)−1 [35,36], where ksp is a complex wave vector defined by ksp = ω/c(ε/(1 + ε))1/2. ε is the complex dielectric function, and this relation applies to the single air–ScAlN interface. Figure 4a shows the SPhP propagation length in the ordinary direction at varying Sc doping concentrations. The propagation length initially decreases with an increasing wavenumber, exhibits a dip near 850 cm−1 and then increases at higher wavenumbers. Below 800 cm−1, increasing the Sc doping concentration results in a reduced propagation length. A similar trend is observed for the extraordinary direction, as shown in Figure 4b, although the overall propagation lengths are shorter compared to those in the conventional case. Equally, for Sc0.1Al0.9N, the highest propagation length can reach over 1000 μm, closely related to the long lifetime of optical phonons.

3.3. Localized SPhPs in ScAlN Nanoresonators

With such a large propagation length of the SPhPs in ScAlN, the introduction of nanostructuring onto the surface will induce localized surface phonon resonances and thus greatly modulate the optical properties. We calculated the absorption spectra of a ScxAl1-xN (x = 0, 0.04, 0.07, 0.10 and 0.20) nanoresonator with P = 7 μm, D = 3 μm and H = 1.5 μm. In Figure 5, it is obvious that pure AlN exhibits three absorption modes: mode A (777 cm−1), mode B (806 cm−1) and mode C (867 cm−1). By analyzing the electric field distribution further, we find that modes A and C are excited by the mirror-symmetric in-plane electric field, and mode B is the higher-order hybrid mode, as shown in Figure 6. However, resonance modes B and C gradually disappear with an increasing Sc doping concentration. For Sc0.04Al0.96N, only modes A and C are observed, while only one mode, mode A, remains when the Sc incorporation increases to 10%, and no mode is observed at 20% doping. Moreover, from Figure 5, it is obvious that upon Sc doping, the phonons in the AlN films are shifted to lower frequencies, and simultaneously, the peak widths increase drastically.
To display the relationship between the two absorption peaks (A and C in Figure 5) and the structural parameters intuitively, taking the Sc0.07Al0.93N nanoresonator as an example, the normalized |E|-field distributions of the two resonant wavenumbers (762 cm−1 and 849 cm−1) are calculated. From Figure 7a, it is observed that the |E|-field of absorption peak A (762 cm−1) is mainly concentrated around the nanoresonator, especially on the topside surface, while the second absorption peak C (849 cm−1) is mainly concentrated at the four corners of the nanoresonator. Additionally, peak C shows a much lower field intensity, indicating that the localization of this mode is relatively weak. And changes in the diameter and height of the ScAlN nanoresonator will affect the absorption spectra.
Furthermore, we studied the absorption characteristics of a Sc0.07Al0.93N nanoresonator with varying geometrical parameters (4 μm ≤ P ≤ 8 μm, 2 μm ≤ D ≤ 6 μm and 1 μm ≤ H ≤ 5 μm). Figure 8a shows the absorption spectra pertinent to different periods P with D = 3 μm and h = 1.5 μm. The position of mode A (TD1)’s absorption peak blueshifts almost linearly (Δk ≈ 20 cm−1/ΔP = 0.6 μm) as P increases. But almost no change is observed for the two modes when P ≥ 6.0 μm. It is obvious that a new mode D (TD1) appears at k ≈ 785 cm−1 and is coupled with mode A (TD1) when P = 4.6 μm, which leads to greater spreading of mode A. However, after this point, the peak widths decrease drastically with increasing period. Figure 8b,c display the absorption spectra as the diameter varies with P = 7 μm and H = 1.5 μm and the height with P = 7 μm and D = 3 μm varies. Mode A exhibits an obvious redshift and blueshift with an increasing height and diameter, especially for height; when H = 1.4 μm, mode C appears and is coupled with mode A at H = 2.2 μm, while the position of mode C’s absorption peak shifts slightly with an increasing diameter. As the diameter increases, the peak widths of the two modes decrease gradually but increase drastically with increasing height. These correspond to the field distribution analysis in Figure 7.
By gradually changing the period, diameter and height, we find that the resonant amplitude is gradually enhanced with increasing structural parameters until it reaches the maximum value. As the height increases, the two modes progressively couple into broadband absorption modes, with this effect becoming particularly pronounced when H ≥ 2.2 μm.

4. Conclusions

To summarize, we have investigated the effects of the Sc doping concentration and geometric morphology on the infrared phonons, SPhPs and absorption spectra of ScAlN using first-principles calculations and FEM simulations. The incorporation of Sc into AlN will enlarge the lattice parameters and decrease the highest optical phonon frequencies. The absorption peak of the infrared dielectric function shifts to a lower wavenumber, and its amplitude decreases at a high Sc doping concentration, mainly due to weakened photon–phonon interactions. ScAlN can host SPhPs at low Sc doping concentrations, and their propagation length is over 1000 μm, providing great avenues for modulating the localized resonance modes and absorption spectra of nanoresonators. The geometric morphology also determines the localized SPhPs and absorption spectra. These findings could help exploit potential application scenarios for ScAlN metasurfaces as mid-IR photodetectors and thermal emitters.

Author Contributions

Conceptualization, H.Z. and J.-Y.Y.; methodology, H.Z. and T.C.; software, H.Z. and T.C.; validation, H.Z., T.C. and X.D.; formal analysis, H.Z.; investigation, M.L.; resources, M.L.; data curation, H.Z. and T.C.; writing—original draft preparation, H.Z. and X.D.; writing—review and editing, L.L. and J.-Y.Y.; visualization, X.D.; supervision, L.L. and J.-Y.Y.; project administration, M.L.; funding acquisition, J.-Y.Y. All authors have read and agreed to the published version of the manuscript.

Funding

Shandong University Outstanding Young Scholar (82363194).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding authors.

Acknowledgments

J.-Y.Y. is grateful for the support from Shandong University Outstanding Young Scholar (82363194).

Conflicts of Interest

The authors have no conflicts of interest to disclose.

References

  1. Zou, Y.; Gao, C.; Zhou, J.; Liu, Y.; Xu, Q.; Qu, Y.; Liu, W.; Soon, J.B.W.; Cai, Y.; Sun, C. Aluminum scandium nitride thin-film bulk acoustic resonators for 5G wideband applications. Microsyst. Nanoeng. 2022, 8, 124. [Google Scholar] [CrossRef]
  2. Song, Y.; Perez, C.; Esteves, G.; Lundh, J.S.; Saltonstall, C.B.; Beechem, T.E.; Yang, J.I.; Ferri, K.; Brown, J.E.; Tang, Z.; et al. Thermal Conductivity of Aluminum Scandium Nitride for 5G Mobile Applications and Beyond. ACS Appl. Mater. Interfaces 2021, 13, 19031–19041. [Google Scholar] [CrossRef]
  3. Yang, G.; Wang, H.; Mu, S.; Xie, H.; Wang, T.; He, C.; Shen, M.; Liu, M.; Van de Walle, C.G.; Tang, H.X. Unveiling the Pockels coefficient of ferroelectric nitride ScAlN. Nat. Commun. 2024, 15, 9538. [Google Scholar] [CrossRef]
  4. Wang, P.; Wang, D.; Mondal, S.; Mi, Z. Ferroelectric N-polar ScAlN/GaN heterostructures grown by molecular beam epitaxy. Appl. Phys. Lett. 2022, 121, 023501. [Google Scholar] [CrossRef]
  5. Yang, D.-P.; Tang, X.-G.; Sun, Q.-J.; Chen, J.-Y.; Jiang, Y.-P.; Zhang, D.; Dong, H.-F. Emerging ferroelectric materials ScAlN: Applications and prospects in memristors. Mater. Horiz. 2024, 11, 2802–2819. [Google Scholar] [CrossRef]
  6. Liu, X.; Wang, D.; Kim, K.-H.; Katti, K.; Zheng, J.; Musavigharavi, P.; Miao, J.; Stach, E.A.; Olsson, R.H.; Jariwala, D. Post-CMOS compatible aluminum scandium nitride/2D channel ferroelectric field-effect-transistor memory. Nano Lett. 2021, 21, 3753–3761. [Google Scholar] [CrossRef]
  7. Gopakumar, G.; Abdin, Z.U.; Kumar, R.; Dzuba, B.; Nguyen, T.; Manfra, M.J.; Malis, O. Conduction-band engineering of polar nitride semiconductors with wurtzite ScAlN for near-infrared photonic devices. J. Appl. Phys. 2024, 135, 165701. [Google Scholar] [CrossRef]
  8. Zhu, S.; Zhong, Q.; Li, N.; Hu, T.; Dong, Y.; Xu, Z.; Zhou, Y.; Fu, Y.H.; Singh, N. Integrated ScAlN photonic circuits on silicon substrate. In Proceedings of the 2020 Conference on Lasers and Electro-Optics (CLEO), San Jose, CA, USA, 10–15 May 2020; pp. 1–2. [Google Scholar]
  9. Moram, M.; Zhang, S. ScGaN and ScAlN: Emerging nitride materials. J. Mater. Chem. A 2014, 2, 6042–6050. [Google Scholar] [CrossRef]
  10. Deng, R.; Jiang, K.; Gall, D. Optical phonon modes in Al1−xScxN. J. Appl. Phys. 2014, 115, 013506. [Google Scholar] [CrossRef]
  11. Mock, A.L.; Jacobs, A.G.; Jin, E.N.; Hardy, M.T.; Tadjer, M.J. Long-wavelength dielectric properties and infrared active optical phonon modes of molecular beam epitaxy ScxAl1−xN determined by infrared spectroscopic ellipsometry. Appl. Phys. Lett. 2020, 117, 232107. [Google Scholar] [CrossRef]
  12. Zhao, H.; Cheng, T.; Wang, G.; Zhang, L.; Ma, M.; Liu, J.; Yang, J.-Y.; Liu, L. Temperature-dependent infrared dielectric function and localized surface phonon polaritons in polar AlN. Opt. Express 2025, 33, 19810–19819. [Google Scholar] [CrossRef]
  13. Reyes-González, J.E.; Dellby, N.; Plotkin-Swing, B.; Wang, P.; Pandey, A.; Mi, Z.; Lagos, M.J. Unveiling phonon dispersion behavior of AlN/GaN heterostructures using EELS. Microsc. Microanal. 2023, 29, 354–355. [Google Scholar] [CrossRef]
  14. Ratchford, D.C.; Winta, C.J.; Chatzakis, I.; Ellis, C.T.; Passler, N.C.; Winterstein, J.; Dev, P.; Razdolski, I.; Matson, J.R.; Nolen, J.R.; et al. Controlling the infrared dielectric function through atomic-scale heterostructures. ACS Nano 2019, 13, 6730–6741. [Google Scholar] [CrossRef]
  15. Zhang, Y.; Gao, X.; Xia, H.; Mei, J.; Cui, Z.; Lai, J.; Chen, C. Tunable infrared surface phonon–plasmon coupling in graphene-integrated polar semiconductor heterostructure. APL Photonics 2023, 8, 126104. [Google Scholar] [CrossRef]
  16. Sharmila; Pandey, A.; Singh, N.; Garg, P.; Raman, R.; Kumar, P.S. Investigation of surface and interface phonon polariton in nitrogen-implanted GaN epi-layers. J. Appl. Phys. 2025, 137, 125302. [Google Scholar] [CrossRef]
  17. Barnett, J.; Wendland, D.; Lewin, M.; Wirth, K.G.; Heßler, A.; Taubner, T. Investigation of low-confinement surface phonon polariton launching on SiC and SrTiO3 using scanning near-field optical microscopy. Appl. Phys. Lett. 2022, 120, 211107. [Google Scholar] [CrossRef]
  18. Cheng, T.; Zhao, H.; Liu, L.; Yang, J.-Y. Temperature-dependent in-plane thermal conductivity of SrTiO3 membranes enhanced by surface phonon polaritons. Int. J. Therm. Sci. 2024, 202, 109077. [Google Scholar] [CrossRef]
  19. Ambacher, O.; Mihalic, S.; Yassine, M.; Yassine, A.; Afshar, N.; Christian, B. Review: Structural, elastic, and thermodynamic properties of cubic and hexagonal ScxAl1−xN crystals. J. Appl. Phys. 2023, 134, 160702. [Google Scholar] [CrossRef]
  20. Jiang, C. First-principles study of ternary bcc alloys using special quasi-random structures. Acta Mater. 2009, 57, 4716–4726. [Google Scholar] [CrossRef]
  21. Hohenberg, P.; Kohn, W. Inhomogeneous electron gas. Phys. Rev. 1964, 136, B864–B871. [Google Scholar] [CrossRef]
  22. Kohn, W.; Sham, L.J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, A1133–A1138. [Google Scholar] [CrossRef]
  23. Perdew, J.P. Density-functional approximation for the correlation energy of the inhomogeneous electron gas. Phys. Rev. B 1986, 33, 8822. [Google Scholar] [CrossRef]
  24. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  25. Dong, H.; Li, Z.; Sun, B.; Zhou, Y.; Liu, L.; Yang, J.-Y. Thermal transport in disordered wurtzite ScAlN alloys using machine learning interatomic potentials. Mater. Today Commun. 2024, 39, 109213. [Google Scholar] [CrossRef]
  26. Furuta, K.; Hirata, K.; Anggraini, S.A.; Akiyama, M.; Uehara, M.; Yamada, H. First-principles calculations of spontaneous polarization in ScAlN. J. Appl. Phys. 2021, 130, 024104. [Google Scholar] [CrossRef]
  27. Dinh, D.V.; Lähnemann, J.; Geelhaar, L.; Brandt, O. Lattice parameters of ScxAl1−xN layers grown on GaN(0001) by plasma-assisted molecular beam epitaxy. Appl. Phys. Lett. 2023, 122, 152103. [Google Scholar] [CrossRef]
  28. Togo, A.; Chaput, L.; Tanaka, I.; Hug, G. First-principles phonon calculations of thermal expansion in Ti3SiC2, Ti3AlC2, and Ti3GeC2. Phys. Rev. B 2010, 81, 174301. [Google Scholar] [CrossRef]
  29. Lyddane, R.H.; Sachs, R.G.; Teller, E. On the polar vibrations of alkali halides. Phys. Rev. 1941, 59, 673–676. [Google Scholar] [CrossRef]
  30. Schubert, M.; Tiwald, T.; Herzinger, C. Infrared dielectric anisotropy and phonon modes of sapphire. Phys. Rev. B 2000, 61, 8187. [Google Scholar] [CrossRef]
  31. Huber, A.J.; Deutsch, B.; Novotny, L.; Hillenbrand, R. Focusing of surface phonon polaritons. Appl. Phys. Lett. 2008, 92, 203104. [Google Scholar] [CrossRef]
  32. Caldwell, J.D.; Vurgaftman, I.; Tischler, J.G.; Glembocki, O.J.; Owrutsky, J.C.; Reinecke, T.L. Atomic-scale photonic hybrids for mid-infrared and terahertz nanophotonics. Nat. Nanotechnol. 2016, 11, 9–15. [Google Scholar] [CrossRef]
  33. Huber, A.; Ocelic, N.; Kazantsev, D.; Hillenbrand, R. Near-field imaging of mid-infrared surface phonon polariton propagation. Appl. Phys. Lett. 2005, 87, 081103. [Google Scholar] [CrossRef]
  34. Cheng, T.; Zhao, H.; Kong, W.; Liu, L.; Yang, J.-Y. Leveraging surface phonon polaritons for enhanced Q-factor of mid-infrared BaTiO3 nanoresonators. J. Mater. Chem. C 2024, 12, 12266–12274. [Google Scholar] [CrossRef]
  35. Kalfagiannis, N.; Stoner, J.L.; Hillier, J.; Vangelidis, I.; Lidorikis, E. Mid- to far-infrared sensing: SrTiO3, a novel optical material. J. Mater. Chem. C 2019, 7, 7851–7857. [Google Scholar] [CrossRef]
  36. Yang, J.-Y.; Cheng, T.; Fei, T.; Zhang, C.; Liu, L. Temperature-induced surface phonon polaritons dissipation in perovskite SrTiO3. Opt. Lett. 2021, 46, 4244–4247. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Crystal structure transition of AlN and ScAlN with Sc doping and (b) schematic and parametric diagram of ScAlN nanoresonator structure defined by period (P), diameter (D) and height (H).
Figure 1. (a) Crystal structure transition of AlN and ScAlN with Sc doping and (b) schematic and parametric diagram of ScAlN nanoresonator structure defined by period (P), diameter (D) and height (H).
Materials 18 03906 g001
Figure 2. A comparison of the theoretically calculated phonon dispersion relation for (a) AlN and (b) Sc0.125Al0.875N from first-principles calculations. The high-symmetry-point coordinates are Γ (0,0,0), M (0.5,0,0), K (0.3333,0.3333,0) and A (0,0,0.5).
Figure 2. A comparison of the theoretically calculated phonon dispersion relation for (a) AlN and (b) Sc0.125Al0.875N from first-principles calculations. The high-symmetry-point coordinates are Γ (0,0,0), M (0.5,0,0), K (0.3333,0.3333,0) and A (0,0,0.5).
Materials 18 03906 g002
Figure 3. The (a,b) ordinary (⊥c) and (c,d) extraordinary (∥c) complex dielectric functions of ScAlN thin films under ambient conditions measured by Ref. [11].
Figure 3. The (a,b) ordinary (⊥c) and (c,d) extraordinary (∥c) complex dielectric functions of ScAlN thin films under ambient conditions measured by Ref. [11].
Materials 18 03906 g003
Figure 4. The SPhP propagation length of ScAlN in the (a) ordinary and (b) extraordinary directions with varying Sc doping concentrations.
Figure 4. The SPhP propagation length of ScAlN in the (a) ordinary and (b) extraordinary directions with varying Sc doping concentrations.
Materials 18 03906 g004
Figure 5. The calculated absorption spectra of the ScxAl1-xN (0 ≤ x ≤ 0.2) nanoresonator at a normal incident angle for the TM wave.
Figure 5. The calculated absorption spectra of the ScxAl1-xN (0 ≤ x ≤ 0.2) nanoresonator at a normal incident angle for the TM wave.
Materials 18 03906 g005
Figure 6. The charge density of (a) AlN at mode A (777 cm−1), mode B (806 cm−1) and mode C (867 cm−1); (b) Sc0.04Al0.96N at mode A (765 cm−1) and mode C (853 cm−1); (c) Sc0.07Al0.93N at mode A (762 cm−1) and mode C (849 cm−1); and (d) Sc0.10Al0.90N at mode A (762 cm−1) and (e) the schematic charge configurations of the three modes.
Figure 6. The charge density of (a) AlN at mode A (777 cm−1), mode B (806 cm−1) and mode C (867 cm−1); (b) Sc0.04Al0.96N at mode A (765 cm−1) and mode C (853 cm−1); (c) Sc0.07Al0.93N at mode A (762 cm−1) and mode C (849 cm−1); and (d) Sc0.10Al0.90N at mode A (762 cm−1) and (e) the schematic charge configurations of the three modes.
Materials 18 03906 g006
Figure 7. The normalized |E|-field distributions for the two absorption peaks at (a) A 762 cm−1 and (b) C 849 cm−1.
Figure 7. The normalized |E|-field distributions for the two absorption peaks at (a) A 762 cm−1 and (b) C 849 cm−1.
Materials 18 03906 g007
Figure 8. Tailoring the mode A (TD1) and mode C (TD2) resonance modes of a Sc0.07Al0.93N nanoresonator with varying structural parameters of (a) P, (b) D and (c) H at a normal incident angle for a TM electromagnetic wave.
Figure 8. Tailoring the mode A (TD1) and mode C (TD2) resonance modes of a Sc0.07Al0.93N nanoresonator with varying structural parameters of (a) P, (b) D and (c) H at a normal incident angle for a TM electromagnetic wave.
Materials 18 03906 g008
Table 1. The fitted frequency of LO-TO phonons and the corresponding Reststrahlen band (RB) width.
Table 1. The fitted frequency of LO-TO phonons and the corresponding Reststrahlen band (RB) width.
ScxAl1-xNOrdinary Dielectric FunctionExtraordinary Dielectric Function
ωTO [cm−1]ωLO [cm−1]RB Width [cm−1]ωTO [cm−1]ωLO [cm−1]RB Width [cm−1]
AlN671.34910.50239.16615.00879.40264.40
Sc0.04Al0.96N660.85892.40231.55615.00868.70253.70
Sc0.07Al0.93N660.52885.60225.08615.00862.00247.00
Sc0.10Al0.90N651.00880.40229.40615.00854.00239.00
Sc0.20Al0.80N636.80862.70225.90635.00770.00235.00
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhao, H.; Cheng, T.; Duan, X.; Lv, M.; Yang, J.-Y.; Liu, L. Localized Surface Phonon Polaritons and Infrared Optical Absorption of ScAlN Nanoresonators. Materials 2025, 18, 3906. https://doi.org/10.3390/ma18163906

AMA Style

Zhao H, Cheng T, Duan X, Lv M, Yang J-Y, Liu L. Localized Surface Phonon Polaritons and Infrared Optical Absorption of ScAlN Nanoresonators. Materials. 2025; 18(16):3906. https://doi.org/10.3390/ma18163906

Chicago/Turabian Style

Zhao, Huanhuan, Tao Cheng, Xinlei Duan, Mingxin Lv, Jia-Yue Yang, and Linhua Liu. 2025. "Localized Surface Phonon Polaritons and Infrared Optical Absorption of ScAlN Nanoresonators" Materials 18, no. 16: 3906. https://doi.org/10.3390/ma18163906

APA Style

Zhao, H., Cheng, T., Duan, X., Lv, M., Yang, J.-Y., & Liu, L. (2025). Localized Surface Phonon Polaritons and Infrared Optical Absorption of ScAlN Nanoresonators. Materials, 18(16), 3906. https://doi.org/10.3390/ma18163906

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop