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

Aluminum scandium nitride on 8-inch Si wafers: material characterization and photonic device demonstration

Open Access Open Access

Abstract

The anisotropic optical properties of aluminum scandium nitride (Al1−xScxN) thin films for both ordinary and extraordinary light are investigated. A quantitative analysis of the band structures of the wurtzite Al1−xScxN is carried out. In addition, Al1−xScxN photonic waveguides and bends are fabricated on 8-inch Si substrates. With x = 0.087 and 0.181, the light propagation losses are 5.98 ± 0.11 dB/cm and 8.23 ± 0.39 dB/cm, and the 90° bending losses are 0.05 dB/turn and 0.08 dB/turn at 1550 nm wavelength, respectively.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Aluminum nitride (AlN) has a large bandgap (Eg) of 6.2 eV and a broad transparency window spanning from ultraviolet to mid-infrared [12]. The non-centrosymmetric crystal structure of AlN contributes to its piezoelectric properties (d33 = 5.53 pm/V and d31 = −2.65 pm/V) as well as its strong second-order nonlinear susceptibility (χ(2) = 4.7 pm/V) [35]. The second-order nonlinear optical property enables applications such as second harmonic generation (SHG) and parametric down-conversion [6]. Furthermore, its thermal stability and compatibility with the complementary metal oxide semiconductor (CMOS) process enable large scale, low cost and highly reliable wafer-level manufacturing of integrated photonic devices.

The doping of scandium (Sc) into AlN alters its band structure and optical properties, exhibiting a tunable Eg regarding various Sc concentrations [79]. In addition, Sc doping increases the lattice asymmetry of AlN, potentially enhances the piezoelectric coefficients and second-order optical nonlinearity such as optical birefringence [10]. When the doping concentration is up to 43%, the piezoelectric coefficient d33 is four times higher than AlN films, and the piezoelectric coefficient d31 is also increased significantly. Al1−xScxN has been conventionally used in micro-electro-mechanical systems (MEMS) and radio-frequency (RF) devices due to its outstanding piezoelectric properties. The unique properties of Al1−xScxN confer its significant potential for photonic applications in fields such as sensing, communication and information processing. However, the application of Al1−xScxN in integrated photonics remains largely unexplored mainly due to two reasons: the limited research on optical properties of the material and the incomplete photonic device library.

The exploration of the optical properties of Al1−xScxN holds immense importance in the realm of integrated photonics [11]. Although the optical properties such as Eg, refractive indices (n) and extinction coefficients (k) of Al1−xScxN have been reported, to the best of our knowledge, its anisotropic properties of those parameters have not been extensively studied yet [9]. The optical anisotropy of Al1−xScxN exhibits a pronounced effect on near-band-edge transitions and excitonic effects. Obtaining in-depth knowledge on the optical anisotropy provides valuable guidance for the design and optimization of photonic devices. Besides, as for the Al1−xScxN photonic device library, there are few devices reported in recent decade [1213]. The optical waveguides and waveguide bends are essential components of photonic integrated circuits. With a relatively high propagation loss of 9 dB/cm at x = 0.108 reported by Zhu et al. at λ = 1550 nm [13], the research on Al1−xScxN waveguides and bends is still in its infancy.

In this paper, we present the investigation of the anisotropic properties of Eg, n and k for Al1−xScxN at various Sc concentrations. The band structure of wurtzite Al1−xScxN is analyzed quantitatively. In addition, Al1−xScxN photonic waveguides and bends are designed according to the material characterization results. Finally, the photonic devices are fabricated on an 8-inch Si substrate using CMOS-compatible fabrication platform. For x = 0.087 and 0.181, the lowest propagation losses at λ = 1550 nm are extracted to be 5.98 ± 0.11 dB/cm and 8.23 ± 0.39 dB/cm and the 90° bending losses with 50 μm bend radius are 0.05 dB/turn and 0.08 dB/turn, respectively.

2. Optical properties of Al1−xScxN

The 400 nm-thick Al1−xScxN film was deposited on a 3 μm-thick silicon dioxide (SiO2) layer which had been grown on an 8-inch Si (100) wafer. The Al0.913Sc0.087N and Al0.819Sc0.181N films were deposited from AlSc targets with nominal Sc contents of 9.6 at.% and 20 at.%, respectively, by an RF magnetron sputtering system. In the X-ray diffraction (XRD) characterization, the peak of Al1−xScxN (002) exhibits a relatively narrow full width at half maximum (FWHM) of 1.84°. In addition, there is no other characteristic peak of Al1−xScxN appearing within the angular range of the 2θ scan. The presence of a single (002) diffraction peak typically indicates a consistent orientation of the crystallites in the sample towards the (001) crystal plane. This indicates that the film has a good c-axis orientation which is perpendicular to the substrate [14].

Spectroscopic ellipsometry (SE) is a powerful optical measurement technique that relies on characterizing changes in the polarization state of linearly polarized light reflected from a thin film sample [1517]. This non-destructive method plays a crucial role in the characterization of various materials and devices. The SE spectra amplitude ratio φ and phase difference Δ were measured to describe the changes in the polarization state of linearly polarized light after reflection from the Al1−xScxN films [18]. The parameters φ and Δ satisfy the equation [19]:

$$\rho = \tan \varphi \cdot {e^{i\Delta }} = {r_p}/r{}_s,$$
where rp and rs are defined as the complex reflection coefficients of parallel (P) and perpendicular (S) polarized light, respectively. The electric fields of the incident plane waves P and S to the plane of incidence in the finite-element simulations are defined with unit amplitudes.

The SE spectra of Al1−xScxN films with different Sc concentrations were measured by a variable-angle spectral ellipsometer at room temperature. The optical properties were investigated by scanning the photon energy (hv) from 1.24 eV to 5.905 eV at the incidence angle θ = 60°, 65°and 70°. It is necessary to model the Al1−xScxN film as a uniaxial anisotropic material to accurately fit the SE data [2021]. The SE spectra were fitted using the anisotropic dispersion equations in the completeEASE software by Mueller matrix measurement. The n and k for extraordinary and ordinary lights were determined by minimizing the mean square error, with the results plotted in Figs. 1(a) and 1(b), respectively. The measured VASE data are presented in the Fig. S1 of the Supplement 1. The minimized mean square error (MSE) values for x = 0, 0.087, 0.181 are 21.1, 17.6, and 20, respectively.

 figure: Fig. 1.

Fig. 1. The dispersions of (a) no, ne and (b) ko, ke for Al1−xScxN films at photon energy ranges from 1.24 to 5.905 eV. The (c) Δn and (d) Δk plotted as a function of hv for Al1−xScxN films.

Download Full Size | PDF

The electric field (E) of propagating light in wurtzite Al1−xScxN can be separated into two components: parallel to c-axis (Ec) and perpendicular to c-axis (Ec) which are commonly known as the extraordinary and ordinary configurations, respectively [20,22]. The n and k of ordinary light are represented as no and ko, while those of extraordinary light are denoted by ne and ke, respectively. To accurately characterize the anisotropy of the Al1−xScxN films, the differences of the n and k are represented by birefringence Δn = neno and dichroism Δk = keko, respectively.

Semiconductors generally have relatively high stimulated intrinsic absorption coefficients and exhibit continuous absorption spectra [23]. When the energy of photons is sufficiently high (equal to or greater than the Eg), photons can interact with electrons in the valence band, causing electron transitions to the conduction band. This process is known as intrinsic absorption and serves as the primary mechanism for optical absorption in semiconductors. In the region where the photon frequency (ν) is less than the Eg divided by the Planck constant (h), it is difficult to excite electrons from the valence band to the conduction band, resulting in a small absorption coefficient. The no and ne can be fitted using the Cauchy dispersion formula or empirical formulas like the Sellmeier equation [11,24]. Figure 1(a) illustrates the increase in the n with increasing Sc doping concentration. The spectra of all samples exhibit a clear distinction between the transparent region and the absorption region with a critical point that closely approximates the Eg. As the energy of photons approaches the Eg, more electrons can transit from the valence band to the conduction band, resulting in an increase in the optical absorption coefficient (α). There is a positive correlation between a and k. As the photon energy increases, distinct extinction coefficient values appear near the Eg of Al1−xScxN, as shown in Fig. 1(b). In the region where  > Eg, the n undergoes a sudden change, exhibiting anomalous dispersion. It is worth noting that in this region, the no and ne exhibit two distinct peaks, resulting in the intersection of refractive indices and leading to the disappearance of birefringence at particular wavelengths.

There is a relationship between Eg and α for semiconductors [25],

$${({\alpha h\nu } )^{\frac{1}{s}}} = B({h\nu - {E_\textrm{g}}} ),$$
where B is a proportionality constant. The value of s depends on the type of semiconductor material. When the semiconductor is a direct band gap material, s equals to 1/2. It should be noted that Al1−xScxN can be considered as a direct band gap material when x < 0.25 [7,9]. Based on the data reported by Zhang et al. [7], as the Sc doping concentration increases, the corresponding Eg decreases. The curves in Figs. 1(a) and 1(b) depict a prominent redshift of n and k as a function of hv. Utilizing the Eq. (2) proposed by Tauc et al., it can be inferred that α increases as a function of hv when hv > Eg, and the value of α at short wavelengths increases as Eg decreases. Figure 2(b) shows a decreasing trend of Eg as x increases. Therefore, when x increases, the value of k at short wavelengths also increases.

 figure: Fig. 2.

Fig. 2. (a) The direct Eg of ordinary and extraordinary lights are extracted by linear fitting for Al1−xScxN films (x = 0.087 and 0.181). (b) Eg of Al1−xScxN as a function of x in Al1−xScxN. The purple dashed line corresponds to the theoretical Eg reported by Zhang et al. [7]. The orange dashed line corresponds to the calculation formula reported by Baeumler et al. [9]. Blue dots and red dots are the direct Eg for this work. (c) Band structure and respective optical transitions between valence bands and conduction band in wurtzite Al1−xScxN. (d) Δcr as a function of x in Al1−xScxN.

Download Full Size | PDF

The Δn and Δk are investigated to provide more insights for the anisotropy of Al1−xScxN thin films. The extrema in the Δn and Δk are typically associated with optical transitions governed by the polarization selection rules [20]. As shown in Fig. 1(c), two extrema are observed in the birefringence Δn curve near the Eg and the minimum and maximum values for x = 0.087 occur at 5.79 eV and 5.56 eV, respectively. For x = 0.181, the minimum and maximum Δn values are observed at 5.34 eV and 5.06 eV, respectively. As the peak of no exceeds the limit of the spectrometer, the minimum Δn value cannot be obtained for AlN (x = 0). The minimum and maximum of Δn curve align with the positions of peaks in the ne and no. These extrema can be attributed to the energy of optical transitions. Similarly, Fig. 1(d) illustrates an obvious peak in the Δk near the Eg. As x increases, the peaks of Δn and Δk also increase, which means that the optical birefringence phenomenon contributed by its non-centrosymmetric crystal structure of Al1−xScxN becomes more pronounced. The noticeable Δn and Δk observed in Al1−xScxN at short wavelengths make it promising for applications in ultraviolet photonic devices.

The Eg of a semiconductor material can be determined by analyzing its absorption spectrum. By substituting α = 4πk/λ into Eq. (2) and plotting the value of (αhν)2 on y-axis against the value of on x-axis, Eq. (2) can be regarded as a linear equation. The value of Eg can be obtained by the intercept of the line on x-axis. By performing linear regression on the linear portion of the graph, a linear equation can be obtained as demonstrated in Fig. 2(a). When x = 0.087 and x = 0.181, the determined band gap energies for ordinary light Eg,o in Al1−xScxN are 5.73 eV and 5.32 eV, respectively. For extraordinary light, the band gap energies Eg,e are determined to be 5.50 eV and 5.06 eV. In the previous papers, the Eg that is not distinguishable between ordinary light and extraordinary light is denoted as Eg,eff [7,9]. Within the range of 0 < x < 0.25, Baeumler et al. extended Vegard's law by incorporating a nonlinear term to describe the bending of the Eg with respect to x in theoretical predictions for wurtzite Al1−xScxN [9]. An empirical relation is derived to describe the composition dependence of the Eg which is indicated by the orange dashed line in Fig. 2(b). In addition, the Eg reported by Zhang et al. are indicated by the purple dashed line [7]. Both two dashed lines exhibit decreasing trend of the Eg with increasing x. The Eg obtained in this work are marked with dots. As compared to Eg,e, it is worth noting that Eg,o values are closer to the Eg,eff values from literature. It is conjectured that this is because luminescence detections approach Ec condition instead of Ec condition, resulting in the determined Eg,eff that are closer to Eg,o [2021].

The band structure of wurtzite Al1−xScxN is illustrated in Fig. 2(c). Due to the hexagonal crystal-field and spin-orbital splitting acting on the triply degenerate p-like valence band, the top of the valence band in Al1−xScxN is divided into three bands: $\mathrm{\Gamma}_{7v}^u$, Γ9 v and $\mathrm{\Gamma}_{7v}^l$. The corresponding interband transitions from these valence bands to the bottom of the conduction band Γ7c are represented by transitions labeled as A, B, and C [20,22].

Anisotropy is attributed to the ordered valence band at the Brillouin zone and a significant crystal-field splitting is observed based on the fitting data. According to the previous findings by Suzuki et al. and Jiang et al. [20,26], it has been established that the spin-orbit splitting Δso in AlN is very small. Therefore, based on the band structure, it is inferred that the crystal-field splitting energy Δcr is approximately equal to Eg,eEg,o. The Δcr are found to be −0.23 eV and −0.26 eV for x = 0.087 and x = 0.181, respectively. According to the data reported by Jiang et al. [20], the Δcr of AlN is approximately −0.2 eV. Thus, it can be inferred that the Δcr increases with the increase in Sc concentration, as plotted in Fig. 2(d). Due to the negative Δcr of Al1−xScxN, the polarization of extraordinary light selectively allows the A transition (Ec). In contrast, for ordinary light polarization with the E perpendicular to c-axis (Ec), both B and C transitions are allowed [20].

3. Waveguide design

The refractive index of a core material plays a crucial role in simulating single mode conditions of a photonic waveguide, which ensures efficient light transmission and mode confinement. Based on the SE measurement, the n of the Al1−xScxN thin films were further fitted by the Cauchy dispersion model in the wavelength range from 250 to 2000nm, as illustrated in Fig. 3(a). There is an approximately linear relationship (n = ax + b) between the refractive indices and the Sc doping concentration with x < 0.25. The slope and intercept of this linear relationship are illustrated in Fig. 3(b). When λ > 900 nm, these two values tend to stabilize. The relationship between refractive indices and x at λ = 1550 nm can be expressed by the following equations:

$${n_e} = 0.533x + 2.1530\quad(x < 0.25),$$
$${n_o} = 0.565x + 2.0829\quad(x < 0.25).$$

 figure: Fig. 3.

Fig. 3. (a) The no and ne at different Sc concentrations are fitted using the Cauchy dispersion formula from λ=250 to 2000nm. (b) When λ > 900 nm, the refractive indices have an approximately linear relationship with the Sc concentration. (c),(d) The neff simulation of Al1−xScxN with x = 0.087 and 0.181.

Download Full Size | PDF

To minimize modal dispersion, the geometric structure of Al1−xScxN channel waveguides is designed to support only the first-order mode. Based on the fitted n, the effective refractive indices (neff) of 400-nm-thick Al1−xScxN waveguides are calculated using the Ansys Lumerical MODE solver. The sidewall angle of the simulated waveguide is fixed at 70 degrees based on fabrication results. The n of the SiO2 cladding is 1.44. For TE mode, the E field of propagating light is perpendicular to c-axis (Ec), thus the no was chosen for the simulation. On the contrary, if the E of propagating light is parallel to c-axis (Ec), the ne was chosen to simulate the transverse magnetic (TM) mode [27].

Figures 3(c),(d) present the simulation results of the neff as a function of the top waveguide width (W) for TE0, TE1 and TM modes for Al1−xScxN waveguides at x = 0.087 and 0.181 at λ = 1550 nm. The height of the waveguide is fixed at 400 nm. Light is better confined within the Al1−xScxN waveguides as W increases, leading to an increase in the neff. When W is further increased, high-order modes will be generated. In addition, the n of Al1−xScxN increases as x increases, which results in a smaller W required for the single mode behaviors. The W for single TE mode condition is simulated to be 0.84 μm when x = 0.087, whereas this value decreases to 0.78 μm at x = 0.181.

4. Device fabrication and characterization

To fabricate the Al1−xScxN photonic devices, an i-line photolithography stepper was used to pattern the 400 nm-thick Al1−xScxN thin films. After that, the Al1−xScxN channel waveguides were formed by chlorine-based inductively coupled plasma (ICP) etching process. A 2 μm-thick SiO2 cladding was deposited using plasma enhanced chemical vapor deposition (PECVD) after the photoresist stripping process. The three-dimensional (3D) schematic of the as-fabricated device is illustrated in Fig. 4(a). The SEM measurement results of critical dimension and a schematic diagram of the design layout are shown in Fig. S2 of Supplement 1. The 45-degree tilted SEM image of the waveguide is presented in Fig. S5 of Supplement 1.

 figure: Fig. 4.

Fig. 4. (a) 3D schematic of the fabricated Al1−xScxN waveguide. The thickness of the Al1−xScxN waveguide and the upper cladding are 0.4 μm and 2 μm, respectively. (b) Photograph of the 8-inch Al1−xScxN photonic device wafer. (c) XSEM image of an Al0.819Sc0.181N waveguide. (d),(e) Optical microscopy images of devices to test the waveguide propagation loss and the bending loss, respectively.

Download Full Size | PDF

Figure 4(b) shows a photograph of the Al0.913Sc0.087N photonic devices on an 8-inch wafer. In general, the fabricated optical waveguides are not perfect rectangles due to the dry etching process. Figure 4(c) displays a cross-sectional scanning electron microscopy (XSEM) image of an Al0.819Sc0.181N waveguide. The waveguide has a sidewall angle of around 70°. As compared to the designed values, the top waveguide width W is reduced by approximately 70 nm for x = 0.087 and around 100 nm for x = 0.181 due to process errors. Figures 4(d),(e) show the optical microscopy images of the testing structures used for measuring waveguide propagation loss and bending loss. The spiral waveguides with lengths of 0.4, 0.8, 1.2 and 1.6 cm are employed to measure waveguide propagation loss. The structures shown in Fig. 4(e) are used to evaluate bending loss with the number of 90° bends being 6, 18, and 30. The design parameters and coupling loss for grating couplers are provided in Table S1 of Supplement 1. The simulated grating coupling efficiency as well as the impact of process error on it are presented in Fig. S3 of Supplement 1. The measured data of the grating coupler is shown in Fig. S4 of Supplement 1.

The propagation loss and bending loss of the channel waveguides were tested. A tunable laser in the Keysight 8164B Mainframe was used to emit the light. The light was passed through a polarization controller and was then coupled into the waveguide using a single mode optical fiber. The grating couplers used in the experiment were designed for the TE polarization mode. The optical power was obtained from the Keysight 8164B Lightwave Measurement System.

At λ = 1550 nm, the propagation losses of the Al1−xScxN channel waveguides with x = 0.087 and x = 0.181 were obtained by the cut-back methods, as shown in Figs. 5(a) and 5(b), respectively. The W used in the figures are the designed widths rather than the actual widths of the fabricated waveguides. The lowest propagation losses for the TE single mode waveguides are 5.98 ± 0.11 dB/cm and 8.23 ± 0.39 dB/cm, respectively. The error bar represents the error in linear fitting. It can be observed that the propagation losses for x = 0.181 are higher than the ones for x = 0.087, which might be attributed to the larger surface and sidewall roughness of the Al1−xScxN waveguides with a larger x as the uneven surfaces enlarge the scattering of the optical field. After optimizing the etching process, it is possible to suppress the propagation loss significantly [28]. The transmission spectrums and propagation losses for x = 0.087 and 0.181 in C band are shown in Fig. S6 and Fig. S7 of Supplement 1.

 figure: Fig. 5.

Fig. 5. (a),(b) The P vs L characteristics of the Al1−xScxN waveguides at λ = 1550 nm for x = 0.087 and 0.181. The waveguide propagation losses were extracted by linear fitting. (c) Benchmarking of Al1−xScxN waveguide propagation loss on Si substrates at λ = 1550 nm. (d),(e) The bending losses of the Al1−xScxN waveguides with x = 0.087 and x = 0.181 extracted by linear fitting.

Download Full Size | PDF

Under TE single mode conditions, the propagation loss decreases as the waveguide width increases. When x = 0.087, considering the 70 nm process error, the designed maximum W is reduced from 0.9 μm to 0.83 μm. This value is smaller than the single mode condition of 0.84 μm obtained from the simulation. Consequently, under the TE mode at λ = 1550 nm, the structures remain in the single mode condition. The propagation loss decreases as W increases due to the smaller sidewall scattering effect.

However, when x = 0.181, the waveguide width of 0.9 μm exhibits the highest propagation loss. This might be due to that the higher mode is generated for the devices with W = 0.9 μm as the actual waveguide width is around 0.8 μm, which just exceeds the limit of the simulated single TE mode condition (0.78 μm) but is not enough to well confine the high-order mode. In this case, the scattering of high-order mode during propagation is large, resulting in increased propagation loss.

The optical propagation losses of Al1−xScxN waveguides on Si substrates at λ = 1550 nm are benchmarked with the results from literature [13,2931], as shown in Fig. 5(c). The propagation losses of AlN (x = 0) waveguides are also summarized for comparison. It can be observed that the average waveguide propagation loss for AlN is approximately 0.6 dB/cm [2931], which is much smaller than those of the Al1−xScxN waveguides.

The high propagation loss in Al1−xScxN is caused by multiple factors, including scattering and absorption. Generally, a higher Sc concentration may lead to a larger optical propagation loss of the waveguide.

The scattering loss could be mainly attributed into three parts. Firstly, the abnormal grain growth, which is induced by complexion formation at crystallite interfaces, becomes more severe in Al1−xScxN thin films with higher Sc concentration. This results in the formation of the abnormally oriented grains (AOGs) which enhance the scattering of light [32]. Secondly, the reduction in c-axis texture of the film with increasing Sc content may affect the propagation path of light within the crystal, making it more prone to deflection and scattering, further increasing the propagation loss [33]. Thirdly, in the fabrication process of the waveguide, the high hardness of Al1−xScxN makes etching difficult, and the immaturity of the etching process leads to rough sidewalls, resulting in significant scattering loss.

In optical waveguides, the absorption characteristics of the material can cause gradual reduction of light power during propagation. Due to the imperfect growth process of Al1−xScxN thin films, the introduction of Sc into AlN may introduce new defects or impurity levels, which can act as centers for light absorption, further increasing absorption loss.

To reduce waveguide losses, it is necessary to optimize the etching process such as adjusting the dry etching parameters and improving the selection of etchants. Additionally, optimizing the growth conditions of Al1−xScxN thin films, including adjusting growth parameters and incorporating post-growth annealing treatments, which can enhance crystallinity and mitigate the impact of abnormal grain orientation.

In addition, it is observed that the propagation loss values from this work are lower than that from Ref. [13]. The propagation loss might be further suppressed by optimizing the thin-film deposition, lithography and dry etching process conditions, or by implementing high temperature annealing process after waveguide formation, which are subjects of our future works.

Furthermore, we have investigated the bending losses of Al1−xScxN waveguides at λ = 1550 nm. Measurements were conducted on 90° bends with varying numbers of 50 μm radius in TE mode. The results are plotted in Figs. 5(d) and 5(e). It can be observed that the bending losses exhibit slight fluctuations for the waveguides with different W. In addition, the bending losses for x = 0.087 and x = 0.181 were measured to be 0.05 dB/turn and 0.08 dB/turn, respectively. The greater bending loss could be attributed to the increased scattering losses caused by the clustering of Sc and the rough surfaces of waveguides made from Al1−xScxN with higher values of x. Although increasing the radius beyond 50 μm could further decrease the bending loss slightly, it will also increase the footprint of the photonic devices significantly. Therefore, we selected a bending radius of 50 μm for the actual design in this experiment. In the future, we will conduct systematic experimental studies on the bending loss as a function of bending radii.

5. Conclusion

In summary, the anisotropic optical properties of Al1−xScxN films are investigated and Al1−xScxN photonic devices are demonstrated on 8-inch Si wafer. The Eg, n and k of ordinary light and extraordinary light in Al1−xScxN with various x are investigated for the first time. In addition, the propagation loss of Al1−xScxN waveguides and the waveguide bending loss are characterized. The single mode waveguides with x = 0.087 and 0.181 exhibit propagation losses of 5.98 ± 0.11 dB/cm and 8.23 ± 0.39 dB/cm, respectively. The 90° bending losses with a radius of 50 μm are 0.05 dB/turn and 0.08 dB/turn at λ = 1550 nm, respectively, for x = 0.087 and 0.181. This work paves the way for large-scale integrated Al1−xScxN photonic circuits on Si platforms.

Funding

Shanghai Collaborative Innovation Center of Intelligent Sensing Chip Technology; Shanghai Key Laboratory of Chips and Systems for Intelligent Connected Vehicle; National Natural Science Foundation of China (62204149, 62205193, U23A20356).

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. M. Soltani, R. Soref, T. Palacios, et al., “AlGaN/AlN integrated photonics platform for the ultraviolet and visible spectral range,” Opt. Express 24(22), 25415–25423 (2016). [CrossRef]  

2. H. Yamashita, K. Fukui, S. Misawa, et al., “Optical properties of AlN epitaxial thin films in the vacuum ultraviolet region,” J. Appl. Phys. 50(2), 896–898 (1979). [CrossRef]  

3. A. Sanz-Hervás, M. Clement, E. Iborra, et al., “Degradation of the piezoelectric response of sputtered c-AlN thin films with traces of non-(0002) x-ray diffraction peaks,” Appl. Phys. Lett. 88(16), 161915 (2006). [CrossRef]  

4. X. Liu, A. W. Bruch, and H. X. Tang, “Aluminum nitride photonic integrated circuits: from piezo-optomechanics to nonlinear optics,” Adv. Opt. Photonics 15(1), 236–317 (2023). [CrossRef]  

5. N. Li, C. P. Ho, S. Zhu, et al., “Aluminium nitride integrated photonics: a review,” Nanophotonics 10(9), 2347–2387 (2021). [CrossRef]  

6. W. H. P. Pernice, C. Xiong, C. Schuck, et al., “Second harmonic generation in phase matched aluminum nitride waveguides and micro-ring resonators,” Appl. Phys. Lett. 100(22), 223501 (2012). [CrossRef]  

7. S. Zhang, D. Holec, W. Y. Fu, et al., “Tunable optoelectronic and ferroelectric properties in Sc-based III-nitrides,” J. Appl. Phys. 114(13), 133510 (2013). [CrossRef]  

8. M. Moram and S. Zhang, “ScGaN and ScAlN: Emerging nitride materials,” J. Mater. Chem. A. 2(17), 6042–6050 (2014). [CrossRef]  

9. M. Baeumler, Y. Lu, N. Kurz, et al., “Optical constants and band gap of wurtzite Al1−xScxN/Al2O3 prepared by magnetron sputter epitaxy for scandium concentrations up to x = 0.41,” J. Appl. Phys. 126(4), 045715 (2019). [CrossRef]  

10. V. Yoshioka, J. Lu, Z. Tang, et al., “Strongly enhanced second-order optical nonlinearity in CMOS-compatible Al1−xScxN thin films,” APL Mater. 9(10), 101104 (2021). [CrossRef]  

11. D. H. Jundt, “Temperature-dependent Sellmeier equation for the index of refraction, ne, in congruent lithium niobate,” Opt. Lett. 22(20), 1553–1555 (1997). [CrossRef]  

12. J. Liu, W. Shin, P. Wang, et al., “Ferroelectric ScAlN: Epitaxy, properties, and emerging photonic device applications,” in Photonics Conference (IEEE, 2022), pp. 1–2.

13. S. Zhu, Q. Zhong, N. Li, et al., “Integrated ScAlN photonic circuits on silicon substrate,” in Conference on Lasers and Electro-Optics (Optical Society of America, 2020), pp. 1–2.

14. J. Su, S. Fichtner, M. Z. Ghori, et al., “Growth of highly c-Axis oriented AlScN films on commercial substrates,” Micromachines 13(5), 783 (2022). [CrossRef]  

15. C. Buchheim, R. Goldhahn, M. Rakel, et al., “Dielectric function and critical points of the band structure for AlGaN alloys,” Phys. Status Solidi B 242(13), 2610–2616 (2005). [CrossRef]  

16. P. Schley, R. Goldhahn, A. Winzer, et al., “Dielectric function and Van Hove singularities for in-rich in x Ga1−xN alloys: Comparison of N-and metal-face materials,” Phys. Rev. B 75(20), 205204 (2007). [CrossRef]  

17. E. Sakalauskas, H. Behmenburg, C. Hums, et al., “Dielectric function and optical properties of Al-rich AlInN alloys pseudomorphically grown on GaN,” J. Phys. D 43(36), 365102 (2010). [CrossRef]  

18. H. Fujiwara, Spectroscopic Ellipsometry: Principles and Applications (John Wiley & Sons Publishing, 2007).

19. A. I. Plekhanov, V. P. Chubakov, and P. A. Chubakov, “Polarization anisotropy of single crystal opal films,” Phys. Solid State 53(6), 1145–1151 (2011). [CrossRef]  

20. W. Jiang, W. Lin, S. Li, et al., “Optical anisotropy of AlN epilayer on sapphire substrate investigated by variable-angle spectroscopic ellipsometry,” Opt. Mater. 32(9), 891–895 (2010). [CrossRef]  

21. D. Gonçalves and E. A. Irene, “Fundamentals and applications of spectroscopic ellipsometry,” Quim. Nova 25(5), 794–800 (2002). [CrossRef]  

22. T. Onuma, S. Chichibu, T. Sota, et al., “Exciton spectra of an AlN epitaxial film on (0001) sapphire substrate grown by low-pressure metalorganic vapor phase epitaxy,” Appl. Phys. Lett. 81(4), 652–654 (2002). [CrossRef]  

23. M. E. Levinshtein, S. L. Rumyantsev, and M. S. Shur, Properties of Advanced Semiconductor Materials: GaN, AIN, InN, BN, SiC, SiGe, M. E. Levinshtein, S. L. Rumyantsev, and M. S. Shur, eds. (John Wiley & Sons Publishing, 2001).

24. P. Langhoff and M. Karplus, “Pade summation of the Cauchy dispersion equation,” J. Opt. Soc. Am. 59(7), 863–871 (1969). [CrossRef]  

25. P. Makuła, M. Pacia, and W. Macyk, “How to correctly determine the band gap energy of modified semiconductor photocatalysts based on UV-Vis spectra,” J. Phys. Chem. Lett. 9(23), 6814–6817 (2018). [CrossRef]  

26. M. Suzuki, T. Uenoyama, and A. Yanase, “First-principles calculations of effective-mass parameters of AlN and GaN,” Phys. Rev. B 52(11), 8132–8139 (1995). [CrossRef]  

27. A. W. Snyder and W. R. Young, “Modes of optical waveguides,” J. Opt. Soc. Am. 68(3), 297–309 (1978). [CrossRef]  

28. T. Barwicz and H. A. Haus, “Three-dimensional analysis of scattering losses due to sidewall roughness in microphotonic waveguides,” J. Light. Technol. 23(9), 2719–2732 (2005). [CrossRef]  

29. C. Xiong, W. H. Pernice, and H. X. Tang, “Low-loss, silicon integrated, aluminum nitride photonic circuits and their use for electro-optic signal processing,” Nano Lett. 12(7), 3562–3568 (2012). [CrossRef]  

30. A. Soltani, A. Stolz, J. Charrier, et al., “Dispersion properties and low infrared optical losses in epitaxial AlN on sapphire substrate in the visible and infrared range,” J. Appl. Phys. 115(16), 163515 (2014). [CrossRef]  

31. S. Zhu, Q. Zhong, T. Hu, et al., “Aluminum nitride ultralow loss waveguides and push-pull electro-optic modulators for near infrared and visible integrated photonics,” in Optical Fiber Communications Conference and Exhibition (Optical Society of America, 2019), pp. 1–3.

32. C. S. Sandu, F. Parsapour, S. Mertin, et al., “Abnormal grain growth in AlScN thin films induced by complexion formation at crystallite interfaces,” Phys. Status Solidi. (a) 216(2), 1800569 (2019). [CrossRef]  

33. O. Zywitzki, T. Modes, S. Barth, et al., “Effect of scandium content on structure and piezoelectric properties of AlScN films deposited by reactive pulse magnetron sputtering,” Surf. Coat. Technol. 309, 417–422 (2017). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       The Supplementary Materials are shown in Fig. S1-S7

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 (5)

Fig. 1.
Fig. 1. The dispersions of (a) no, ne and (b) ko, ke for Al1−xScxN films at photon energy ranges from 1.24 to 5.905 eV. The (c) Δn and (d) Δk plotted as a function of hv for Al1−xScxN films.
Fig. 2.
Fig. 2. (a) The direct Eg of ordinary and extraordinary lights are extracted by linear fitting for Al1−xScxN films (x = 0.087 and 0.181). (b) Eg of Al1−xScxN as a function of x in Al1−xScxN. The purple dashed line corresponds to the theoretical Eg reported by Zhang et al. [7]. The orange dashed line corresponds to the calculation formula reported by Baeumler et al. [9]. Blue dots and red dots are the direct Eg for this work. (c) Band structure and respective optical transitions between valence bands and conduction band in wurtzite Al1−xScxN. (d) Δcr as a function of x in Al1−xScxN.
Fig. 3.
Fig. 3. (a) The no and ne at different Sc concentrations are fitted using the Cauchy dispersion formula from λ=250 to 2000nm. (b) When λ > 900 nm, the refractive indices have an approximately linear relationship with the Sc concentration. (c),(d) The neff simulation of Al1−xScxN with x = 0.087 and 0.181.
Fig. 4.
Fig. 4. (a) 3D schematic of the fabricated Al1−xScxN waveguide. The thickness of the Al1−xScxN waveguide and the upper cladding are 0.4 μm and 2 μm, respectively. (b) Photograph of the 8-inch Al1−xScxN photonic device wafer. (c) XSEM image of an Al0.819Sc0.181N waveguide. (d),(e) Optical microscopy images of devices to test the waveguide propagation loss and the bending loss, respectively.
Fig. 5.
Fig. 5. (a),(b) The P vs L characteristics of the Al1−xScxN waveguides at λ = 1550 nm for x = 0.087 and 0.181. The waveguide propagation losses were extracted by linear fitting. (c) Benchmarking of Al1−xScxN waveguide propagation loss on Si substrates at λ = 1550 nm. (d),(e) The bending losses of the Al1−xScxN waveguides with x = 0.087 and x = 0.181 extracted by linear fitting.

Equations (4)

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

ρ = tan φ e i Δ = r p / r s ,
( α h ν ) 1 s = B ( h ν E g ) ,
n e = 0.533 x + 2.1530 ( x < 0.25 ) ,
n o = 0.565 x + 2.0829 ( x < 0.25 ) .
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.