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

Analytic ellipsometric measurement for materials under bulk encapsulation

Open Access Open Access

Abstract

Accurate measurement of the dielectric functions of emerging optical materials is of great importance for advancements in solid-state physics. However, it is rather challenging since most materials are highly active in ambient conditions, which makes in-situ measurements tough. Here, we report an analytical ellipsometry method (AEM) accessible in ambient conditions for measuring the dielectric functions of chemically reactive materials under bulk encapsulation. Taking the highly pursued low-loss plasmonic materials, such as sodium films, as an example, the effectiveness and measuring errors of the proposed AEM have been systematically demonstrated. This verifies AEM’s superiority in overcoming the limitations of traditional spectroscopic ellipsometry methodologies, which include complex multi-parameter fitting procedures and the issue of potentially unphysical results, especially in newly developed low-loss materials. Our results will provide a generalized and convenient ellipsometric measurement strategy for sensitive materials under bulk encapsulation.

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

1. Introduction

In the fields of solid-state physics, the complex dielectric function is a fundamental cornerstone for exploring light-matter interactions in emerging materials [1], as well as for the rationally designing various photonic and optoelectronic devices, such as photonic chips [2,3], biosensing [4,5], photochemistry [6,7], etc.

Recent decades have evidenced great advancements in commercial spectroscopic ellipsometry (SE) techniques, leading to a significant expansion of the artificial material database. High-quality materials with low optical loss or poor ambient tolerance, including two-dimensional materials [8,9], perovskites [10,11], and other related organic and inorganic materials, are noted for their highly desired physical properties. However, most of these materials often necessitate in-situ vacuum-based SE measurement integrated with the fabrication processes, significantly raising infrastructure costs and operational complexity, thus limiting practical applications. An alternative strategy to manage reactive materials outside the vacuum involves applying stable encapsulating films to the targets. However, this introduces additional fitting parameters, complicating the SE optical models and making the fitting of the imaginary part of dielectric functions rather challenging, particularly for materials with low absorption coefficients or losses. Consequently, there is an urgent demand for ambient available alternative methods to traditional SE, especially for research involving low-loss and highly sensitive materials.

Specifically, as for plasmonic materials, the integration of various optical fitting models of SE with well-developed metallic samples, from polycrystalline and single crystalline films [12] to high-quality alloys [13,14], shows that conventional noble metal-based plasmonics are almost approaching the inherent optical loss limit predicted by theories [15,16]. Recently, Yang Wang and coworkers utilized the thermo-assisted spin coating method to overcome the fabrication bottleneck, creating high-quality sodium films encapsulated by quartz substrates [17]. This experimental breakthrough led to a record-low-loss plasmonic material, surpassing noble metal silver in the near-infrared and achieving an optical damping rate of 10 meV through standard multi-parameter SE modeling, which serves as crucial evidence and new referencing systems for the ongoing explorations such as low threshold nanolasors [18], integrated optics [19] and single particle strong coupling [20], etc. However, applying standard SE measurements to bulk dispersive encapsulated systems is still challenging, which will introduce significant measuring errors or even unphysical fitting solutions [17] (see Supplement 1 for more fitting details). Therefore, there is an urgent need for a precise and convenient SE technique that does not rely on fitting models, which is crucial for advancing the study of encapsulated material systems and thus of significant interest across various branches of solid-state physics.

Here, we demonstrate an analytical ellipsometry method (AEM) for the measurement of dielectric functions of encapsulated materials, which effectively eliminates fitting errors and avoids nonphysical results. Taking the highly pursued low-loss plasmonic materials, such as sodium films, as an example, which are prepared by the thermo-assisted spin-coating method proposed in our previous work, with a thickness of hundreds of microns approximately. After encapsulation with fused silica and side packaging with ultraviolet curing glue, the sodium films can be transferred freely out of glove boxes for in-air SE measurements by RC2 ellipsometer (J.A. Woollam Corporation). We measured the dielectric functions of the sodium films across a broad spectral range from 0.50 to 4.97 eV (250 nm to 2.5 µm). In addition, by fitting the Drude model to the dielectric functions of more than 200 sodium samples, the statistical Drude parameters of sodium (ɛ, ħωp, ħγ) can be obtained, which provides direct experimental evidence that sodium is an ideal candidate with a low damping rate beyond noble metals. Finally, the measurement errors of the proposed method can be analytically evaluated to be less than 1.7% for the real part and 4.4% for the imaginary part of the dielectric function. In conclusion, our method provides a simple and precise alternative to traditional methods for low-loss and highly sensitive materials, beneficial for both precise optical designs in photonics and fundamental studies in solid-state physics.

2. Analytical ellipsometry method (AEM)

To establish an analytic formula of the dielectric functions of sodium as a function of the SE raw data ρm(ω), the dielectric functions of encapsulated metals ${\varepsilon _m}(\omega )$ can be expressed as the following formula by a simple formulation from Snell's Law and Fresnel Formula:

$${\varepsilon _m}(\omega )= si{n^2}\alpha \left[ {1 + ta{n^2}\beta {{\left( {\frac{{T - {\rho_m}(\omega )}}{{T + {\rho_m}(\omega )}}} \right)}^2}} \right].$$

More detailed derivations are available in Supplement 1, Sec. 2. As depicted in Fig. 1, α is the incident angle, β is the refractive angle on silica obtained by $sin\beta = \frac{{sin\alpha }}{{\sqrt {{\varepsilon _{sub}}(\omega )} }}$, and the parameter T can be derived as the following formula:

$$T = \frac{{si{n^2}\beta {{\left( {cos\alpha + \sqrt {{\varepsilon_{sub}}(\omega )} cos\beta } \right)}^2}}}{{si{n^2}({\alpha + \beta } )co{s^2}({\alpha - \beta } )}}.$$

 figure: Fig. 1.

Fig. 1. The schematic diagram of the in-air ellipsometry measurement on the encapsulated surface. The light beams depicted in dark red are the measurement light beams needed to collect.

Download Full Size | PDF

${\varepsilon _{sub}}(\omega )$, the dielectric function of the isotropic quartz, can be obtained by the following formula:

$${\varepsilon _{sub}}(\omega )= si{n^2}{\theta _0}\left\{ {1 + ta{n^2}{\theta_0}{{\left[ {\frac{{1 - {\rho_{sub}}(\omega )}}{{1 + {\rho_{sub}}(\omega )}}} \right]}^2}} \right\},$$
where ${\theta _0}$ is the incident angle with respect to the surface normal of the sample, and ${\rho _{sub}}(\omega )$ is the polarization state change of the surface reflected beam. By combining Eqs. (23) with Eq. (1), one can retrieve the ${\varepsilon _m}(\omega )$ as a function of the incident angle (α and ${\theta _0}$) and the polarization state variation (${\rho _m}(\omega )$ and ${\rho _{sub}}(\omega )$ measured from conventional SE). Through the analytical ellipsometry method (AEM), the systematic errors of the previous work can be effectively suppressed and thus more precise measured dielectric functions can be expected.

To minimize the depolarization effect due to the fused silica substrate, there physical phenomena that generate partially polarized light upon light reflection, including surface light scattering, multiple reflections, and incident angle variation, are taken into account in our experimental design and analysis. Eventually, a 1.5-mm-thick fused silica with a surfaces roughness of ∼100 pm is used as the substrate (more details of thickness choice are available in Supplement 1, Sec. 3). Besides, to evaluate the validity of the proposed AEM, a pre-evaluation procedure of the proposed AEM is conducted on conventional plasmonic metals, such as gold and silver, before the measurements on sodium (more experiment details are available in Supplement 1, Sec. 4). Figs. 2(a) and (b) illustrate the measured ε1 and ε2 of the gold film and silver film respectively, obtained by analytically measuring without the fused silica (solid line) and the AEM with the fused silica (dotted line). It is depicted that, the dielectric functions of noble metals measured with fused silica coincide well with the corresponding standard data, evidencing the validity of the AEM. Furthermore, measurements at multiple angles of incidence are also conducted on sodium samples (see Supplement 5), which also show the validity of the AEM.

 figure: Fig. 2.

Fig. 2. Evaluation the validity of the AEM. The dielectric function of (a) Au and (b) Ag measured with and without covering fused silica respectively.

Download Full Size | PDF

3. Results and discussion

The dielectric function of materials reflects the intrinsic light-matter interactions. Specifically, it includes electromagnetic field confinement and intrinsic optical loss for plasmonic metals. Fig. 3(a) shows ${\varepsilon _1}$ of the sodium film across a spectral range of 0.50–4.97 eV in red circles, while Figs. 3(b) and (c) show ${\varepsilon _2}$ in the energy range of 1.77–4.97 eV and 0.50–1.77 eV in yellow and blue circles, respectively. For comparison, data from Wang (left-triangle) [17], Hodgson (rhombus) [21], Smith (up-triangle) [22], Palmer (circles) [23], Monin (down-triangle) [24], and Inagaki (squares) [25] are plotted. The dielectric functions are sensitive to sample quality and measurement methods, which may induce differences among data from our work and previous literature.

 figure: Fig. 3.

Fig. 3. (a) The ${\varepsilon _1}$ and (b)-(c) ${\varepsilon _2}$ of the dielectric functions of sodium from 0.50 eV to 4.97 eV (250-2500 nm). The measured data of the dielectric functions of sodium reported in the literature are shown for comparison. The statistical data of Drude model parameters (d) ${\varepsilon _\infty }$, (e) $\hbar {\omega _p}$, and (f) $\hbar \gamma $ are obtained from over 200 sample points.

Download Full Size | PDF

The ${\varepsilon _1}$ is related to the capability of field confinement nearby the sodium/quartz interface, which is determined by the density of free carriers theoretically. Since the free electron density of sodium is generally up to 3 × 1028 /m3 [26], the ${\varepsilon _1}$ of sodium fabricated and measured by different groups are close to each other. However, the imaginary parts measured by different research groups show distinct diversity. The ${\varepsilon _2}$ is associated with intrinsic optical loss, including interband and intraband absorption. In Fig. 3(b), the ${\varepsilon _2}$ shows an inconspicuous Lorentz-like peak around 2.8 eV basically consistent with the literature, corresponding to the weak interband transition in sodium. Unlike the interband absorption, the intraband absorption of free carriers within the conduction band mainly determines the ${\varepsilon _2}$ in the infrared spectral range, which is of great importance in the performance of surface plasmon polariton devices. As shown in Fig. 3(c), the ${\varepsilon _2}$ in the infrared spectral range is smaller than the data of other works, indicating the superiority of our fabrication method and high sample quality.

Based on the near free electron gas behaviors of sodium in the near-infrared range, the measured dielectric functions can be described by Drude model:

$${\varepsilon _m}(\omega )= {\varepsilon _1}(\omega )+ i{\varepsilon _2}(\omega )= {\varepsilon _\infty } - \frac{{\omega _p^2}}{{\omega ({\omega + i\gamma } )}}.$$

The parameter ${\varepsilon _\infty }$ accounts for the net contribution from the positive ion cores. The volume plasma frequency ${\omega _p}$ is related to the effective mass of the electron m* and the electron density N through ${\omega _p} = \sqrt {N{e^2}/{\varepsilon _0}{m^\mathrm{\ast }}} $. And the electron relaxation rate $\gamma $ describes the effective electron scattering rate regarded as a figure of merit for optical loss, with a corresponding relaxation time τ = 1/$\gamma $. The dielectric function ${\varepsilon _m}(\omega )$ in Eq. (4) can be decomposed into real ${\varepsilon _1}(\omega )$ and imaginary part ${\varepsilon _2}(\omega )$, which can be simplified as follows under the condition of $\mathrm{\omega } \gg \mathrm{\gamma }$ but below the onset of the interband response,

$${\varepsilon _1}(\omega )= {\varepsilon _\infty } - \frac{{\omega _p^2}}{{{\omega ^2} + {\gamma ^2}}} \approx {\varepsilon _\infty } - \frac{{\omega _p^2}}{{{\omega ^2}}},$$
$${\varepsilon _2}(\omega )= \frac{{\omega _p^2 \times \gamma }}{{\omega ({{\omega^2} + {\gamma^2}} )}} \approx \frac{{\omega _p^2 \times \gamma }}{{{\omega ^3}}}.$$

Equations (5) and (6) allow for the direct calculations of ${\varepsilon _\infty },\textrm{}{\omega _p}$, and $\mathrm{\gamma }$ from ${\varepsilon _1}(\omega )$ and ${\varepsilon _2}(\omega )$. Using this approach in the region of 800 nm to 2.5 µm, it could be obtained that the fitting parameters of sodium are ${\varepsilon _\infty }$ = 1.18 ± 0.23, $\hbar {\omega _p}$ = 5.79 ± 0.08 eV, and $\hbar \mathrm{\gamma }$ = 13.5 ± 7.5 meV. Fig. 3(d-f) demonstrate the statistical data of three parameters from over 200 sample points. Each parameter has a distribution close to the Gauss profile. The peak width is originating from sample quality variations and random errors.

The intraband absorption of free carriers provides information about the electron scattering rate γ, contributed by electron-phonon scattering, electron-electron scattering, surface roughness scattering, and other possible scattering mechanisms. The average $\hbar \mathrm{\gamma }$ extracted from the Drude model is 13.5 meV, and the lowest $\hbar \mathrm{\gamma }$ is merely 6 meV. As a result, the maximum electron relaxation time τ of these sodium samples is ∼ 109 fs. Furthermore, the maximum free path of the electron of sodium is ∼ 116 nm, which can be calculated by $l = {v_f}\tau $, where ${v_f} = $ 1.07 × 108 cm/s is the Fermi velocity of sodium. For comparison, the electron scattering rate of sodium is lower than that of gold (35 ∼ 80 meV) [2729] and silver (15 ∼ 50 meV) [28,30] in the optical region, which means that sodium metal is a more promising plasmonic candidate than noble metals. In addition, since the smaller ${m^\mathrm{\ast }}$ corresponds to more smooth metal surfaces with the same electron density N [29], the fact that the plasma frequency $\hbar {\omega _p}$ of our sodium sample is slightly higher than that of literature (5.2 ∼ 5.8 eV) [2125] demonstrates the high optical quality of our sodium samples among previous works as well. Another parameter of the Drude model is the infinite dielectric constant ${\varepsilon _\infty }$, which is related to the conformity between the metal and free electron structure. The ${\varepsilon _\infty }$ of our sodium samples is close to 1, which agrees well with the previous literature (1.0 ∼ 1.3) [2125] and means that sodium is very close to the free electron structure.

To evaluate the effectiveness of the measured data, we finally perform the error analysis of the proposed AEM measurements. One of errors is due to a surface layer between sodium films and substrate as shown in Fig. 4(a), since sodium is chemically reactive and easy to react with oxygen. In our sample preparation method, with the experiment condition of O2 ∼ 0.1 ppm, the thickness of the oxide layer is only ∼ 0.004 nm, which would be applied to estimate the errors in the dielectric function of sodium. Besides, referring to [31] , the refractive index of oxidation layer is assumed to be from 1.4 to 1.6. As Fig. 4(b) and (c) show, thanks to the efficiency of the fabrication method in removing oxide layers, the error related to interface oxide layer in the real part of the dielectric function is within 0.004%, and that in the imaginary part is within 0.006% (more details are available in Supplement 1, Sec. 6).

 figure: Fig. 4.

Fig. 4. Errors of the AEM. (a) The schematic diagram of the error due to a surface layer. (b) The relative error of ε1 and (c) the relative error of ε2 of sodium samples covered with an 0.004-nm oxidation layer with a refractive index of 1.4, 1.5, and 1.6. (d) The schematic diagram of the error due to surface roughness-induced depolarization effect. (e) The relative error of ε1 and (f) the relative error of ε2 of sodium samples encapsulated by quartz substrates with a surface roughness of ∼100 pm. (g) The schematic diagram of the error due to depolarization effect resulting from substrate dispersion and focusing beam. The light beam depicted in dark red is a beam of a certain wavelength with an incident angle α, while the light beam indicated in light red is a beam of other wavelength with an incident angle α+θ. (h) The relative error of ε1 and (i) the relative error of ε2 of sodium samples caused by substrate dispersion and focusing beam.

Download Full Size | PDF

The second error here is related to surface roughness-induced depolarization effect as shown in Fig. 4(d), which is inevitable but can be minimized by decreasing the surface roughness of quartz substrates. To quantitatively evaluate this error, a simplified optical model with an equivalent rough surface layer is proposed (more details are available in Supplement 1, Sec. 7), where the dielectric function of this layer is calculated by effective medium theories. As shown in Fig. 4(e) and (f), the relative error related to surface roughness (∼100 pm) in the real part of the dielectric function is < 0.7%, and that in the imaginary part is < 1.0%, which suggests that the impact of surface roughness can be considered negligible in our work.

Another potential error is due to depolarization effect resulting from the substrate dispersion and focusing beam. For the encapsulated measuring system, to eliminate this error, a large-area (∼ 4 mm) and high-quality sample surface together with a well-defined ultrathick substrate (see Supplement 1, Fig. S12 in Sec. 8) is ideally needed to fulfill the non-fusing measuring requirement, which is too heavy for the spin coater and raises significant challenges in the sample fabrication process. As an alternative scheme, this depolarization effect can be minimized by carefully optimizing the thickness of the bulky substrate. In this case, the angle of incident light at certain wavelengths detected is actually α+θ due to the substrate dispersion as shown in Fig. 4(g), but the incident angle is considered to be α. Thus, the relative error could be calculated with the analytic expression in Eq. (7), reading as

$$\frac{{\delta \varepsilon }}{\varepsilon } = \frac{{\frac{{\partial \varepsilon }}{{\partial \alpha }} \cdot \delta \theta }}{\varepsilon } \times 100{\%}.$$

Fig. 4(h) and (i) show, as the thickness of the fused silica substrate is 1.5 mm and the incident angle is 65°, the relative error of the ε12) is < 1.0% (< 3.4%). Note that these relative errors would increase as the substrate thickness increases as predicted in Supplement 4. That is why the 1.5-mm-thick fused silica was chosen as the spin-coating substrate and encapsulation layer.

According to the error analysis, the total relative error of the ε1 and ε2 of sodium, is less than 1.7% and 4.4%, respectively. Furthermore, for film materials that can be grown on thick substrate and formed in a large area with high quality, the total relative error of the ε12) by this method could be decreased to < 0.7% (< 1.0%). Therefore, compared to the traditional fitting method, the measured dielectric function of encapsulated materials, especially the imaginary part, is more reliable and precise.

4. Conclusion

Our work introduces a novel analytical ellipsometry method (AEM) for accurately measuring dielectric functions of chemically reactive materials under bulk encapsulation without the multi-parameter fitting errors and the issue of potentially unphysical results in traditional spectroscopic ellipsometry methodologies. The successful application of this method to sodium films, analyzing critical issues like surface overlayer effects and depolarization effects, demonstrates substantial improvements in precision. Specifically, our method reduced measurement errors to less than 1.7% for the real part and 4.4% for the imaginary part of the dielectric function, showcasing a significant advancement in the precision of optical measurements. Furthermore, by fitting the Drude model to the dielectric functions of more than 200 sodium samples, we observed a remarkably low damping rate in sodium, with Drude model parameters indicating an average electron relaxation rate (ℏγ) of 13.5 ± 7.5 meV and the lowest reaching merely 6 meV, pointing to sodium's exceptional potential as a low-loss plasmonic material beyond conventional noble metals. In conclusion, our results provide a generalized and convenient ellipsometry measurement strategy for highly sensitive materials under bulk encapsulation, particularly beneficial for precise optical designs in photonics and fundamental studies in solid-state physics.

Funding

National Key Research and Development Program of China (2021YFA1400700); National Natural Science Foundation of China (12022403, 62375123).

Acknowledgment

The authors thank the micro-fabrication center of National Laboratory of Solid State Microstructures (NLSSM) for technique support.

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. F. Liu, W. Wu, Y. Bai, et al., “Disassembling 2D van der Waals crystals into macroscopic monolayers and reassembling into artificial lattices,” Science 367(6480), 903–906 (2020). [CrossRef]  

2. S. I. Bozhevolnyi, V. S. Volkov, E. Devaux, et al., “Channel plasmon subwavelength waveguide components including interferometers and ring resonators,” Nature 440(7083), 508–511 (2006). [CrossRef]  

3. E. Ozbay, “Plasmonics: merging photonics and electronics at nanoscale dimensions,” Science 311(5758), 189–193 (2006). [CrossRef]  

4. K. Kneipp, Y. Wang, H. Kneipp, et al., “Single molecule detection using surface-enhanced Raman scattering (SERS),” Phys. Rev. Lett. 78(9), 1667–1670 (1997). [CrossRef]  

5. J. Homola, S. S. Yee, and G. Gauglitz, “Surface plasmon resonance sensors: review,” Sens. Actuators, B 54(1-2), 3–15 (1999). [CrossRef]  

6. S. Linic, U. Aslam, C. Boerigter, et al., “Photochemical transformations on plasmonic metal nanoparticles,” Nat. Mater. 14(6), 567–576 (2015). [CrossRef]  

7. Y. Zhang, S. He, W. Guo, et al., “Surface-plasmon-driven hot electron photochemistry,” Chem. Rev. 118(6), 2927–2954 (2018). [CrossRef]  

8. F. J. G. de Abajo, “Graphene plasmonics: challenges and opportunities,” ACS Photonics 1(3), 135–152 (2014). doi:10.1021/ph400147y

9. C. Lian, S. Hu, J. Zhang, et al., “Integrated plasmonics: broadband dirac plasmons in borophene,” Phys. Rev. Lett. 125(11), 116802 (2020). [CrossRef]  

10. P. Löper, M. Stuckelberger, B. Niesen, et al., “Complex refractive index spectra of CH 3 NH 3 PbI 3 Perovskite thin films determined by spectroscopic ellipsometry and spectrophotometry,” J. Phys. Chem. Lett. 6(1), 66–71 (2015). [CrossRef]  

11. M. S. Alias, I. Dursun, M. I. Saidaminov, et al., “Optical constants of CH_3NH_3PbBr_3 perovskite thin films measured by spectroscopic ellipsometry,” Opt. Express 24(15), 16586 (2016). [CrossRef]  

12. J. H. Park, P. Ambwani, M. Manno, et al., “Single-crystalline silver films for plasmonics,” Adv. Mater. 24(29), 3988–3992 (2012). [CrossRef]  

13. D. A. Bobb, G. Zhu, M. Mayy, et al., “Engineering of low-loss metal for nanoplasmonic and metamaterials applications,” Appl. Phys. Lett. 95(15), 151102 (2009). [CrossRef]  

14. M. G. Blaber, M. D. Arnold, and M. J. Ford, “A review of the optical properties of alloys and intermetallics for plasmonics,” J. Phys.: Condens. Matter 22(14), 143201 (2010). [CrossRef]  

15. R. T. Beach and R. W. Christy, “Electron-electron scattering in the intraband optical conductivity of Cu, Ag, and Au,” Phys. Rev. B 16(12), 5277–5284 (1977). [CrossRef]  

16. G. R. Parkins, W. E. Lawrence, and R. W. Christy, “Intraband optical conductivity σ (ω, T) of Cu, Ag, and Au: Contribution from electron-electron scattering,” Phys. Rev. B 23(12), 6408–6416 (1981). [CrossRef]  

17. Y. Wang, J. Yu, Y.-F. Mao, et al., “Stable, high-performance sodium-based plasmonic devices in the near infrared,” Nature 581(7809), 401–405 (2020). [CrossRef]  

18. J. Xu, T. Zhang, Y. Li, et al., “Room-temperature low-threshold plasmonic nanolaser through mode-tailoring at communication wavelengths,” Laser and Photonics Reviews 17(2), 2200322 (2023). [CrossRef]  

19. J. Ni, C. Huang, L.-M. Zhou, et al., “Multidimensional phase singularities in nanophotonics,” Science 374(6566), eabj0039 (2021). [CrossRef]  

20. W. Li, R. Liu, J. Li, et al., “Highly efficient single-exciton strong coupling with plasmons by lowering critical interaction strength at an exceptional point,” Phys. Rev. Lett. 130(14), 143601 (2023). [CrossRef]  

21. J. N. Hodgson, “The optical effective mass and absorption edge of electrons in sodium,” J. Phys. Chem. Solids 24(10), 1213–1216 (1963). [CrossRef]  

22. N. V. Smith, “Optical constants of sodium and potassium from 0.5 to 4.0 eV by split-beam ellipsometry,” Phys. Rev. 183(3), 634–644 (1969). [CrossRef]  

23. R. E. Palmer and S. E. Schnatterly, “Observation of surface plasmons and measurement of the optical constants for sodium and potassium,” Phys. Rev. B 4(8), 2329–2339 (1971). [CrossRef]  

24. J. Monin and G.-A. Boutry, “Optical and photoelectric properties of alkali metals,” Phys. Rev. B 9(4), 1309–1327 (1974). [CrossRef]  

25. T. Inagaki, L. C. Emerson, E. T. Arakawa, et al., “Optical properties of solid Na and Li between 0.6 and 3.8 eV,” Phys. Rev. B 13(6), 2305–2313 (1976). [CrossRef]  

26. M. Fox, Optical Properties of Solids, 2nd ed, Oxford Master Series in Condensed Matter Physics No. v. 3 (Oxford University Press, 2010).

27. M.-L. Thèye, “Investigation of the optical properties of Au by means of thin semitransparent Films,” Phys. Rev. B 2(8), 3060–3078 (1970). [CrossRef]  

28. P. B. Johnson and R. W. Christy, “Optical Constants of the Noble Metals,” Phys. Rev. B 6(12), 4370–4379 (1972). [CrossRef]  

29. R. L. Olmon, B. Slovick, T. W. Johnson, et al., “Optical dielectric function of gold,” Phys. Rev. B 86(23), 235147 (2012). [CrossRef]  

30. H. U. Yang, J. D’Archangel, M. L. Sundheimer, et al., “Optical dielectric function of silver,” Phys. Rev. B 91(23), 235137 (2015). [CrossRef]  

31. “RefractiveIndex INFO,” 2015, https://refractiveindex.info

Supplementary Material (1)

NameDescription
Supplement 1       Supporting_Information

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. The schematic diagram of the in-air ellipsometry measurement on the encapsulated surface. The light beams depicted in dark red are the measurement light beams needed to collect.
Fig. 2.
Fig. 2. Evaluation the validity of the AEM. The dielectric function of (a) Au and (b) Ag measured with and without covering fused silica respectively.
Fig. 3.
Fig. 3. (a) The ${\varepsilon _1}$ and (b)-(c) ${\varepsilon _2}$ of the dielectric functions of sodium from 0.50 eV to 4.97 eV (250-2500 nm). The measured data of the dielectric functions of sodium reported in the literature are shown for comparison. The statistical data of Drude model parameters (d) ${\varepsilon _\infty }$, (e) $\hbar {\omega _p}$, and (f) $\hbar \gamma $ are obtained from over 200 sample points.
Fig. 4.
Fig. 4. Errors of the AEM. (a) The schematic diagram of the error due to a surface layer. (b) The relative error of ε1 and (c) the relative error of ε2 of sodium samples covered with an 0.004-nm oxidation layer with a refractive index of 1.4, 1.5, and 1.6. (d) The schematic diagram of the error due to surface roughness-induced depolarization effect. (e) The relative error of ε1 and (f) the relative error of ε2 of sodium samples encapsulated by quartz substrates with a surface roughness of ∼100 pm. (g) The schematic diagram of the error due to depolarization effect resulting from substrate dispersion and focusing beam. The light beam depicted in dark red is a beam of a certain wavelength with an incident angle α, while the light beam indicated in light red is a beam of other wavelength with an incident angle α+θ. (h) The relative error of ε1 and (i) the relative error of ε2 of sodium samples caused by substrate dispersion and focusing beam.

Equations (7)

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

ε m ( ω ) = s i n 2 α [ 1 + t a n 2 β ( T ρ m ( ω ) T + ρ m ( ω ) ) 2 ] .
T = s i n 2 β ( c o s α + ε s u b ( ω ) c o s β ) 2 s i n 2 ( α + β ) c o s 2 ( α β ) .
ε s u b ( ω ) = s i n 2 θ 0 { 1 + t a n 2 θ 0 [ 1 ρ s u b ( ω ) 1 + ρ s u b ( ω ) ] 2 } ,
ε m ( ω ) = ε 1 ( ω ) + i ε 2 ( ω ) = ε ω p 2 ω ( ω + i γ ) .
ε 1 ( ω ) = ε ω p 2 ω 2 + γ 2 ε ω p 2 ω 2 ,
ε 2 ( ω ) = ω p 2 × γ ω ( ω 2 + γ 2 ) ω p 2 × γ ω 3 .
δ ε ε = ε α δ θ ε × 100 % .
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.