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

Fabrication of broadband anti-reflective layers by mask-free etching TiO2 films

Open Access Open Access

Abstract

We present a simple way to make TiO2 anti-reflective layers on top of silicon substrates. Surfaces of TiO2 films have been modified by radio frequency plasma with CF4 as an etchant. Mask-free etching process on the polycrystalline films leads to the formation of random sub-wavelength textures. The reflection of the etched samples are significantly suppressed in the wavelength range of 400~800 nm (2.9~4.6%, 3% compared with 34% on bare silicon at the wavelength of 600 nm). We have numerically simulated the optical properties of TiO2 layers using the finite-difference time-domain method. The anti-reflective effects are attributed to random roughness on TiO2 surfaces. The etching porcess increases the surface roughness, therefore, the gradient of refractive index between air and silicon substrate is reduced. As a result, the Fresnel reflection is supressed. Our results demonstrate an efficient way of anti-reflective coating for solar cells.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Study of solar cells has become increasingly important due to the upcoming energy shortage and various environmental problems [1,2]. When making these devices, special care must be taken to reduce the Fresnel reflection occurring at the interface between air and absorbing layers [3]. For instance, over 30 percent of the energy impinging on bare silicon surface is reflected and wasted because of the high refractive index of silicon. Traditional anti-reflective (AR) coatings consist of alternative layers with different refractive indices [4]. Reflection at different interfaces within the AR films interfere destructively when the optical thickness of each layer is λ/4. These kinds of AR layers usually work in narrow bands which cover only small portions of the solar spectrum. The optimization of AR performance involves precise control of refractive index and thickness of each layer of the films [5,6].

An alternative for reducing reflection is sub-wavelength structures. Using methods such as nanoimprinting [7,8] and nanosphere lithography [9–11], various structures have been fabricated for AR coatings. Sub-wavelength pyramidal and honeycomb structures were fabricated on silicon substrates with colloidal lithography followed by a reactive ion etching process. Reflection less than 1.5% were obtained with optimal structural parameters [9]. Inspired by the structures in the corneas of nocturnal moths [12], sub-wavelength nipple arrays were directly patterned on silicon using spin-coated silica colloidal monolayers as etching masks [13]. The resulted textures showed excellent AR properties over a wide range of wavelenghs. Sub-wavelength structures can also be patterned on transparent substrates. Using close packed nanoparticle arrays as masks, plasma etching was performed to generate nanopillar arrays on quartz substrates. Effects of structure parameters on AR performance were analyzed in detail [14]. Numerical methods such as rigorous coupled-wave analysis (RCWA), finite element method (FEM), and finite-difference time domain (FDTD) method [9,15] are often used to study the optical properties of these structures.

The main idea of sub-wavelength structures is to generate a graded refractive index between air and substrates, therefore, Fresnel reflection can be supressed [16,17]. From this point of view, random structures can also be used for AR coatings as long as the structures are sub-wavelength in scale [18].Black silicon can be fabricated with different etching process with or without masks [19]. The nano-structures on the silicon surface effectively suppress reflection, so the silicon substrates appear black.

Besides directly made on substrates, sub-wavelength structures can also be fabricated on additional coating layers. Using a square array of TiO2 nano-cylinders, an average reflectivity (weighed over the AM1.5 solar spectrum in the 420~980 nm spectral range) of 2.8% was demonstrated [20]. With carefully controlled synthesis process, TiO2 nano rods dramatically reduced the reflection of silicon substrates [21]. Along with suppressing reflection, AR layers may have additional functionalities such as passivation [7] and self cleaning [22,23].

In this paper, we present a simple method to make TiO2 AR layers on top of silicon substrates. We use radio frequency plasma to etch the surfaces of TiO2 films and get wage-like textures. The etched films can effectively suppress the reflection to 2.9~4.6% at the wavelength range of 400~800 nm. We analyze the optical properties with a 3D FDTD method, and attribute the enhanced AR effect to increased roughness of the TiO2 surfaces. TiO2 is one of the most widely used materials for AR coatings due to its properties such as high refractive index, excellent visible and near infrared transmittance, chemical resistant and stable, good mechanical properties, non-toxic, and low cost [24–26]. High efficient and low cost methods of making TiO2 AR layers can be very helpful to the development of solar cells.

2. Experiments

The experiments started with commercial polycrystalline TiO2 films [27]. The films were prepared on silicon wafers by the Chemical Vapor Deposition (CVD) method. They were well crystallized in the anatase form. The surface morphology was examined by scanning electron micrographs (SEM, Hitachi S-4500). It is clear that the film is composed of well faceted nanocrystals that are stacked in random directions [Fig. 1(a)]. The microcrystals are hundreds of nanometers in size. The thickness of the film is about 800 ~ 900 nm [Fig. 1(c)].

 figure: Fig. 1

Fig. 1 Micrographs of the TiO2 films: (a) top view of the unetched film, (b) top view of the etched film, c) cross section of the unetched film, (d) cross section of the etched film.

Download Full Size | PDF

We used CF4 plasma to etch the TiO2 film to modify its surface morphology. This method has been proved to be an efficient way for roughening surfaces of TiO2 films [27]. After 40s etching, we got a structure shown in Figs. 1(b) and 1(d). Wedge like structures are formed in the process. Some wedges may stack together and form wedge arrays, but both the sizes and directions of the wedges are generally random [Fig. 1(b)]. The typical lateral size of a single wedge from the top view is about 300 nm*80 nm, [Fig. 1(b)]. The average height of the wedges is around 600 nm [Fig. 1(d)], which is a little smaller than the original thickness of the film.

We measured the reflection with a UV-Vis spectrometer (Shimadzu UV-2450) attached with an integrating sphere (ISR-2200). The schematic of the configuration inside the integrating sphere is shown in Fig. 2(a). The measurement beam impinges on the sample at an 8 ° incident angle. Both of the specular and diffuse reflection are collected. The measured results are shown in Fig. 2. In the wavelength range of 400 ~ 800 nm, the reflection of a bare silicon substrate is above 30% (blue dashed line). The reflection of a TiO2 film as purchased varies from 3% to 8%. The oscillations in the reflection are due to the interference between the reflection at air/TiO2 and TiO2/Si interfaces (red dotted line). After 40 s etching process, the reflection of the TiO2 film droped to 2.9~4.6% at the wavelength range of 400~800 nm, meanwhile, the oscillations in the reflection no longer exist (black solid line). We can tell from the black surface of the etched film that the incident light is trapped and absorbed [Inset of Fig. 2].

 figure: Fig. 2

Fig. 2 (a) Schematic of the configuration inside the integrating shpere. (b) Measured reflection of bare a Si substrate, unetched TiO2 film, and etched TiO2 film. Inset shows the photograph of etched TiO2 film.

Download Full Size | PDF

3. Numerical simulation

Qualitatively, the AR effects can be attributed to the graded refractive index. Since the lateral size of the wedge-like structures are smaller than wavelengths in the 400~800 nm region (especially, the width of the structures are only around 80nm, well below the wavelengths), the TiO2 film can be evaluated with the effective medium theory [13,15]. From the top of TiO2 film to the bottom, the ratio of TiO2/air increases continuously. As a result, there is graded refractive index inside the TiO2 layer.

In order to understand the underlying physics of the experimental results, we numerically simulated the optical properties with the finite-difference time-domain (FDTD) method (Lumerical FDTD Solutions). The configuration of the simulation is demonstrated in Fig. 3. A plane wave source with a wavelength range of 400~800 nm is placed 1200 nm above the top surface of silicon substrate. A reflection monitor (frequency domain filed and power monitor) is placed 100 nm above the source. The plane wave propagates downwards to the AR structure, so that both the specular and diffuse reflection are collected by the reflection monitor. A field profile monitor (frequency domain filed and power monitor) is placed perpendicular to the surface of substrate. The electric field profiles in the cross sections are recorded in the frequency domain.

 figure: Fig. 3

Fig. 3 Schematic of the configuration in FDTD simulations

Download Full Size | PDF

We constructed 4 models of TiO2 layers [Fig. 4] on silicon substrates and calculated the corresponding reflection. The refractive index of TiO2 in the simulation follows the dispersion formula [28]:

n21=4.6796λ2λ20.20021482.

Therefore, the refractive index varies from 2.45 to 2.69 in the wavelength range of 400~800 nm. We believe that the key factor of the AR effects is the roughness of the film surfaces, so random roughness is used to simulate the textures of the TiO2 film surfaces [Fig. 1]. The roughness used in the simulation is characterized by the root-mean-squre (rms) roughness (σrms) and correlation length (lc). The correlation function is

H(r)H(r+δ)=σrms 2exp((δlc)2).

The simulated reflection of the bare silicon substrate [Fig. 4(a)] is above 30% (blue dashed lines in Fig. 5), which is in accordance with the measured result [blue dashed line in Fig. 2(b)]. When the silicon substrate is covered by a flat TiO2 film [Fig. 4(b)], we can expect AR effect because the refractive index of TiO2 is between the refractive indices of air and silicon. Following the conventional strategy of single-layer AR coating, the thickness of the TiO2 layer is set to be 60 nm, so that the optical length is nL = λ/4 at the wavelength of 600 nm. The calculated reflection is shown with cayan dot-dashed line in Fig. 5(a). The minimum reflection is about 5% at the wavelength of 600 nm, while the reflection increases at other wavelengths, such as 38% at the wavelength of 400 nm. The calculated reflection is similar to the reported experimental results [29].

 figure: Fig. 4

Fig. 4 Illustration of the sub-wavelength structures used in numerical sumulation, (a) bare Si substrate, (b) flat TiO2 film, (c) TiO2 film with roughness of rms=80 nm, (d) TiO2 film with roughness of rms=200 nm.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 Comparison between the measured and simulated reflection of TiO2 films, (a) film as purchased, (b) film etched for 40 s by CF4 plasma.

Download Full Size | PDF

The AR effect enhances when roughness is introduced to the surface of the film. We used σrms = 80 nm and lc = 100 nm [Eq. 2] to generate random roughness shown in Fig. 4(c). In this case, the subwavelengh textures on the surface generate gradually varing effective rafractive index [16] in the vertical direction, so that the reflection can be further reduced. The simulation result is shown with the black solid line in Fig. 5(a). The reflection reduces to below 10%, which is in accordance with the experimental result of the unetched film [red dotted line].

We increased the rougness of the surface to σrms = 200 nm, and got a rougher film shown in Fig. 4(d). The average thickness of the film is also reduced to 400 nm. This model is used to simulate the roughened and thinner film after the etching process [Figs. 1(b) and 1(d)]. The simulated result is shown with the black solid line in Fig. 5(b). The calculated reflection is round 4%, which is in accordance with the experimental result of the etched film [red dotted line]. The gradient of the effective refractive index reduces with the increase of the roughness. That leads to the improvement of AR effect.

Using numerical simulation, we can see the effects of the surface roughness on the propagation of light [Fig. 6]. The top surface of silicon substrate was defined as the z = 0 nm position, and the plane wave source propagating along the −z direction was placed at z = 1200 nm. The field monitor is 2-D Y-normal and placed at the position of y=500 nm [Fig. 4]. The electric field (|E|2) distributions were calculated at the frequency of 500 THz (which corresponds to the wavelength of about 600 nm in the air). The electric field intensity above the source is contributed by the reflection. Without TiO2 layer, there is considerable intensity above the air/Si interface owing to the high reflection of the bare silicon substrate [Fig. 6(a)]. With a smooth TiO2 film, the reflection is reduced, that leads to less electric field intensity above the air/TiO2 interface [Fig. 6(b)]. When roughness is introduced to the TiO2 films, we can hardly see any electric field intensity above the air/TiO2 interface [Figs. 6(c) and 6(d)]. Because of the AR effect generated by the rough films, we can also see more electric field intensity in the silicon substrates. The sub wavelength strucures in the rough films induce modifications of the electric field distribution, and the field intensity is no longer uniform along the x direction.

 figure: Fig. 6

Fig. 6 Simulated Electric field distribution (|E|2) in the cross sections near the surfaces, (a) bare Si substrate, (b) flat TiO2 film, (c) TiO2 film with rms= 80 nm (d) TiO2 film with rms=200 nm.

Download Full Size | PDF

The simulated results are in good accordance to the experimental results, the remaining differences between them can be attributed to the following reasons: owing to the limitation of computing power, we generated random structure in a small area (1 µm2) then used periodic boundary conditions; there are fine structures at the surface of the microcrystals and wedge like structures in TiO2 films [Fig. 1], which can further modify the near field optical properties.

4. Discussion

We believe that the method presented here is helpful to improve the performance of silicon solar cells. Firstly, the reflection is suppressed to below 4.6%, while the reflection of a conventional textured silicon surface is above 10% [25,30]. The etched film exhibits AR effects in the whole visible band, while conventional single-layer AR coatings, such SiN [31], TiO2, and ITO [32], can minimize reflection in a narrow band. Secondly, excellent AR effects can be obtained with elaborate structures such as moth-eye structures [9], black silicon [19], and TiO2 arrays [20], however, the fabrication processes usually involve multi-step procedures, and the parameters need to be precisely controlled. Mask-free dry etching is relatively easy to implement. Thirdly, textured surfaces of substrates may have more charge recombination sites and need additional passivation [33], while the AR coating in this paper does not destroy the silicon surface. Fourthly, CVD and dry etching are widely used industrial methods, so we may expect high throughput and process speed. Meanwhile, the cost of TiO2 AR coating is relatively low compared to materials such as ITO. The method presented here can be integrated into the current fabrication process of silicon solar cells. The idea for generating graded refractive index can even be applied to already finished solar cells. Finally, in standard commercial photovoltaic modules, solar cells are commonly encapsulated with polymers such as ethylene vinyl acetate (EVA) [34]. The wedge-like structure may maintain their shapes after encapsulation. Moreover, because the high refractive index of TiO2, the encapsulated structures may still generate graded refractive index. However, owing to the modification of environmental refractive index, more study is needed to optimize the nanostructures to keep the excellent AR effects. The method presented in this paper may also be applied in optical devices such as LEDs and detectors.

5. Conclusion

We have investigated the AR effects of TiO2 films on top of silicon substrates. The polycrystalline TiO2 films were roughened by an mask-free ething process, then exhibited broad band AR properies. We numerically analyzed the reflection of the films with the FDTD method, and showed that the AR performance can be attributed to random roughness. TiO2 layers reduce the contrast of refractive index between air and silicon substrates. The rough surfaces of TiO2 films further reduce the gradient of the refractive index and improve AR performance.

Funding

National Natural Science Foundation of China (91233204, 51372036); The 111 Project, China (B13013); Open Project of Key Laboratory of Infrared System Detection and Imaging Technology, Shanghai Institute of Technical Physics, Chinese Academy of Sciences; Japan Science and Technology Agency (JST).

References

1. A. Polman, M. Knight, E. C. Garnett, B. Ehrler, and W. C. Sinke, “Photovoltaic materials: Present efficiencies and future challenges,” Science 352, aad4424 (2016). [CrossRef]   [PubMed]  

2. N.-G. Park, “Perovskite solar cells: an emerging photovoltaic technology”, Mater. Today 18, 65–72 (2015). [CrossRef]  

3. H. K. Raut, V. A. Ganesh, A. S. Nair, and S. Ramakrishna, “Anti-reflective coatings: A critical, in-depth review,” Energy & Environ. Sci. 4, 3779–3804 (2011). [CrossRef]  

4. M. Victoria, C. Domínguez, I. Antón, and G. Sala, “Antireflective coatings for multijunction solar cells under wide-angle ray bundles,” Opt. Express 20, 8136–8147 (2012). [CrossRef]   [PubMed]  

5. N.-F. Wang, T.-W. Kuo, Y.-Z. Tsai, S.-X. Lin, P.-K. Hung, C.-L. Lin, and M.-P. Houng, “Porous SiO2/MgF2 broadband antireflection coatings for superstrate-type silicon-based tandem cells,” Opt. Express 20, 7445–7453 (2012). [CrossRef]   [PubMed]  

6. X. Yan, D. J. Poxson, J. Cho, R. E. Welser, A. K. Sood, J. K. Kim, and E. F. Schubert, “Enhanced Omnidirectional Photovoltaic Performance of Solar Cells Using Multiple-Discrete-Layer Tailored- and Low-Refractive Index Anti-Reflection Coatings,” Adv. Fun. Mater. 23, 583–590 (2013). [CrossRef]  

7. J. Barbé, A. F. Thomson, E.-C. Wang, K. McIntosh, and K. Catchpole, “Nanoimprinted TiO2 sol-gel passivating diffraction gratings for solar cell applications,” Prog. Photovoltaics: Res. Appl. 20, 143–148 (2012). [CrossRef]  

8. S. M. Kang, S. Jang, J.-K. Lee, J. Yoon, D.-E. Yoo, J.-W. Lee, M. Choi, and N.-G. Park, “Moth-Eye TiO2 Layer for Improving Light Harvesting Efficiency in Perovskite Solar Cells,” Small 12, 2443–2449 (2016). [CrossRef]   [PubMed]  

9. H. L. Chen, S. Y. Chuang, C. H. Lin, and Y. H. Lin, “Using colloidal lithography to fabricate and optimize sub-wavelength pyramidal and honeycomb structures in solar cells,” Opt. Express 15, 14793–14803 (2007). [CrossRef]   [PubMed]  

10. C. K. Huang, K. W. Sun, and W.-L. Chang, “Efficiency enhancement of silicon solar cells using a nano-scale honeycomb broadband anti-reflection structure,” Opt. Express 20, A85–A93 (2012). [CrossRef]   [PubMed]  

11. X. Wang, Z. Yang, P. Gao, X. Yang, S. Zhou, D. Wang, M. Liao, P. Liu, Z. Liu, S. Wu, J. Ye, and T. Yu, “Improved optical absorption in visible wavelength range for silicon solar cells via texturing with nanopyramid arrays,” Opt. Express 25, 10464–10472 (2017). [CrossRef]   [PubMed]  

12. D. Stavenga, S. Foletti, G. Palasantzas, and K. Arikawa, “Light on the moth-eye corneal nipple array of butterflies,” Proc. R. Soc. London, Ser. B 273, 661–667 (2006). [CrossRef]  

13. C.-H. Sun, P. Jiang, and B. Jiang, “Broadband moth-eye antireflection coatings on silicon,” Appl. Phys. Lett. 92, 061112 (2008). [CrossRef]  

14. S. Ji, K. Song, T. B. Nguyen, N. Kim, and H. Lim, “Optimal Moth Eye Nanostructure Array on Transparent Glass Towards Broadband Antireflection,” ACS Appl. Mater. Interfaces 5, 10731–10737 (2013). [CrossRef]   [PubMed]  

15. K. Han and C.-H. Chang, “Numerical Modeling of Sub-Wavelength Anti-Reflective Structures for Solar Module Applications,” Nanomaterials 4, 87–128 (2014). [CrossRef]   [PubMed]  

16. S. Chattopadhyay, Y. Huang, Y. Jen, A. Ganguly, K. Chen, and L. Chen, “Anti-reflecting and photonic nanostructures,” Mater. Sci. Eng. R: Reports 69, 1–35 (2010). [CrossRef]  

17. D. Hu, D. Liu, J. Zhang, L. Wu, and W. Li, “Preparation and stability study of broadband anti-reflection coatings and application research for CdTe solar cell,” Opt. Mater. 77, 132 – 139 (2018). [CrossRef]  

18. D. A. Boyd, J. A. Frantz, S. S. Bayya, L. E. Busse, W. Kim, I. Aggarwal, M. Poutous, and J. S. Sanghera, “Modification of nanostructured fused silica for use as superhydrophobic, IR-transmissive, anti-reflective surfaces,” Opt. Mater. 54, 195–199 (2016). [CrossRef]  

19. X. Liu, P. R. Coxon, M. Peters, B. Hoex, J. M. Cole, and D. J. Fray, “Black silicon: fabrication methods, properties and solar energy applications,” Energy Environ. Sci. 7, 3223–3263 (2014). [CrossRef]  

20. P. Spinelli, B. Macco, M. A. Verschuuren, W. M. M. Kessels, and A. Polman, “Al2O3/TiO2 nano-pattern antireflection coating with ultralow surface recombination,” Appl. Phys. Lett. 102, 233902 (2013). [CrossRef]  

21. D. M. Andoshe, S. Choi, Y.-S. Shim, S. H. Lee, Y. Kim, C. W. Moon, D. H. Kim, S. Y. Lee, T. Kim, H. K. Park, M. G. Lee, J.-M. Jeon, K. T. Nam, M. Kim, J. K. Kim, J. Oh, and H. W. Jang, “A wafer-scale antireflective protection layer of solution-processed TiO2 nanorods for high performance silicon-based water splitting photocathodes,” J. Mater. Chem. A 4, 9477–9485 (2016). [CrossRef]  

22. K. Nakata, M. Sakai, T. Ochiai, T. Murakami, K. Takagi, and A. Fujishima, “Antireflection and self-cleaning properties of a Moth-Eye-Like Surface Coated with TiO2 Particles,” Langmuir 27, 3275–3278 (2011). [CrossRef]   [PubMed]  

23. X.-T. Zhang, O. Sato, M. Taguchi, Y. Einaga, T. Murakami, and A. Fujishima, “Self-cleaning particle coating with antireflection properties,” Chem. Mater. 17, 696–700 (2005). [CrossRef]  

24. G. S. Vicente, A. Morales, and M. Gutierrez, “Preparation and characterization of sol-gel TiO2 antireflective coatings for silicon,” Thin Solid Films 391, 133 – 137 (2001). [CrossRef]  

25. J. Višniakov, A. Janulevičius, A. Maneikis, I. Matulaitienė, A. Selskis, S. Stanionytė, and A. Suchodolskis, “Antireflection TiO2 coatings on textured surface grown by HiPIMS,” Thin Solid Films 628, 190–195 (2017). [CrossRef]  

26. B. S. Richards, “Comparison of TiO2 and other dielectric coatings for buried-contact solar cells: a review,” Prog. Photovoltaics: Res. Appl. 12, 253–281 (2004). [CrossRef]  

27. X. Zhang, M. Jin, Z. Liu, S. Nishimoto, H. Saito, T. Murakami, and A. Fujishima, “Preparation and Photocatalytic Wettability Conversion of TiO2 -Based Superhydrophobic Surfaces,” Langmuir 22, 9477–9479 (2006). [CrossRef]   [PubMed]  

28. M. N. Polyanskiy, “Refractive index database,” https://refractiveindex.info. Accessed on 2018-10-11.

29. E. Shi, L. Zhang, Z. Li, P. Li, Y. Shang, Y. Jia, J. Wei, K. Wang, H. Zhu, D. Wu, S. Zhang, and A. Cao, “TiO2-Coated Carbon Nanotube-Silicon Solar Cells with Efficiency of 15%,” Sci. Reports 2884 (2012).

30. D. Z. Dimitrov and C.-H. Du, “Crystalline silicon solar cells with micro/nano texture,” Appl. Surf. Sci. 266, 1–4 (2013). [CrossRef]  

31. A. F. Braña, H. Gupta, R. K. Bommali, P. Srivastava, S. Ghosh, and R. P. Casero, “Enhancing efficiency of c-si solar cell by coating nano structured silicon rich silicon nitride films,” Thin Solid Films 662, 21 – 26 (2018). [CrossRef]  

32. K. Ali, S. A. Khan, and M. Z. M. Jafri, “Structural and optical properties of ITO/TiO2 anti-reflective films for solar cell applications,” Nanoscale Res. Lett. 9, 175 (2014). [CrossRef]   [PubMed]  

33. Y. Yu, Z. Zhang, X. Yin, A. Kvit, Q. Liao, Z. Kang, X. Yan, Y. Zhang, and X. Wang, “Enhanced photoelectrochemical efficiency and stability using a conformal TiO2 film on a black silicon photoanode,” Nat. Energy 2, 17045 (2017). [CrossRef]  

34. A. Skoczek, T. Sample, and E. D. Dunlop, “The results of performance measurements of field-aged crystalline silicon photovoltaic modules,” Prog. Photovoltaics: Res. Appl. 17, 227–240 (2008). [CrossRef]  

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 (6)

Fig. 1
Fig. 1 Micrographs of the TiO2 films: (a) top view of the unetched film, (b) top view of the etched film, c) cross section of the unetched film, (d) cross section of the etched film.
Fig. 2
Fig. 2 (a) Schematic of the configuration inside the integrating shpere. (b) Measured reflection of bare a Si substrate, unetched TiO2 film, and etched TiO2 film. Inset shows the photograph of etched TiO2 film.
Fig. 3
Fig. 3 Schematic of the configuration in FDTD simulations
Fig. 4
Fig. 4 Illustration of the sub-wavelength structures used in numerical sumulation, (a) bare Si substrate, (b) flat TiO2 film, (c) TiO2 film with roughness of rms=80 nm, (d) TiO2 film with roughness of rms=200 nm.
Fig. 5
Fig. 5 Comparison between the measured and simulated reflection of TiO2 films, (a) film as purchased, (b) film etched for 40 s by CF4 plasma.
Fig. 6
Fig. 6 Simulated Electric field distribution (|E|2) in the cross sections near the surfaces, (a) bare Si substrate, (b) flat TiO2 film, (c) TiO2 film with rms= 80 nm (d) TiO2 film with rms=200 nm.

Equations (2)

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

n 2 1 = 4.6796 λ 2 λ 2 0.2002148 2 .
H ( r ) H ( r + δ ) = σ r m s   2 exp ( ( δ l c ) 2 ) .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.