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

Ultrathin planar hematite film for solar photoelectrochemical water splitting

Open Access Open Access

Abstract

Hematite holds promise for photoelectrochemical (PEC) water splitting due to its stability, low-cost, abundance and appropriate bandgap. However, it suffers from a mismatch between the hole diffusion length and light penetration length. We have theoretically designed and characterized an ultrathin planar hematite/silver nanohole array/silver substrate photoanode. Due to the supported destructive interference and surface plasmon resonance, photons are efficiently absorbed in an ultrathin hematite film. Compared with ultrathin hematite photoanodes with nanophotonic structures, this photoanode has comparable photon absorption but with intrinsically lower recombination losses due to its planar structure and promises to exceed the state-of-the-art photocurrent of hematite photoanodes.

© 2015 Optical Society of America

1. Introduction

Photoelectrochemical (PEC) water splitting is an effective method to convert solar energy to chemical energy that can be stored and used when desired. Hematite (α-Fe2O3) is a promising photoanode material for PEC water splitting due to its stability (in water and electrolyte), low-cost, abundance and near optimal band gap of 2.1 eV [1]. However, hematite suffers from a mismatch between the hole diffusion length (~20 nm) and the light penetration length (~120 nm at 550 nm) [1]. Only holes generated in the proximity of the hematite/water interface (also called the semiconductor liquid junction (SCLJ)) can contribute to water splitting, the rest is lost through recombination with photogenerated electrons. The reported photocurrents produced by hematite photoanodes are far lower than the theoretical maximum of 12.6 mA cm−2 [2] because of the significant recombination loss associated with the discrepancy of the light absorption and the hole diffusion length scales.

Many studies have tried to overcome the challenge by shortening the distance a hole needs to travel and enhancing photon absorption near the SCLJ. One route is to use hematite films with thicknesses on the order of the wavelength of incident sunlight and feature sizes on the order of 101 nm, including random nanoparticle aggregations [3, 4], ordered cauliflower [5], wormlike [6], nanowire [7] and nanorod [8, 9] structures. These thick films have near total photon absorption. However, these structures do not have well-engineered geometries and morphologies for hole transport [1, 10], hence, hematite photoanodes with these structures still exhibit excessive bulk recombination losses [11].

The other route is to use ultrathin hematite films where all of the hematite is in close proximity of the SCLJ. To enhance the photon absorption in such ultrathin films, one approach is to employ nanophotonic structures. Wang et al. [12] and Qiu et al. [10] coated ultrathin hematite films on fluorine-doped tin oxide (FTO) two-dimensional nanophotonic crystals. These host-guest structures have a graded refractive index that can reduce the total reflectivity and thus, increase the photon absorption in hematite films. Kim et al. [13] illustrated a hematite photoanode with a nanobeam photonic structure that supports Mie resonances to enhance the photons absorbed near the SCLJ. However, these nanophotonic structures have both high surface and interfacial areas between the hematite layer and electron collection layer which increases the amount of surface and interface recombination [14, 15]. In addition, nanopatterning typically requires chemical etching, laser ablation or catalytic vapor deposition, which creates surface defects [16]. These factors increase recombination possibilities in nanophotonic structures. Therefore, studies have also sought to use ultrathin planar hematite films to reduce both bulk and surface recombination losses. To enhance the photon absorption in the ultrathin planar hematite film, surface plasmon resonances supported by noble metal nanoparticles [17, 18] have been suggested. However, these ultrathin planar hematite films have shown reduced photon absorption and thus lower photocurrents compared to hematite films with nanophotonic structures [10]. For example, Iandolo et al. [18] reported a photocurrent of only 0.38 mA cm−2 at 1.6 V against the reversible hydrogen electrode (1.6 VRHE). Recently, Dotan et al. [11] coated a 26 nm thick hematite film on a silver-gold alloy substrate and achieved a photocurrent of 3.02 mA cm−2 at 1.6 VRHE. The photon absorption in the ultrathin planar hematite film is enhanced due to the destructive interference between the hematite film and the metallic substrate. The interference can be realized in a film much thinner than a quarter wavelength attributed to non-trivial phase shifts of propagating electromagnetic waves at the water/hematite and hematite/metallic substrate interfaces [11, 19]. They then combined this effect with a V-shape cell configuration to achieve a photocurrent of 3.98 mA cm−2 at 1.6 VRHE which is the state-of-the-art photocurrent for ultrathin planar hematite photoanodes. Nevertheless, this photocurrent is still lower than the best reported photocurrent of hematite photoanodes with nanophotonic structures (4.36 mA cm−2 at 1.6 VRHE, reported by Qiu et al. [10]).

Here, we first quantitatively investigate the spectral absorption feature of ultrathin hematite films on various metallic substrates. We show that destructive interference resonances in ultrathin hematite films can only be achieved in short wavelengths (below 450 nm). Then we propose an ultrathin planar hematite/silver nanohole array/silver substrate structure for solar photoelectrochemical water splitting. We theoretically show that this structure has both high broadband photon absorption comparable to nanophotonic structures due to supported destructive interference and surface plasmon resonance, and lower recombination loss, characteristic of planar ultrathin films. Thus, this structure promises to exceed the state-of-the-art performance of ultrathin hematite photoanodes.

2. Methods

To harness a broad band of solar energy and reduce recombination losses, we propose the hematite/metal nanohole array/metal substrate planar structure shown in Fig. 1(a). Figure 1(b) shows the optical constants (m=n+iκ, n is the refractive index, κ is the absorption index) of hematite from 300 to 590 nm (300 nm corresponds to the cut-on wavelength of solar spectrum, 590 nm corresponds to the hematite band gap of 2.1 eV) [11]. Hematite is absorbing below 450 nm, while nearly lossless above 450 nm (κis less than one tenth of n). The photon absorption in an ultrathin hematite film can be enhanced below 450 nm (and approach unity at some wavelengths) through the destructive interference that exists in ultrathin absorbing dielectric/metal structures, as recently described by Dotan et al. [11] and Kats et al. [19]. Above 450 nm, where hematite/metal does not support this interference resonance, the surface plasmon resonance supported by the metal nanohole array can be coupled into the hematite film to enhance the photon absorption.

 figure: Fig. 1

Fig. 1 Ultrathin planar hematite/metal nanohole array/metal substrate structure for solar PEC water splitting. (a) Layer 1 is water, layer 2 is hematite, layer 3 is dielectric (refractive index n = 1.4) filled metal nanohole array with hexagonal lattice and layer 4 is metal substrate. (b) Optical constants (m=n+iκ) of hematite.

Download Full Size | PDF

To quantitatively optimize the proposed structure, we used the photocurrent as a figure of merit. We assumed that only photons absorbed within 20 nm distance from the SCLJ can contribute to the photocurrent and that each of these photons can generate one electron/hole pair without any recombination loss. Therefore, the photocurrent is expressed as

Jph=e300nm590nmα(λ)I(λ)hc/λdλ
where e is the unit charge, α is the photon absorption in the hematite film within 20 nm of the SCLJ, I is the ASTM Air Mass 1.5 solar spectrum, h is Planck’s constant, c is the speed of light and λ is the wavelength. We do not aim to predict photocurrents exactly, but use it to evaluate and compare optical designs. Although the assumption is simplified, this model for photocurrent (Eq. (1)) was accepted by many studies [12, 13] and hence, is useful to fairly compare the functional optical properties of our structures with those in literature.

We also calculated the solar-averaged absorption defined as

αsolar=300nm590nmα(λ)I(λ)dλ300nm590nmI(λ)dλ
where α is the photon absorption in the hematite film within 20 nm of the SCLJ as defined before. Equations (1) and (2) show that photocurrent and solar-averaged absorption are positively related because I(λ) increases with λ from 300 to 590 nm. Therefore, the solar-averaged absorption can also be used as a figure of merit and was used for more information later.

We determined α(λ) using the finite difference time domain (FDTD) method [20]. For the nanohole array, we used a hexagonal lattice unit cell and periodic boundary conditions (along x and y).

3. Results and discussion

First, we calculated how the photocurrent varies with hematite thicknesses in a tri-layer structure (shown in the inset of Fig. 2(a)). For the substrate (layer 3), we considered gold (Au), silver (Ag) and aluminum (Al) because they are reflective and can support a surface plasmon resonance in the 300 to 590 nm wavelength range. As shown in Fig. 2(a), the optimal hematite thickness on Au and Ag substrates is 20 nm, while on Al it is 30 nm. We also observed that the photoanode with an Ag substrate produces the best photocurrent, while photoanodes with Al and Au substrates are second and third best, respectively. This is further verified by the solar-averaged absorption variations with hematite thicknesses shown in Fig. 2(b).

 figure: Fig. 2

Fig. 2 Photocurrent and photon absorption of hematite films on Au, Al and Ag substrates. (a) Photocurrent and (b) solar-averaged absorption variations with hematite film thicknesses. Spectral total absorption, αtot, and the spectral absorption within 20 nm of the SCLJ, α, for hematite films on (c) Au, (d) Al and (e) Ag substrates. (f) Phasor diagrams of the reflected partial waves at 400 and 550 nm for 20 nm hematite/Ag structure.

Download Full Size | PDF

To better understand how metal substrates and hematite thicknesses impact destructive interference and spectral absorption, we calculated the spectral total absorption, αtot(λ), (in both the hematite and the substrate) and the spectral absorption within 20 nm of the SCLJ for various substrate choices and hematite thicknesses. In the 20 nm hematite/Au structure, strong interference resonances occur below 470 nm (indicated by high total absorption) as shown in Fig. 2(c). However, the Au substrate absorbs a large portion of photons owing to the presence of the interband transition at 470 nm, which reduces the photon absorption within 20 nm of the SCLJ (the absorption in hematite in this case). Therefore, the photocurrent is relatively low. In contrast, most photons are absorbed by the hematite layer in structures with Al and Ag substrates as shown in Figs. 2(d) and 2(e), and consequently these structures reach higher optimal photocurrents.

Figure 2(d) shows that the resonant wavelength (indicated by the location of peak total absorption) is shifted to longer wavelengths with thicker hematite layers. This means photon absorption at longer wavelengths (closer to the band gap), which is desirable, can be enhanced by increasing the hematite thickness. However, the absorption is spread over the entire thickness, such that a lower fraction within 20 nm of the SCLJ is collected. In the case with the Al substrate, the photon absorption within 20 nm of the SCLJ increases at long wavelengths while decreases at short wavelengths with increasing hematite thicknesses. Therefore, an optimal hematite thickness exists by balancing the absorption over the entire spectrum. The optimal hematite thickness is larger on the Al substrate (30 nm) than on the Ag substrate (20 nm), hence, photon absorption within 20 nm of the SCLJ is lower for the hematite/Al structure than for hematite/Ag structure. Comparing Figs. 2(c)–2(e), the hematite/Ag structure clearly achieves the highest overall useful photon current generation of the three structures.

Nevertheless, the 20 nm hematite/Ag structure still absorbs poorly at longer wavelengths (near the bandgap) as shown in Fig. 2(e), because the destructive interference can only be realized over a narrow spectrum (which is also the situation for two other substrates). This is further verified by the phasor diagrams of the reflected partial waves at 400 and 550 nm for this structure shown in Fig. 2(f). The total reflection coefficient can be expressed in the form of partial waves as,

r=l=0rl
where r0=r12, rl=t12r23lr21l1t21exp(4πm2h2li/λ) for l>0, rpq=(mpmq)/(mp+mq), tpq=2mp/(mp+mq), and m2 and h2 is the optical constants and thickness of layer 2. The phasor of the first partial wave,r0, begins at the origin; the second phasor, r1, begins at the end of r0, etc. The total reflection coefficient is the final value of the phasor trajectory in the complex plane. Only the first three phasors were calculated since the rest were negligible.We observe that the phasor trajectory returns sharply after the first partial wave, r0, and thus, the sum of the secondary partial waves partially cancel r0 at 400 nm indicating the existence of destructive interference. In contrast, the first three partial waves all move away from the origin indicating no supported destructive interference and thus, weak absorption at 550 nm.

To enhance the absorption above 450 nm, where hematite/metal does not support destructive interference, we introduced an array of nanoholes in the metal substrate to support a surface plasmon resonance that couples into the hematite film.

We optimized the height, radius and lattice constant of the Ag nanohole array using the particle swarm optimization technique [21]. The thickness of the hematite was kept constant at 20 nm. To make the structure more practical and compatible with common fabrication techniques, the nanohole areas are assumed to be filled with a dielectric material having refractive index of 1.4 (approximately corresponding to the refractive index of polydimethylsiloxane or glass). Although the contact area between the hematite layer and the metal substrate is reduced, the destructive effects on the photocurrent are expected to be minimal [17, 22, 23]. Due to the change from a solid Ag substrate to a dielectric-filled Ag nanohole array, the interference condition is slightly altered and the photon absorption below 450 nm (in the destructive interference region) is reduced correspondingly, as shown in Fig. 3(a). However, above 450 nm, where the solar flux is higher, the photon absorption is significantly enhanced due to the surface plasmon resonance. The optimization procedure maximizes the photocurrent by balancing photon absorption below and above 450 nm. The optimal height, radius and lattice constant of the dielectric-filled Ag nanohole array is 30 nm, 90 nm and 260 nm, respectively. This optimized structure exhibits high photon absorption over a broader spectrum compared to hematite on a solid Ag substrate.

 figure: Fig. 3

Fig. 3 Photon absorption in the hematite/dielectric-filled Ag nanohole/Ag substrate structure. (a) Absorption in hematite within 20 nm of the SCLJ for bare 20 nm hematite film, 20 nm hematite/Ag substrate and 20 nm hematite/dielectric-filled Ag nanohole/Ag substrate. The optimal height, radius and lattice constant of Ag nanohole are 30 nm, 90 nm and 260 nm. The electric field magnitude distributions at 550 nm in y = 0 plane for (b) 20 nm hematite/dielectric-filled Ag nanohole/Ag substrate and (c) 20 nm hematite/Ag substrate.

Download Full Size | PDF

Figures 3(b) and 3(c) show the electric field distribution as evidence of the surface plasmon resonance at 550 nm with the nanohole array. When surface plasmon resonances are generated, the collective charge oscillations at the Ag surface create a sharp field enhancement that is spatially confined at the edge of nanoholes and at the Ag/hematite interface, and decays exponentially into the hematite film. These highly intense electromagnetic fields contribute to the photon absorption enhancement above 450 nm.

Finally, we compared the photocurrent and solar-averaged absorption of our proposed structure with recent results from Dotan et al. [11] (state-of-the-art photocurrent for ultrathin planar hematite photoanodes) and Qiu et al. [10] (state-of-the-art photocurrent for ultrathin nanophotonic hematite photoanodes). We report the photocurrent and the solar-averaged absorption of the structure in Dotan et al. [11] based on spectral photon absorption in the hematite film within 20 nm of the SCLJ as in this work. As shown in Fig. 4, the structure proposed in this work exceeds both the calculated photocurrent and the solar-averaged absorption. A direct comparison with Qiu et al. [10] cannot be abtained because they did not report spectral absorption in the hematite film within 20 nm of the SCLJ. Nevertheless, our proposed structure has comparable photon absorption properties to their ultrathin hematite films with nanophotonic structures. Moreover, our proposed structure should exhibit lower recombination losses because of its planar structure [14–16]. Therefore, our structure promises to exceed the state-of-the-art photocurrent of ultrathin hematite photoanodes.

 figure: Fig. 4

Fig. 4 Experimental and theoretical photocurrents (left axis) and solar-averaged absorption (right axis) comparing prior nanophotonic and planar structured hematite films reported in Qiu et al. [10] and Dotan et al. [11] with the ultrathin planar hematite film proposed and modeled in this work.

Download Full Size | PDF

4. Conclusions

In conclusion, we have theoretically designed and characterized an ultrathin planar hematite photoanode for solar PEC water splitting. We have demonstrated that the use of an optimized dielectric-filled silver nanohole array as a substrate improves broad band absorption due to supported destructive interference and surface plasmon resonance. Our proposed structure has higher solar-averaged absorption compared to previous ultrathin hematite film photoanodes with planar and nanophotonic structures. In addition, the planar nature of our design should enable lower recombination losses. This design has the potential to exceed the state-of-the-art photocurrent of ultrathin hematite photoanodes and aid in making solar PEC water splitting a more efficient and viable renewable energy alternative.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (Nos. 51236004 and 51321002) (for design and modeling) and as part of the Solid-State Solar-Thermal Energy Conversion Center (S3TEC), an Energy Frontier Research Center funded by the U.S. Department of Energy, Office of Science, Basic Energy Sciences under Award DE-SC0001299 (for theoretical framework).

References and links

1. K. Sivula, F. Le Formal, and M. Grätzel, “Solar water splitting: progress using hematite (α-Fe2 O3) photoelectrodes,” ChemSusChem 4(4), 432–449 (2011). [CrossRef]   [PubMed]  

2. Z. Chen, T. F. Jaramillo, T. G. Deutsch, A. Kleiman-Shwarsctein, A. J. Forman, N. Gaillard, R. Garland, K. Takanabe, C. Heske, M. Sunkara, E. W. McFarland, K. Domen, E. L. Miller, J. A. Turner, and H. N. Dinh, “Accelerating materials development for photoelectrochemical hydrogen production: standards for methods, definitions, and reporting protocols,” J. Mater. Res. 25(1), 3–16 (2010). [CrossRef]  

3. K. Sivula, R. Zboril, F. Le Formal, R. Robert, A. Weidenkaff, J. Tucek, J. Frydrych, and M. Grätzel, “Photoelectrochemical water splitting with mesoporous hematite prepared by a solution-based colloidal approach,” J. Am. Chem. Soc. 132(21), 7436–7444 (2010). [CrossRef]   [PubMed]  

4. J. Brillet, M. Grätzel, and K. Sivula, “Decoupling feature size and functionality in solution-processed, porous hematite electrodes for solar water splitting,” Nano Lett. 10(10), 4155–4160 (2010). [CrossRef]   [PubMed]  

5. S. D. Tilley, M. Cornuz, K. Sivula, and M. Grätzel, “Light-induced water splitting with hematite: improved nanostructure and iridium oxide catalysis,” Angew. Chem. Int. Ed. Engl. 49(36), 6405–6408 (2010). [CrossRef]   [PubMed]  

6. J. Y. Kim, G. Magesh, D. H. Youn, J. W. Jang, J. Kubota, K. Domen, and J. S. Lee, “Single-crystalline, wormlike hematite photoanodes for efficient solar water splitting,” Sci. Rep. 3, 2681 (2013). [PubMed]  

7. Y. Ling, G. Wang, D. A. Wheeler, J. Z. Zhang, and Y. Li, “Sn-doped hematite nanostructures for photoelectrochemical water splitting,” Nano Lett. 11(5), 2119–2125 (2011). [CrossRef]   [PubMed]  

8. S. H. Shen, J. G. Jiang, P. H. Guo, C. X. Kronawitter, S. S. Mao, and L. J. Guo, “Effect of Cr doping on the photoelectrochemical performance of hematite nanorod photoanodes,” Nano Energy 1(5), 732–741 (2012). [CrossRef]  

9. M. Li, J. J. Deng, A. W. Pu, P. P. Zhang, H. Zhang, J. Gao, Y. Y. Hao, J. Zhong, and X. H. Sun, “Hydrogen-treated hematite nanostructures with low onset potential for highly efficient solar water oxidation,” J. Mater. Chem. A Mater. Energy Sustain. 2(19), 6727–6733 (2014). [CrossRef]  

10. Y. Qiu, S. F. Leung, Q. Zhang, B. Hua, Q. Lin, Z. Wei, K. H. Tsui, Y. Zhang, S. Yang, and Z. Fan, “Efficient photoelectrochemical water splitting with ultrathin films of hematite on three-dimensional nanophotonic structures,” Nano Lett. 14(4), 2123–2129 (2014). [CrossRef]   [PubMed]  

11. H. Dotan, O. Kfir, E. Sharlin, O. Blank, M. Gross, I. Dumchin, G. Ankonina, and A. Rothschild, “Resonant light trapping in ultrathin films for water splitting,” Nat. Mater. 12(2), 158–164 (2012). [CrossRef]   [PubMed]  

12. K. X. Wang, Z. F. Yu, V. Liu, M. L. Brongersma, T. F. Jaramillo, and S. H. Fan, “Nearly total solar absorption in ultrathin nanostructured iron oxide for efficient photoelectrochemical water splitting,” ACS Photonics 1(3), 235–240 (2014). [CrossRef]   [PubMed]  

13. S. J. Kim, I. Thomann, J. Park, J. H. Kang, A. P. Vasudev, and M. L. Brongersma, “Light trapping for solar fuel generation with Mie resonances,” Nano Lett. 14(3), 1446–1452 (2014). [CrossRef]   [PubMed]  

14. J. Tang, Z. Huo, S. Brittman, H. Gao, and P. Yang, “Solution-processed core-shell nanowires for efficient photovoltaic cells,” Nat. Nanotechnol. 6(9), 568–572 (2011). [CrossRef]   [PubMed]  

15. R. Yu, Q. F. Lin, S. F. Leung, and Z. Y. Fan, “Nanomaterials and nanostructures for efficient light absorption and photovoltaics,” Nano Energy 1(1), 57–72 (2012). [CrossRef]  

16. F. Xiu, H. Lin, M. Fang, G. F. Dong, S. P. Yip, and J. C. Ho, “Fabrication and enhanced light-trapping properties of three-dimensional silicon nanostructures for photovoltaic applications,” Pure Appl. Chem. 86(5), 557–573 (2014). [CrossRef]  

17. I. Thomann, B. A. Pinaud, Z. Chen, B. M. Clemens, T. F. Jaramillo, and M. L. Brongersma, “Plasmon enhanced solar-to-fuel energy conversion,” Nano Lett. 11(8), 3440–3446 (2011). [CrossRef]   [PubMed]  

18. B. Iandolo, T. J. Antosiewicz, A. Hellman, and I. Zorić, “On the mechanism for nanoplasmonic enhancement of photon to electron conversion in nanoparticle sensitized hematite films,” Phys. Chem. Chem. Phys. 15(14), 4947–4954 (2013). [CrossRef]   [PubMed]  

19. M. A. Kats, R. Blanchard, P. Genevet, and F. Capasso, “Nanometre optical coatings based on strong interference effects in highly absorbing media,” Nat. Mater. 12(1), 20–24 (2012). [CrossRef]   [PubMed]  

20. A. Taflove and S. C. Hagness, Computational Eletrodynamics: the Finite-Difference Time-Domain Method (Artech House, 2005).

21. J. Robinson and Y. Rahmat-Samii, “Particle swarm optimization in electromagnetics,” IEEE Trans. Antenn. Propag. 52(2), 397–407 (2004). [CrossRef]  

22. N. Beermann, L. Vayssieres, S. Lindquist, and A. Hagfeldt, “Photoelectrochemical studies of oriented nanorod thin films of hematite,” J. Electrochem. Soc. 147(7), 2456–2461 (2000). [CrossRef]  

23. S. Jeong, M. D. McGehee, and Y. Cui, “All-back-contact ultra-thin silicon nanocone solar cells with 13.7% power conversion efficiency,” Nat. Commun. 4, 2950 (2013). [CrossRef]   [PubMed]  

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 Ultrathin planar hematite/metal nanohole array/metal substrate structure for solar PEC water splitting. (a) Layer 1 is water, layer 2 is hematite, layer 3 is dielectric (refractive index n = 1.4) filled metal nanohole array with hexagonal lattice and layer 4 is metal substrate. (b) Optical constants ( m = n + i κ ) of hematite.
Fig. 2
Fig. 2 Photocurrent and photon absorption of hematite films on Au, Al and Ag substrates. (a) Photocurrent and (b) solar-averaged absorption variations with hematite film thicknesses. Spectral total absorption, αtot, and the spectral absorption within 20 nm of the SCLJ, α, for hematite films on (c) Au, (d) Al and (e) Ag substrates. (f) Phasor diagrams of the reflected partial waves at 400 and 550 nm for 20 nm hematite/Ag structure.
Fig. 3
Fig. 3 Photon absorption in the hematite/dielectric-filled Ag nanohole/Ag substrate structure. (a) Absorption in hematite within 20 nm of the SCLJ for bare 20 nm hematite film, 20 nm hematite/Ag substrate and 20 nm hematite/dielectric-filled Ag nanohole/Ag substrate. The optimal height, radius and lattice constant of Ag nanohole are 30 nm, 90 nm and 260 nm. The electric field magnitude distributions at 550 nm in y = 0 plane for (b) 20 nm hematite/dielectric-filled Ag nanohole/Ag substrate and (c) 20 nm hematite/Ag substrate.
Fig. 4
Fig. 4 Experimental and theoretical photocurrents (left axis) and solar-averaged absorption (right axis) comparing prior nanophotonic and planar structured hematite films reported in Qiu et al. [10] and Dotan et al. [11] with the ultrathin planar hematite film proposed and modeled in this work.

Equations (3)

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

J ph = e 300 nm 590 nm α ( λ ) I ( λ ) h c / λ d λ
α solar = 300 nm 590 nm α ( λ ) I ( λ ) d λ 300 nm 590 nm I ( λ ) d λ
r = l = 0 r l
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.