Next Article in Journal
Design of Inner Matching Three-Stage High-Power Doherty Power Amplifier Based on GaN HEMT Model
Previous Article in Journal
Wide Field of View Under-Panel Optical Lens Design for Fingerprint Recognition of Smartphone
Previous Article in Special Issue
Gold Nanoparticle-Modified Carbon-Fiber Microelectrodes for the Electrochemical Detection of Cd2+ via Fast-Scan Cyclic Voltammetry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Deposition of FeOOH Layer on Ultrathin Hematite Nanoflakes to Promote Photoelectrochemical Water Splitting

Shandong Provincial Key Laboratory of Optical Communication Science and Technology, School of Physics Science and Information Technology, Liaocheng University, Liaocheng 252059, China
*
Author to whom correspondence should be addressed.
Micromachines 2024, 15(3), 387; https://doi.org/10.3390/mi15030387
Submission received: 10 January 2024 / Revised: 10 March 2024 / Accepted: 12 March 2024 / Published: 13 March 2024
(This article belongs to the Special Issue Microelectrodes and Microdevices for Electrochemical Applications)

Abstract

:
Hematite is one of the most promising photoanode materials for the study of photoelectrochemical (PEC) water splitting because of its ideal bandgap with sufficient visible light absorption and stability in alkaline electrolytes. However, owing to the intrinsically high electron-hole recombination, the PEC performance of hematite is still far below that expected. The efficient charge separation can be achieved via growth of FeOOH on hematite photoanode. In this study, hematite nanostructures were successfully grown on the surface of iron foil by the simple immersion deposition method and thermal oxidation treatment. Furthermore, cocatalyst FeOOH was successfully added to the hematite nanostructure surface to improve charge separation and charge transfer, and thus promote the photoelectrochemical water splitting. By utilizing the FeOOH overlayer as a cocatalyst, the photocurrent density of hematite exhibited a substantial 86% increase under 1.5 VRHE, while the onset potential showed an apparent shift towards the cathodic direction. This can be ascribed to the high reaction area for the nanostructured morphology and high electrocatalytic activity of FeOOH that enhanced the amount of photogenerated holes and accelerated the kinetics of water splitting.

1. Introduction

Photoelectrochemical (PEC) water splitting is a promising method of hydrogen energy production that utilizes the renewable resource of solar energy to split water into hydrogen and oxygen to achieve zero carbon emissions [1,2,3,4,5,6]. Among many photoanodes of metal oxides, hematite (α-Fe2O3) stands out among various photoanodes due to its low cost, high chemical stability in alkaline electrolytes, high earth abundance, and appropriate bandgap (~2.2 eV) [7,8,9,10]. For hematite, in theory, its solar-to-hydrogen energy conversion efficiencies can reach 16.8% [11,12,13]; however, it is far below this theoretical value in practice. This is mainly due to the low absorption coefficient, short excited-state lifetime (3–10 ps), slow kinetics for the oxygen evolution process, short carrier diffusion length (2–4 nm), and the poor conductivity of hematite (about 10−2 cm2 V−1s−1 at 20 °C) [14,15,16,17].
Therefore, numerous researchers have come up with various approaches to solve those problematic flaws above, such as control morphology [18], ion doping [11,14], construction of heterojunction/homojunction [6,15], and deposition of cocatalysts [19,20]. Specifically, the use of nanostructures can shorten the transport distance of the holes, so people have prepared hematite nanostructures with different micromorphology, such as nanowire, nanorods, nanotubes, nanocubes, and nanosheets [20,21,22,23,24]. In order to effectively reduce the surface recombination caused by the surface state, using a cocatalyst is an important means of reducing overpotential, accelerating reaction kinetics, inhibiting corrosion, and accelerating surface charge separation [25,26,27]. Various cocatalyst materials have been studied and deposited to the surface of hematite, including IrO2 [28], cobalt phosphate (Co-Pi) [29], NiFeOx [30], and nickel(II) hydroxide [31]. FeOOH came to the attention of researchers because of its low cost of synthesis and good compatibility with hematite [25,32]. For example, Ye et al. reported that hematite photoanodes decorated with nanostructured FeOOH displayed four times enhancement of photocurrent density in the PEC water oxidation [33]. Bi and co-workers reported the decoration of FeOOH and Au cocatalysts on hematite nanoflake photoanodes showed a photocurrent density of 6.5 mA cm−2 at 1.6 VRHE under AM 1.5 G simulated sunlight [25]. Recently, researchers combined bulk and surface engineering to change the photocurrent density of hematite photoanodes. Mai et al. introduced a Ti dopant into Fe2O3 crystal structure to enhance charge separation efficiency and bulk kinetics, and deposited FeOOH on the Ti-doped Fe2O3 to promote the charge transfer efficiency and surface kinetics [34]. FeOOH can facilitate the kinetics of water oxidation as an oxygen evolution cocatalyst, which has widely been utilized in photoanodes [35,36,37,38,39].
In this study, in order to shorten the hole transport distance, we successfully grew the nanoflake structure on the surface of the iron foil. In order to solve the problem that a large number of photogenerated holes are accumulated on the surface of photoanode materials due to slow oxygen generation kinetics, we successfully loaded the cocatalyst FeOOH on the surface of the photoanode to improve the transfer efficiency of the photogenerated holes. The loaded FeOOH layer provides a way of avoiding the accumulation of holes on the surface of photoanode materials, so that the holes can effectively participate in the water oxidation reaction and improve the oxygen evolution reaction. Therefore, the micromorphology and surface modification of the sample have certain influence on the PEC performance of hematite, which provides an inspiration for designing efficient photoelectrode devices.

2. Materials and Methods

2.1. Preparation of α-Fe2O3 and α-Fe2O3/FeOOH Nanosheets

We prepared hematite on iron foil (99.99%, Alfa Aesar, Haverhill, MA, USA), which was degreased by sonicating in acetone and ethanol for 10 min, and then dried in an air stream. Then, the iron foil was directly placed in the muffle furnace (KSL-1100X, HFKejing furnace, Hefei, China) in air at 450 °C, and maintained at the desired temperature for 1.5 h with a heating rate of 10 °C min−1. As for the preparation of α-Fe2O3/FeOOH samples, primarily, the iron foils were immersed in 5 mM FeCl3·6H2O (99%, Aladdin, Shanghai, China) aqueous solution for 10 min, 20 min, 30 min, 50 min, and 60 min, following by rinsing with ultrapure water and drying in an air stream. In the control group, the iron foils were soaked in 10 mM FeCl3·6H2O (99%, Aladdin) aqueous solution for 10 min, 20 min, 30 min, 50 min, and 60 min, following by rinsing with ultrapure water and drying in an air stream. After that, the samples were thermally annealed in a muffle furnace (KSL-1100X, HFKejing furnace) in air at 450 °C, maintained at the desired temperature for 1.5 h, and the heating rate was 10 °C min−1. The samples were taken out of the furnace when the temperature in the furnace decreased to room temperature. In the following, α-Fe2O3/FeOOH refers to the samples obtained after the iron foil was immersed in 5 mM FeCl3·6H2O (99%, Aladdin) aqueous solution for 30 min for annealing.

2.2. Material Characterization

The pristine α-Fe2O3 and α-Fe2O3/FeOOH samples were examined for surface morphology using a thermal field emission scanning electron microscope (MERLIN compact, Oberkochen, Germany). Using a 532 nm laser as the excitation source, a Raman spectrometer (iHR 550 HORIBA, Kyoto, Japan) was used to record the Raman spectra of hematite that was soaked and unsoaked for varying durations. X-ray photoelectron spectroscopy (XPS, Escalab Xi+, Waltham, MA, USA) was utilized to illustrate the FeOOH layer on hematite and determine the surface composition of the pristine α-Fe2O3 and α-Fe2O3/FeOOH samples.

2.3. Photoelectrochemical Performance Measurements

The PEC performance experiments used a three-electrode configuration. A saturated calomel electrode (SCE) was employed as the reference electrode, platinum mesh was used for the counter electrode, and the prepared α-Fe2O3 and α-Fe2O3/FeOOH photoanodes served as the working electrodes in all PEC examinations. An electrolyte solution of 1M NaOH (pH 13.6) was utilized. Stirring was used throughout the entire PEC measuring process. Without the use of a light filter, a 500 W xenon lamp with a light intensity of 225 mW cm−2 was used as the light source.
A CHI660E potentiostat (CH Instruments, Austin, TX, USA) was implemented for determining the photocurrent density using linear sweep voltammetry (LSV) at a scan rate of 10 mV s−1. The equation ERHE = ESCE + 0.059 pH + E0SCE, where ESCE is the experimentally observed potential against the reference electrode and E0SCE = 0.245 V at ambient temperature, was used to convert the measured potentials versus SCE to the reversible hydrogen electrode (RHE) scale. In 1 M NaOH electrolyte, Mott–Schottky experiments were carried out using the same photoelectrochemical apparatus in a dark environment. The scan rate was 50 mV s−1 and the measurement frequency was 1000 Hz. Electrochemical impedance spectroscopy (EIS) measurements were carried out at 10,000 to 0.1 Hz using 0.5 V versus SCE in the dark.

3. Results and Discussion

The morphology of α-Fe2O3 nanoflakes and hematite modified with the FeOOH cocatalyst is shown in Figure 1. It can be clearly seen that α-Fe2O3 nanoflakes appear, but it can also be seen that the distribution of the nanoflakes is not uniform and is relatively sparse, as shown in Figure 1a. In combined immersion deposition and thermal oxidation treatment in air, the morphology of samples makes a difference. Figure 1b shows the SEM image of the α-Fe2O3/FeOOH combined with immersion for 10 min and sintering in the muffle furnace at 450 °C for 1.5 h. One can find that there are less nanoflakes than in the sample of α-Fe2O3. However, as shown in Figure 1c, the nanoflakes gradually increase under the immersion time of 20 min. Once the immersion time reaches 30 min, the sample shows uniformly distributed nanoflakes, which can be seen in Figure 1d. The length of hematite nanoflakes is about 1–2 μm and the thickness of the flakes tapers down from the base to the tip. Similarly, in Figure 1e, the sample still exhibits dense nanoflakes under the immersion time of 50 min. Nevertheless, from Figure 1f, the nanoflakes decrease drastically when the immersion time reaches 60 min. In consequence, a suitable immersion time plays an important role in forming uniformly distributed nanoflakes before sintering. It can be seen that after the optimized decoration cocatalyst FeOOH modification, the nanoflakes are much denser than the pristine hematite sample. From the surface morphology of hematite, one can find that the uniformly distributed nanoflakes were obtained after proper immersion treatment. Indeed, in a previous study, pre-polishing was necessary to obtain dense nanoflakes for an iron foil substrate [40]. Therefore, the suitable decoration cocatalyst FeOOH on hematite can not only make the nanoflakes dense and uniform on the iron surface, but also change the surface area of the sample for surface reaction and shorten the distance of hole transport.
Figure 2 displays the Raman spectra of hematite samples sintered under different conditions. Sample (a) was the original control group (annealing at 450 °C for 1.5 h), and samples (b)–(f) were soaked for 10 min, 20 min, 30 min, 50 min, and 60 min, respectively. Each sample displays the identical hematite characteristic peak in the range of 200–800 cm−1. Hematite belongs to the crystallographic space groups, which include A1g modes (223 cm−1 and 497 cm−1) and E1g modes (242, 290, 410, and 612 cm−1) [41,42].
X-ray photoelectron spectroscopy was used to discern whether the pristine hematite and the modified hematite were covered by the FeOOH film. All samples were calibrated with a standard carbon peak of 284.6 eV. Figure 3 presents a summary of the XPS findings for the hematite and α-Fe2O3/FeOOH samples. The Fe 2p spectra of the as-grown sample (Figure 3b) exhibits a typical hematite spectrum with binding energies of 710.6 eV for Fe 2p3/2 and 723.59 eV for Fe 2p1/2. Additionally, there is a shakeup satellite at 718 eV, which corresponds to the characteristic satellite for phases of α-Fe2O3 or γ-Fe2O3. The Raman spectra analysis excluded the potential existence of γ-Fe2O3, as no vibration mode associated with γ- Fe2O3 was observed. In Figure 3e, the FeOOH-decorated α- Fe2O3 sample exhibits signal peaks at 711.02 eV and 724.8 eV, which correspond to Fe 2p3/2 and Fe 2p1/2, respectively. These results are consistent with the valence state of Fe3+ in typical α-Fe2O3 [43,44]. Figure 3c,f show the O 1 s spectra of as-grown hematite and the FeOOH decoration on α-Fe2O3 samples. After the addition of the cocatalyst, the peak of O 1 s can be divided into two peaks. The peaks at 529.8 eV and 531.68 eV correspond to the Fe-O peak in Fe2O3 and the Fe-OH peak in FeOOH, respectively. The signal peak located at 529.8 eV corresponds to the lattice oxygen in Fe2O3 [32]. The above results are also consistent with those reported in other studies in the literature [45,46].
The α-Fe2O3/FeOOH nanostructures were prepared by combining the immersion deposition method and thermal oxidation treatment. The surface of the iron sheet was formed as an ultrathin and dense FeOOH film for decoration to enhance the hematite photoanode PEC water oxidation performance. Utilizing a xenon lamp (225 mW cm−2) to measure the photocurrent density with bias potential in 1 M NaOH electrolyte, the PEC water oxidation activity of several photoanodes was investigated, as shown in Figure 4.
According to Figure 4a, the as-grown α-Fe2O3 nanoarrays exhibited a water oxidation initial potential of 0.9 V when the potential was swept from 0.8 V to 1.6 V under illumination, and the current density was 0.15 mA/cm2 at 1.23 V. For the α-Fe2O3/FeOOH(3) photoanode, it is evident that there is an obvious enhanced photoresponse from 1.1 V to 1.6 V, and the photocurrent density is up to 0.2 mA/cm2 at 1.23 V, while there is a 1.33 times enhancement of the photocurrent density compared to the as-grown hematite. Moreover, it can be seen that the photocurrent density increases significantly and can reach 0.78 mA/cm2 at 1.5V, which is 1.9 times that of the as-grown hematite (0.42 mA/cm2). By adjusting the immersion period, the quantities of FeOOH on the hematite nanoflakes could be customized. Figure 4a also shows the solar water splitting performance of α-Fe2O3/FeOOH photoanodes. It is evident that the photocurrent density increases with lowering the amount of FeOOH (20 min), particularly for bias potentials above 1.3 V. However, the data indicate a decline in photocurrent density between 0.9 and 1.25 V when compared to the as-grown hematite. Compared to the as-prepared hematite, a further immersion time increase of up to 50 min improves the photoresponse; however, it is still less than the optimal time of 30 min. Furthermore, it is evident that the photocurrent density decreases with soaking times of 10 and 60 min compared to pristine hematite.
In order to explore the influence of FeCl3, 10 mM FeCl3 was selected for comparison. After an equal immersion time in 10 mM FeCl3 and then sintering in the same conditions, the PEC water splitting performance is exhibited in Figure 4b. One can find that α-Fe2O3/FeOOH(c) shows the optimal photocurrent density during the measured applied bias potential. Its photocurrent density is at 1.23 V up to 0.56 mA cm−2, and a maximum photocurrent density of 0.92 mA cm−2 is obtained at 1.6 V. In order to provide a more visual comparison of the influence of immersion concentration, Figure 5 summarizes the photocurrent density of α-Fe2O3, α-Fe2O3/FeOOH(3), and α-Fe2O3/FeOOH(c) under the potential of 1.23 V, 1.4 V, and 1.6 V, respectively. Obviously, the α-Fe2O3/FeOOH(3) exhibits the optimal PEC water splitting performance. Therefore, a suitable FeOOH amount on the hematite photoanodes results in an improvement in the PEC water oxidation.
To conduct further electrochemical characterization, Mott–Schottky (M-S) analysis was carried out, as depicted in Figure 6a. The flat band potential and donor density can be ascertained using the M-S measurement. Both the α-Fe2O3 and α-Fe2O3/FeOOH photoanodes exhibit positive slopes, indicating their n-type semiconductor characteristic with electrons as the majority carriers. As for the flat band potential, which was derived from the intercept of the x-axis, it can be seen that 0.14 V shifts to the cathodic direction. The FeOOH layer may be responsible for the shift in the flat band potential due to the passivation of surface states. Surface states induce a decrease in potential within the Helmholtz layer; however, this effect can be reduced through surface passivation. The FeOOH overlayer can efficiently passivate the surface of the nanoflakes by creating a wider contact area.
1 C 2 = 2 e ε 0 ε N d V V fb k T e
The above Mott–Schottky equation can be used to estimate the carrier density [11,40]. It involves the specific capacitance (C), electron charge (e), permittivity of vacuum (ε0), dielectric constant of hematite (ε = 80), donor density (Nd), applied voltage (V), flat band potential (Vfb), Boltzmann constant (kB), and absolute temperature (T). The donor density is derived from the gradient calculated in the Mott−Schottky plots:
N d = 2 e ε 0 ε d 1 C 2 d V 1
The carrier concentrations of α-Fe2O3/FeOOH and α-Fe2O3 were calculated to be 1.48 × 1019 cm−3 and 4.88 × 1018 cm−3, respectively. After adding the cocatalyst, the carrier concentration of the photoanode material increased. The α-Fe2O3/FeOOH structure exhibits a lower flat band potential and a larger carrier density, indicating that the formation of ultrathin FeOOH films could enhance the process of charge separation and transfer.
Electrochemical impedance spectroscopy was used to analyze the process taking place at the interface between the photoelectrodes and the electrolyte. The measurements were conducted under dark conditions. Figure 6b displays Nyquist plots of α-Fe2O3 and α-Fe2O3/FeOOH. The inset in the upper right corner represents the equivalent circuit, where Rs denotes the series resistance, Rct and CPE represent the charge transfer resistance and the capacitance of the depletion layer, respectively, and Ws represents the Warburg impedance. The fitted curves, shown by solid lines, exhibit a high level of agreement with the measured points, which are denoted by symbols. The simulated results show that RCT and CPE of α-Fe2O3 are 2713 Ω and 1.15 × 10−4 F, respectively, and those of α-Fe2O3/FeOOH are 1027 Ω, and 3.56 × 10−4 F. While comparing the hematite nanoflake photoanode with the sample covered in FeOOH film, it was observed that the charge transfer resistance decreased and the capacitance increased. This suggests that the FeOOH layer on the hematite reduced the charge transfer resistance and improved the capacitance at the interface between the photoanode and electrolyte.
In this work, we tried to use the combination of immersion pre-treating the iron foil and annealing in an air environment, which exhibits a more efficient PEC water splitting performance for α-Fe2O3/FeOOH. On the basis of the above results, a possible mechanism for illustrating the photogenerated charge transfer on α-Fe2O3/FeOOH is displayed in Figure 7. Under light illumination, hematite can be excited to induce photogenerated electrons to leap into the conduction band; meanwhile, photogenerated holes form on the valance band. The photoinduced electrons favorably transfer from the conduction band of hematite to the iron substrate via the nanoflake structure, then to the Pt electrode through an external circuit for hydrogen evolution reaction. The photoinduced holes transfer from the valence band of hematite to the interface of the photoanode/electrolyte to participate in the water oxidation reaction. However, during the photogenerated charge transport, the separated photogenerated charges rapidly recombine with each other during the short hole-diffused length of hematite. As for the α-Fe2O3/FeOOH photoanode, there is the ultrathin FeOOH layer to decorate the hematite surface. The FeOOH layer can serve as an oxygen evolution cocatalyst. It could make the photogenerated holes on the valence band of hematite transfer to the FeOOH layer [47,48,49], to finish the water oxidation and produce oxygen, which is shown in the inset of Figure 7. On the other hand, the FeOOH facilitates the suppression of the formation of the surface recombination centers. As a result, the PEC water splitting performance of α-Fe2O3/FeOOH is enhanced.

4. Conclusions

In summary, a hematite nanostructure was successfully prepared on iron foil by immersion precipitation and the thermal oxidation treatment method. Different immersion times and concentrations were explored to ensure a suitable FeOOH amount. The results demonstrate that the suitable FeOOH amount decorated on hematite photoanodes could make the nanoflakes dense and uniform on the iron surface and result in an improvement in PEC water oxidation. When the bias voltage is 1.5 V, the photocurrent density of the α-Fe2O3/FeOOH photoanode is increased by about 86% compared with that of the pristine photoanode. The FeOOH layer on the photoanode surface provides an alternative way to avoid the accumulation of holes on the α-Fe2O3 photoanode surface, so that the holes can effectively participate in the water oxidation process. The surface modification of FeOOH can effectively promote the oxygen evolution reaction of semiconductor α-Fe2O3 and improve the charge transfer efficiency. As a result, we identified a simple way to obtain large-area hematite nanoflakes on iron foil and then further to use FeOOH to enhance the photoelectrochemical performance of hematite.

Author Contributions

Conceptualization, W.Z.; methodology, W.Z. and Y.Z.; investigation, X.M. and L.Z.; writing—original draft preparation, W.Z. and Y.Z.; writing—review and editing, X.M., L.Z. and C.Z.; resources, X.M. and L.Z.; Funding acquisition, C.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (No. 61804072) and the Natural Science Foundation of Shandong Province (No. ZR2018PF013, ZR2019PA002, ZR2017BF009).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Liu, Y.; Yu, Y.X.; Zhang, W.D. Photoelectrochemical properties of Ni-doped Fe2O3 thin films prepared by electrodeposition. Electrochim. Acta 2012, 59, 121–127. [Google Scholar] [CrossRef]
  2. Gong, M.; Li, Y.; Wang, H.; Liang, Y.; Wu, J.Z.; Zhou, J.; Wang, J.; Regier, T.; Wei, F.; Dai, H. An advanced Ni-Fe layered double hydroxide electrocatalyst for water oxidation. J. Am. Chem. Soc. 2013, 135, 8452–8455. [Google Scholar] [CrossRef]
  3. Sivula, K.; Formal, F.L.; Grätzel, M. Solar water splitting: Progress using hematite (alpha-Fe2O3) photoelectrodes. ChemSusChem 2011, 4, 432–449. [Google Scholar] [CrossRef]
  4. Hou, Y.; Zuo, F.; Dagg, A.; Feng, P. Visible light-driven alpha-Fe2O3 nanorod/graphene/BiV1−xMoxO4 core/shell heterojunction array for efficient photoelectrochemical water splitting. Nano Lett. 2012, 12, 6464–6473. [Google Scholar] [CrossRef] [PubMed]
  5. Mishra, M.; Chun, D. α-Fe2O3 as a photocatalytic material: A review. Appl. Catal. A Gen. 2015, 498, 126–141. [Google Scholar] [CrossRef]
  6. Hong, Y.; Jeon, S.; Jeong, H.; Ryu, H. Systematic study of α-Fe2O3 stacked homojunction photoelectrochemical photoelectrode. Ceram. Int. 2021, 47, 7814–7823. [Google Scholar] [CrossRef]
  7. Xu, X.; Pan, L.; Zhang, X.; Wang, L.; Zou, J. Rational design and construction of cocatalysts for semiconductor-based photo-electrochemical oxygen evolution: A comprehensive review. Adv. Sci. 2019, 6, 1801505. [Google Scholar] [CrossRef] [PubMed]
  8. Tilley, S.D. Recent Advances and emerging trends in photo-electrochemical solar energy conversion. Adv. Energy Mater. 2019, 9, 1802877. [Google Scholar] [CrossRef]
  9. Tamirat, A.G.; Rick, J.; Dubale, A.A.; Sub, W.N.; Hwang, B.J. Using hematite for photoelectrochemical water splitting: A review of current progress and challenges. Nanoscale Horiz. 2016, 1, 243–267. [Google Scholar] [CrossRef] [PubMed]
  10. Xi, L.; Bassi, P.S.; Chiam, S.Y.; Mak, W.F.; Tran, P.D.; Barber, J.; Loo, J.S.C.; Wong, L.H. Surface treatment of hematite photoanodes with zinc acetate for water oxidation. Nanoscale 2012, 4, 4430–4433. [Google Scholar] [CrossRef]
  11. Ling, Y.; Wang, G.; Wheeler, D.A.; Zhang, J.Z.; Li, Y. Sn-doped hematite nanostructures for photoelectrochemical water splitting. Nano Lett. 2011, 11, 2119–2125. [Google Scholar] [CrossRef] [PubMed]
  12. Gurudayal; Sabba, D.; Mulmudi, H.K.; Wong, L.; Barber, J.; Grätzel, M.; Mathews, N. Perovskite–hematite tandem cells for efficient overall solar driven water splitting. Nano Lett. 2015, 15, 3833–3839. [Google Scholar] [CrossRef] [PubMed]
  13. Lin, Y.; Yuan, G.; Sheehan, S.; Zhou, S.; Wang, D. Hematite-based solar water splitting: Challenges and opportunities. Energy Environ. Sci. 2011, 4, 4862–4869. [Google Scholar] [CrossRef]
  14. Luo, Z.; Li, C.; Liu, S.; Wang, T.; Gong, J. Gradient doping of phosphorus in Fe2O3 nanoarray photoanodes for enhanced charge separation. Chem. Sci. 2017, 8, 91–100. [Google Scholar] [CrossRef]
  15. Bai, S.; Chu, H.; Xiang, H.; Luo, R.; He, F.; Chen, A. Fabricating of Fe2O3/BiVO4 heterojunction based photoanode modified with NiFe-LDH nanosheets for efficient solar water splitting. Chem. Eng. J. 2018, 350, 148–156. [Google Scholar] [CrossRef]
  16. Kment, S.; Riboni, F.; Pausova, S.; Wang, L.; Han, H.; Hubicka, Z.; Krysa, J.; Schmuki, P.; Zboril, R. Photoanodes based on TiO2 and alpha-Fe2O3 for solar water splitting—Superior role of 1D nanoarchitectures and of combined heterostructures. Chem. Soc. Rev. 2017, 46, 3716–3769. [Google Scholar] [CrossRef]
  17. Forster, M.; Potter, R.J.; Ling, Y.; Yang, Y.; Klug, D.R.; Li, Y.; Cowan, A.J. Oxygen deficient alpha-Fe2O3 photoelectrodes: A balance between enhanced electrical properties and trap-mediated losses. Chem. Sci. 2015, 6, 4009–4016. [Google Scholar] [CrossRef]
  18. Kim, J.Y.; Magesh, G.; Youn, D.H.; Jang, J.W.; Kubota, J.; Domen, K.; Lee, J.S. Single-crystalline, wormlike hematite photoanodes for efficient solar water splitting. Sci. Rep. 2013, 3, 2681. [Google Scholar] [CrossRef]
  19. Qi, Y.; Zhang, J.; Kong, Y.; Zhao, Y.; Chen, S.; Li, D.; Liu, W.; Chen, Y.; Xie, T.; Cui, J.; et al. Unraveling of cocatalysts photodeposited selectively on facets of BiVO4 to boost solar water splitting. Nat. Commun. 2022, 13, 484. [Google Scholar] [CrossRef]
  20. Ye, K.; Wang, Z.; Li, H.; Yuan, Y.; Huang, Y.; Mai, W. A novel CoOOH/(Ti, C)-Fe2O3 nanorod photoanode for photoelectrochemical water splitting. Sci. China Mater. 2018, 61, 887–894. [Google Scholar] [CrossRef]
  21. Tang, P.; Xie, H.; Ros, C.; Han, L.; Biset-Peiró, M.; He, Y.; Kramer, W.; Rodríguez, A.P.; Saucedo, E.; Galán-Mascarós, J.R.; et al. Enhanced photoelectrochemical water splitting of hematite multilayer nanowire photoanodes by tuning the surface state via bottom-up interfacial engineering. Energy Environ. Sci. 2017, 10, 2124–2136. [Google Scholar] [CrossRef]
  22. Schrebler, R.; Ballesteros, L.A.; Gomez, H.; Grez, P.; Cordova, R.; Munoz, E.; Schrebler, R.; Ramos-Barrado, J.R.; Dalchiele, E.A. Electrochemically grown self-organized hematite nanotube arrays for photoelectrochemical water splitting. J. Electrochem. Soc. 2014, 161, H903–H908. [Google Scholar] [CrossRef]
  23. Zheng, C.; Zhu, Z.; Wang, S.; Hou, Y. NaF-assisted hydrothermal synthesis of Ti-doped hematite nanocubes with enhanced photoelectrochemical activity for water splitting. Appl. Surf. Sci. 2015, 359, 805–811. [Google Scholar] [CrossRef]
  24. Xie, X.; Li, K.; Zhang, W.D. Photoelectrochemical properties of Ti-doped hematite nanosheet arrays decorated with CdS nanoparticles. RSC Adv. 2016, 6, 74234–74240. [Google Scholar] [CrossRef]
  25. Wang, L.; Hu, H.; Nguyen, N.T.; Zhang, Y.; Schmuki, P.; Bi, Y. Plasmon-induced hole-depletion layer on hematite nanoflake photoanodes for highly efficient solar water splitting. Nano Energy 2017, 35, 171–178. [Google Scholar] [CrossRef]
  26. Chen, D.; Liu, Z.; Zhang, S. Enhanced PEC performance of hematite photoanode coupled with bimetallic oxyhydroxide NiFeOOH through a simple electroless method. Appl. Catal. B Environ. 2020, 265, 118580. [Google Scholar] [CrossRef]
  27. Liu, W.; Liu, H.; Dang, L.; Zhang, H.; Wu, X.; Yang, B.; Li, Z.; Zhang, X.; Lei, L.; Jin, S. Amorphous cobalt-iron hydroxide nanosheet electrocatalyst for efficient electrochemical and photo-electrochemical oxygen evolution. Adv. Funct. Mater. 2017, 27, 1603904. [Google Scholar] [CrossRef]
  28. Tilley, S.D.; Cornuz, M.; Sivula, K.; Grätzel, M. Light-Induced Water Splitting with Hematite: Improved Nanostructure and Iridium Oxide Catalysis. Angew. Chem. Int. Ed. 2010, 49, 6405–6408. [Google Scholar] [CrossRef] [PubMed]
  29. Klahr, B.; Gimenez, S.; Fabregat-Santiago, F.; Bisquert, J.; Hamann, T.W. Photoelectrochemical and Impedance Spectroscopic Investigation of Water Oxidation with “Co−Pi” Coated Hematite Electrodes. J. Am. Chem. Soc. 2012, 134, 16693–16700. [Google Scholar] [CrossRef] [PubMed]
  30. Du, C.; Yang, X.; Mayer, M.T.; Hoyt, H.; Xie, J.; McMahon, G.; Bischoping, G.; Wang, D. Hematite-Based Water Splitting with Low Turn-On Voltages. Angew. Chem. Int. Ed. 2013, 52, 12692–12695. [Google Scholar] [CrossRef] [PubMed]
  31. Yong, K.M.H.; Hamann, T.W. Enhanced photocatalytic water oxidation efficiency with Ni(OH)2 catalysts deposited on α-Fe2O3 via ALD. Chem. Commun. 2014, 50, 8727–8730. [Google Scholar] [CrossRef] [PubMed]
  32. Kim, J.Y.; Youn, D.H.; Kang, K.; Lee, J.S. Highly conformal deposition of an ultrathin FeOOH layer on a hematite nanostructure for efficient solar water splitting. Angew. Chem. Int. Ed. 2016, 55, 10854–10858. [Google Scholar] [CrossRef]
  33. Yu, Q.; Meng, X.; Wang, T.; Li, P.; Ye, J. Hematite films decorated with nanostructured ferric oxyhydroxide as photoanodes for efficient and stable photoelectrochemical water splitting. Adv. Funct. Mater. 2015, 25, 2686–2692. [Google Scholar] [CrossRef]
  34. Yuan, Y.; Gu, J.; Ye, K.H.; Chai, Z.; Yu, X.; Chen, X.; Zhao, C.; Zhang, Y.; Mai, W. Combining bulk/surface engineering of hematite to synergistically improve its photoelectrochemical water splitting performance. ACS Appl. Mater. Interfaces 2016, 8, 16071–16077. [Google Scholar] [CrossRef] [PubMed]
  35. Chen, P.L.; Zhong, S.M.; Cheng, X.X.; Wang, Z.Q.; Wang, X.T.; Fang, B.Z. Steel slag source-derived FeOOH for enhanced BiVO4 photoelectrochemical water splitting. J. Colloid Interf. Sci. 2024, 665, 417–426. [Google Scholar] [CrossRef]
  36. Li, Y.; Mei, Q.; Liu, Z.J.; Hu, X.S.; Zhou, Z.H.; Huang, J.W.; Bai, B.; Liu, H.; Ding, F.; Wang, Q.Z. Fluorine-doped iron oxyhydroxide cocatalyst: Promotion on the WO3 photoanode conducted photoelectrochemical water splitting. Appl. Catal. B-Environ. 2022, 304, 120995. [Google Scholar] [CrossRef]
  37. Fan, M.M.; Tao, Z.Y.; Zhao, Q.; Li, J.P.; Liu, G.; Zhao, C. Molecular copper phthalocyanine and FeOOH modified BiVO4 photoanodes for enhanced photoelectrochemical water oxidation. Adv. Mater. Technol. 2023, 8, 2201835. [Google Scholar] [CrossRef]
  38. Chen, M.H.; Chang, X.B.; Li, C.; Wang, H.Q.; Jia, L.C. Ni-Doped BiVO4 photoanode for efficient photoelectrochemical water splitting. J. Colloid Interf. Sci. 2023, 640, 162–169. [Google Scholar] [CrossRef]
  39. Li, R.; Wang, Y.M.; Chen, B.J.; Zhang, H.M.; Yan, C.; Xu, X.F.; Humayun, M.; Debecker, D.P.; Wang, C.D. Core-shell structured Co(OH)F@FeOOH enables highly efficient overall water splitting in alkaline electrolyte. Int. J. Hydrog. Energy 2024, 51, 1292–1302. [Google Scholar] [CrossRef]
  40. Zhu, C.; Li, C.; Zheng, M.; Delaunay, J.J. Plasma-induced oxygen vacancies in ultrathin hematite nanoflakes promoting photoelectrochemical water oxidation. ACS Appl. Mater. Interfaces 2015, 7, 22355–22363. [Google Scholar] [CrossRef]
  41. De Faria, D.L.A.; Silva, S.V.; De Oliveira, M.T. Raman microspectroscopy of some iron oxides and oxyhydroxides. J. Raman Spectrosc. 1997, 28, 873–878. [Google Scholar] [CrossRef]
  42. Zhu, C.; Li, C.; Miao, X.; Zhao, L.; Wang, Z.Y.; Delaunay, J.J. Photoelectrochemical water oxidation performance promoted by a cupric oxide-hematite heterojunction photoanode. Int. J. Hydrog. Energy 2020, 45, 33102–33110. [Google Scholar] [CrossRef]
  43. Miao, C.; Ji, S.; Xu, G.; Liu, G.; Zhang, L.D.; Ye, C.H. Micro-nano-structured Fe2O3: Ti/ZnFe2O4 heterojunction films for water oxidation. ACS Appl. Mater. Interfaces 2012, 4, 4428–4433. [Google Scholar] [CrossRef] [PubMed]
  44. Yamashita, T.; Hayes, P. Analysis of XPS spectra of Fe2+ and Fe3+ ions in oxide materials. Appl. Surf. Sci. 2008, 254, 2441–2449. [Google Scholar] [CrossRef]
  45. Li, Z.; Wu, J.; Liao, L.; He, X.; Huang, B.; Zhang, S.; Wei, Y.; Wang, S.; Zhou, W. Surface engineering of hematite nanorods photoanode towards optimized photoelectrochemical water splitting. J. Colloid Interf. Sci. 2022, 626, 879–888. [Google Scholar] [CrossRef] [PubMed]
  46. Yang, P.; Ding, Y.; Lin, Z.; Chen, Z.W.; Li, Y.Z.; Qiang, P.F.; Ebrahimi, M.; Mai, W.J.; Wong, C.P.; Wang, Z.L. Low-cost high-performance solid-state asymmetric supercapacitors based on MnO2 nanowires and Fe2O3 nanotubes. Nano Lett. 2014, 14, 731–736. [Google Scholar] [CrossRef]
  47. Grätzel, M. Photoelectrochemical cells. Nature 2001, 414, 338–344. [Google Scholar] [CrossRef] [PubMed]
  48. Quang, N.D.; Van, P.C.; Majumder, S.; Jeong, J.R.; Kim, D.; Kim, C. Rational construction of S-doped FeOOH onto Fe2O3 nanorods for enhanced water oxidation. J. Colloid Interf. Sci. 2022, 616, 749–758. [Google Scholar] [CrossRef]
  49. Wang, Q.; Zong, X.; Tian, L.; Han, Y.; Ding, Y.; Xu, C.; Tao, R.; Fan, X. Fe2O3/FePO4/FeOOH ternary stepped energy band heterojunction photoanode with cascade-driven charge transfer and enhanced photoelectrochemical performance. ChemSusChem 2022, 15, e202102377. [Google Scholar] [CrossRef]
Figure 1. The morphology of iron foil samples sintered in air for 1.5 h at the same temperature with soaking and without soaking, respectively. (a) Iron foil annealing in air at 450 °C for 1.5 h. (bf) iron foil soaked in 5 mM FeCl3 solution for 10 min, 20 min, 30 min, 50 min, and 60 min, respectively; then, annealed in air at 450 °C for 1.5 h.
Figure 1. The morphology of iron foil samples sintered in air for 1.5 h at the same temperature with soaking and without soaking, respectively. (a) Iron foil annealing in air at 450 °C for 1.5 h. (bf) iron foil soaked in 5 mM FeCl3 solution for 10 min, 20 min, 30 min, 50 min, and 60 min, respectively; then, annealed in air at 450 °C for 1.5 h.
Micromachines 15 00387 g001
Figure 2. Raman spectra of hematite sintered under various soaking times and at the same thermal oxidation temperature in air. Sample (a) was the original control group (annealing at 450 °C for 1.5 h), and samples (bf) were soaked for 10 min, 20 min, 30 min, 50 min, and 60 min, respectively.
Figure 2. Raman spectra of hematite sintered under various soaking times and at the same thermal oxidation temperature in air. Sample (a) was the original control group (annealing at 450 °C for 1.5 h), and samples (bf) were soaked for 10 min, 20 min, 30 min, 50 min, and 60 min, respectively.
Micromachines 15 00387 g002
Figure 3. X-ray photoelectron spectroscopy analysis: (a) survey spectra of the as-grown hematite; (b) Fe 2p peaks of the as-grown hematite; (c) O 1s peaks of the as-grown hematite; (df) survey spectra, Fe 2p peaks, and O 1 s peaks of α-Fe2O3/FeOOH, respectively.
Figure 3. X-ray photoelectron spectroscopy analysis: (a) survey spectra of the as-grown hematite; (b) Fe 2p peaks of the as-grown hematite; (c) O 1s peaks of the as-grown hematite; (df) survey spectra, Fe 2p peaks, and O 1 s peaks of α-Fe2O3/FeOOH, respectively.
Micromachines 15 00387 g003
Figure 4. J-V curves of hematite and α-Fe2O3/FeOOH(x) samples. (a) The α-Fe2O3/FeOOH(x) samples were immersed in 5 mM FeCl3 for different times: α-Fe2O3/FeOOH(1): 10 min; α-Fe2O3/FeOOH(2): 20 min; α-Fe2O3/FeOOH(3): 30 min; α-Fe2O3/FeOOH(4): 50 min; α-Fe2O3/FeOOH(5): 60 min. (b) The α-Fe2O3/FeOOH(x) samples were immersed in 10 mM FeCl3 for different times: α-Fe2O3/FeOOH(a): 10 min; α-Fe2O3/FeOOH(b): 20 min; α-Fe2O3/FeOOH(c): 30 min; α-Fe2O3/FeOOH(d): 50 min; α-Fe2O3/FeOOH(e): 60 min. PEC water splitting experiments were carried out in 1 M NaOH electrolyte under Xenon lamp with illumination irradiance of 225 mW cm−2. Solid and dashed lines represent illuminated and dark conditions, respectively.
Figure 4. J-V curves of hematite and α-Fe2O3/FeOOH(x) samples. (a) The α-Fe2O3/FeOOH(x) samples were immersed in 5 mM FeCl3 for different times: α-Fe2O3/FeOOH(1): 10 min; α-Fe2O3/FeOOH(2): 20 min; α-Fe2O3/FeOOH(3): 30 min; α-Fe2O3/FeOOH(4): 50 min; α-Fe2O3/FeOOH(5): 60 min. (b) The α-Fe2O3/FeOOH(x) samples were immersed in 10 mM FeCl3 for different times: α-Fe2O3/FeOOH(a): 10 min; α-Fe2O3/FeOOH(b): 20 min; α-Fe2O3/FeOOH(c): 30 min; α-Fe2O3/FeOOH(d): 50 min; α-Fe2O3/FeOOH(e): 60 min. PEC water splitting experiments were carried out in 1 M NaOH electrolyte under Xenon lamp with illumination irradiance of 225 mW cm−2. Solid and dashed lines represent illuminated and dark conditions, respectively.
Micromachines 15 00387 g004
Figure 5. Comparison of photocurrent densities under different applied bias potential for α-Fe2O3, α-Fe2O3/FeOOH(3), and α-Fe2O3/FeOOH(c), respectively.
Figure 5. Comparison of photocurrent densities under different applied bias potential for α-Fe2O3, α-Fe2O3/FeOOH(3), and α-Fe2O3/FeOOH(c), respectively.
Micromachines 15 00387 g005
Figure 6. (a) Mott-Schottky plots of α-Fe2O3 and α-Fe2O3/FeOOH. Mott-Schottky measurements were performed in 1 M NaOH solution at a frequency of 1 kHz under dark environment. (b) Electrochemical impedance spectroscopy (EIS) of hematite. EIS measurements were performed in 1 M NaOH at 0.5 V and in the frequency ranging from 10,000 to 1 Hz. Black circles represent α-Fe2O3 photoanode, and blue circles represent α-Fe2O3/FeOOH photoanode.
Figure 6. (a) Mott-Schottky plots of α-Fe2O3 and α-Fe2O3/FeOOH. Mott-Schottky measurements were performed in 1 M NaOH solution at a frequency of 1 kHz under dark environment. (b) Electrochemical impedance spectroscopy (EIS) of hematite. EIS measurements were performed in 1 M NaOH at 0.5 V and in the frequency ranging from 10,000 to 1 Hz. Black circles represent α-Fe2O3 photoanode, and blue circles represent α-Fe2O3/FeOOH photoanode.
Micromachines 15 00387 g006
Figure 7. Schematic representation of charge separation and transfer on α-Fe2O3/FeOOH nanoflakes photoanode.
Figure 7. Schematic representation of charge separation and transfer on α-Fe2O3/FeOOH nanoflakes photoanode.
Micromachines 15 00387 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, W.; Zhang, Y.; Miao, X.; Zhao, L.; Zhu, C. Deposition of FeOOH Layer on Ultrathin Hematite Nanoflakes to Promote Photoelectrochemical Water Splitting. Micromachines 2024, 15, 387. https://doi.org/10.3390/mi15030387

AMA Style

Zhang W, Zhang Y, Miao X, Zhao L, Zhu C. Deposition of FeOOH Layer on Ultrathin Hematite Nanoflakes to Promote Photoelectrochemical Water Splitting. Micromachines. 2024; 15(3):387. https://doi.org/10.3390/mi15030387

Chicago/Turabian Style

Zhang, Wenyao, Ya Zhang, Xiao Miao, Ling Zhao, and Changqing Zhu. 2024. "Deposition of FeOOH Layer on Ultrathin Hematite Nanoflakes to Promote Photoelectrochemical Water Splitting" Micromachines 15, no. 3: 387. https://doi.org/10.3390/mi15030387

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

Article Metrics

Back to TopTop