Next Article in Journal
Ligand Binding Path Sampling Based on Parallel Cascade Selection Molecular Dynamics: LB-PaCS-MD
Previous Article in Journal
Adsorption of Pollutants from Colored Wastewaters after Natural Wool Dyeing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Converting Sewage Water into H2 Fuel Gas Using Cu/CuO Nanoporous Photocatalytic Electrodes

1
Physics Department, College of Science, Jouf University, Sakaka P.O. Box 2014, Saudi Arabia
2
Basic Sciences Research Unit, Jouf University, Sakaka P.O. Box 2014, Saudi Arabia
3
Nanoscale Science, University of North Carolina at Charlotte, Charlotte, NC 28223, USA
4
State Key Lab of Metastable Materials Science and Technology, Yanshan University, Qinhuangdao 066004, China
5
Nanophotonics and Applications Lab, Physics Department, Faculty of Science, Beni-Suef University, Beni-Suef 62514, Egypt
6
Department of Physics, Faculty of Science, Islamic University of Madinah, Prince Naifbin Abdulaziz, Al Jamiah, Madinah 42351, Saudi Arabia
7
Physics Department, Faculty of Science, Sohag University, Sohag 82524, Egypt
8
School of Photovoltaics and Renewable Energy Engineering, University of New South Wales, Sydney, NSW 2052, Australia
9
Electrical Engineering Department, College of Engineering and Petroleum, Kuwait University, Safat 13113, Kuwait
10
Nanomaterials Science Research Laboratory, Chemistry Department, Faculty of Science, Beni-Suef University, Beni-Suef 62514, Egypt
*
Authors to whom correspondence should be addressed.
Materials 2022, 15(4), 1489; https://doi.org/10.3390/ma15041489
Submission received: 13 January 2022 / Revised: 5 February 2022 / Accepted: 10 February 2022 / Published: 16 February 2022

Abstract

:
This work reports on H2 fuel generation from sewage water using Cu/CuO nanoporous (NP) electrodes. This is a novel concept for converting contaminated water into H2 fuel. The preparation of Cu/CuO NP was achieved using a simple thermal combustion process of Cu metallic foil at 550 °C for 1 h. The Cu/CuO surface consists of island-like structures, with an inter-distance of 100 nm. Each island has a highly porous surface with a pore diameter of about 250 nm. X-ray diffraction (XRD) confirmed the formation of monoclinic Cu/CuO NP material with a crystallite size of 89 nm. The prepared Cu/CuO photoelectrode was applied for H2 generation from sewage water achieving an incident to photon conversion efficiency (IPCE) of 14.6%. Further, the effects of light intensity and wavelength on the photoelectrode performance were assessed. The current density (Jph) value increased from 2.17 to 4.7 mA·cm−2 upon raising the light power density from 50 to 100 mW·cm−2. Moreover, the enthalpy (ΔH*) and entropy (ΔS*) values of Cu/CuO electrode were determined as 9.519 KJ mol−1 and 180.4 JK−1·mol−1, respectively. The results obtained in the present study are very promising for solving the problem of energy in far regions by converting sewage water to H2 fuel.

1. Introduction

Photocatalytic materials represent an important class of components for potential applications in renewable-energy-related fields such as solar cells, optoelectronic, and photocatalytic H2 production [1,2,3,4]. On the one hand, the production of H2 gas from sewage water is a very promising field of renewable energy. This process provides H2 gas fuel for different uses of normal life such as cooking and warming, especially in remote regions inside deserts or rural areas. Moreover, this H2 gas is used as a fuel for airplanes and aircrafts, in addition to its normal utilization in industrial factories and companies. On the other hand, this photocatalytic reaction removes contamination (sewage water) through conversion into fuel. In addition, the H2 fuel is clean energy; its combustion has no side effects. The dependence of the world on this clean energy reduces the use of fossil fuels with their harmful produced gases such as CO2, NOx, and SOx [5,6,7].
The photocatalysts applied for H2 generation reaction must be semiconducting materials such as metal oxides, sulfides, nitrides, or organic [8,9]. Metal oxides have many benefits that qualified them to be ideal photocatalytic materials for H2 production such as low cost, stability, and easy preparation [10,11]. There are some methods for enhancing the photocatalytic activity such as increasing the surface area (nanofibers, nanowires, and nanotubes) [12,13,14]. Another way to enhance the photocatalytic activity is through the use of plasmonic materials [15,16], or materials with high thermal capacity such as Cu metal. These materials are effective for light capture and cause electron localization phenomena extended over the composite semiconductor materials, after which the neighbor materials use these phenomena for the H2 production process [17]. Previous studies examined the effect of Cu as a plasmonic material in Cu/ZnO for light capture and enhancement of the photocatalytic properties of ZnO [18].
CuO and Cu2O are promising materials for renewable energy applications, owing to their bandgap values of 0.7–1.6 eV and 2.0–2.2 eV for CuO and Cu2O, respectively [19,20]. These low bandgap values enable these materials to absorb most of the sunlight, which is preferable to large bandgap semiconductors that can absorb 10% to 20% of sunlight [21]. Moreover, the lower bandgap and high absorption efficiency of CuO favor its application (compared with Cu2O) in the renewable energy field [22].
In 2014, Li et al. managed to synthesize a nanoporous CuO layer onto Cu foil through the annealing of Cu (OH)2 nanowires at 500 °C under an oxygen flow. A thick Cu2O interlayer was also formed; this annealing process formed under high oxygen pressure, which is higher than CuO dissociation pressure, so the CuO layer formed at the outer surface of the structure. The Cu/Cu2O/CuO structure is used as an electrode for glucose sensing. Sagadevan et al. prepared CuO nanoparticles via the combustion technique for various annealing temperatures (100 °C and 300 °C), with ascorbic acid used as a capping agent [23]. Ragupathi et al. prepared CuO/g-C3N4 for a water-splitting reaction, but the produced current density (Jph) and the incident photon to current conversion efficiency (IPCE) were very small [24]. Quyen et al. studied the effect of Cu on TiO2 for the photocatalytic water splitting reaction and determined that Cu nanoparticles increased the photocatalytic effect very much and increased the rate of H2 generation, which resulted from increasing oxygen vacancies in TiO2 and the charge transfer process [24]. Shen et al. prepared graphite carbon nitride and decorated this material with CuO for photocatalyst application, and studied the effect of CuO for increasing the efficiency of the catalytic reaction and H2 generation, but the rate of H2 generation was small [25].
Most of the previous studies on CuO relied on the use of an additional sacrificing agent such as Na2SO3, Na2S2O3, HCl, and NaOH [26,27,28]. Moreover, the H2 rate production was limited [29,30,31,32]. Various fabrication methods of CuO have been reported such as physical sputtering, atomic layer deposition, spray pyrolysis, radiofrequency sputtering, and so on. However, these techniques have drawbacks of high-cost, long fabrication time, and complex fabrication processes [33,34].
This study provides the H2 gas fuel from the contamination (sewage water) without using any additional electrolyte. The catalytic electrode is prepared with a low-cost combustion method without using any complex techniques. The prepared electrode has a high Jph in comparison with the previous literature. The produced efficiency is promising for application of this electrode in H2 generation from sewage water in industrial applications.
In this study, Cu/CuO photocathode, prepared by an oxidation/combustion process, was evaluated for H2 production from a sewage water-splitting reaction without using any sacrificing agent. The influence of various factors such as light intensity and wavelength, temperature, and on/off chopped current was assessed on the photocathode performance. Finally, IPCE was estimated under different monowavelength light, and the mechanism for the sewage water splitting reaction was examined.

2. Materials and Methods

2.1. Cu/CuO Nanomaterial Preparation

Prior to sample preparation, the Cu foil was cleaned using water, soap, acetone, and ethanol under ultrasonication for 10 min each. The preparation of Cu/CuO was carried out through oxidation combustion of copper foil (99.9%, thickness 0.3 mm) in Nabertherm box furnace (Nitrex Metal Inc., St-Laurent, QC, Canada) at 550 °C for 1 h. Through this combustion process, the Cu metal is oxidized to generate CuO NP material. Then 1 cm2 of the Cu/CuO was used as electrode for hydrogen generation under wastewater splitting reaction through electrochemical measurements.

2.2. Characterization

X-ray diffraction pattern (XRD, Malvern Panalytical Ltd., Malvern, UK) analyses were carried out using a Bruker D8 advance diffractometer using Cu Kα radiation (wavelength = 0.15418 nm). Field-emission scanning electron microscopy (FE-SEM, Hitachi, S-4800, Schaumburg, IL, USA) was used to assess the morphology. The energy dispersive X-ray analysis (EDAX, Hitachi, Schaumburg, IL, USA) elemental composition and elemental mapping were examined using the EDAX unit attached to the FE-SEM. The optical properties were examined using a double beam spectrophotometer (Perkin Elmer Lamba 950, Shelton, CT, USA).

2.3. Electrochemical Measurements

The H2 generation reaction was carried out from sewage water solution (100 mL, pH 5.5) using a three-electrode cell, in which Cu/CuO nanomaterial (1cm2), a graphite sheet of the same dimensions, and calomel act as working, counter, and reference electrodes, respectively, as shown in Figure 1. All measurements were carried out using a workstation (CHI660E, Tennison Hill Drive, Austin, TX, USA) in a potential range from −1 to 1 V, a xenon lamp acts as a solar simulator. Some parameters were studied such as the light intensity (25 to 100 mW·cm−2), light wavelength (390 to 636 nm), on/off chopped current, and temperature on the electrode performance.
The sewage water was obtained from the drinking water and sanitation of Beni Suef city, Egypt, and the construction was confirmed using gas chromatography–mass spectrometry.

3. Results and Discussion

Morphological, Structural, and Optical Properties

The morphology of the CuO NP, prepared from Cu metal using a simple combustion process in air at 550 °C for 1 h, is displayed in Figure 2a,b. The CuO NP substrate consists of island-like structures with an inter-distance of 100 nm. Each island is highly porous with a pore diameter of about 250 nm. The formation of these pores with high homogeneous distribution on the CuO surface enlarges the surface area. This feature is beneficial for the photocatalytic process through high light absorption efficiency, in which the porous surface acts as a cave for light absorption [7,12].
Figure 2c shows the XRD pattern of the Cu/CuO sample showing the two characteristic peaks of CuO at around 35.6° and 38.5° (JCPDS #41–0254), indicating that CuO NP were formed on the copper foil [35,36,37]. The other three strong diffraction peaks at 43.5°, 50.7°, and 74.5° correspond to the (111), (200), and (220) reflections, respectively, of the face-centered-cubic Cu (JCPDS #02–1225) [36,38]. The crystal size of the CuO nanomaterials is calculated using the Scherrer equation [39,40,41], Equation (1):
D = 0.94 λ / ß cos θ
where ß is the full width at half maximum (), λ is the X-ray wavelength (CuKα = 0.154 nm), and θ is the Bragg’s angle [42]. From the equation, the average crystal size of CuO is about 89 nm.
The EDAX spectrum (Figure 2d) and EDAX mapping images (Figure 2e) confirm the formation of the CuO NP material where Cu and O peaks are obviously present. The Cu and O elements are homogeneously distributed over the whole examined area.
The optical diffuse reflectance of the CuO NP was determined using a double beam spectrophotometer, as shown in Figure 2f. It is evident that the prepared CuO NP has a high light absorption behavior in a wide optical range (Vis and near IR). This is related to the low reflectance behavior of CuO NP in these regions. The bandgap is calculated using the Kubelka–Munk equation (Equation (2)) [12], as shown in the insert of Figure 2f, where K is the molar absorption coefficient, and S is the scattering factor. From Equation (2), a bandgap of 1.38 eV was determined for the CuO thin film, which is in good agreement with a recent research study [43].
K / S = ( 1 R ) 2 2 R
The cross section and roughness is estimated using the modeling program (Image J) as shown in Figure 2g,h. From this figure, the surface roughness appears well with a surface area of 235 µm2 in 38 µm2.
The photoelectrochemical (PEC) activity of the Cu/CuO photoelectrode was assessed in sewage water (chemical composition in Table 1) at room temperature (25 °C) with a sweep rate of 1 mV/s under xenon lamp illumination. The PEC measurements were performed in dark and light without optical filters, as shown in Figure 3a. The broad surface area of the prepared nanotextured electrode generates a high density of electron-hole pairs when exposed to light, leading to the splitting of H2O molecules to conduct the hydrogen electro-generation reaction. The effect of light on the Cu/CuO photoelectrode generate Jph values of −0.07 and 4.7 mA·cm−2 at 0 and 1 V, respectively (Figure 3a), although the density of the dark current is very small for the electrode and can be ignored. Therefore, from the Jph values, the Cu/CuO with the lowest photogeneration voltage (0.56 V) is a highly efficient electrode for water splitting and H2 gas generation.
The substrate Cu metal contributes to a high density of pairs of electrons formed. This will motivate H2O molecules to be broken to produce hydrogen when it reaches the CuO surface. The spectral overlap between CuO absorbance oscillator frequencies and the Cu metal oscillator frequency improves the produced photocurrent.
The stability (time-Jph relation) of the prepared Cu/CuO photoelectrode was studied as presented in Figure 3b. At +0.1 V, the produced Jph value under on and off chopped light is mentioned. From the figure, the electrode has high stability and sensitivity to light. From the magnified part in Figure 3c, the change in the Jph values under on and off chopped light appears well. This confirms the high sensitivity of the electrode to light due to the high effect of the light on the electrode. The reproducibility of the electrode for four runs is shown in Figure 3d, in which the voltage–current relation shows the same behavior until four runs. This reproducibility was carried out at 30 °C and in normal light. The standard deviation (SD) value for the Cu/CuO photoelectrode is 0.3%.
The effect of light intensity (50 to 100 mW·cm−2) on the Cu/CuO photoelectrode is mentioned in Figure 4a,b. This effect appears clearly, in which the Jph values increase from 2.17 to 4.7 mA·cm−2, directly with the light intensity until 100 mW·cm−2. This increase is related to the electron–hole pair formation under the increasing light intensity [44], in which many photons activate the active sites on the photocatalytic materials [45]. The Jph represents the electrons collected on the surface of the photoelectrode; this Jph then represented the water splitting and H2 generation rate [46,47].
The number of photons (N) is directly proportional to the light intensity (P) as shown in Equation (3). This equation depends on other factors such as wavelength ( λ ), Planck constant (h), and light velocity (c). From this equation, the N per second is changed from 4 × 1021 to 8 × 1021 photon/s under light intensity from 50 to 100, mW·cm−2, respectively.
N = λ P / hc
The effect of the light monowavelength (390 to 636 nm) on the produced Jph for Cu/CuO electrodes is presented in Figure 5a. From this figure, the Jph has varied values under the monochromatic light effect. The optimum Jph value is 4.6 mA·cm−2 at 390 nm, in which these values correspond to the optical analyses data (Figure 3). The inset figure in Figure 6a shows this relation clearly.
The incident photon to current conversion efficiency (IPCE) represents the electrons collected on the surface of the photocatalytic materials under the photon flux (Figure 5b). This IPCE value can be calculated from the wavelength values [48], through Equation (4).
IPCE = J ph ( mA · cm 2 ) · 1240   ( V · nm ) P ( mW · cm 2 ) · λ ( nm )
The IPCE is determined at 100 mW·cm−2 for the photoelectrode Cu/CuO and presented in Figure 5b. The IPCE values very much depend on monochromatic light, in which the optimum IPCE value is 14.6% at 390 nm. The IPCE values decrease with increasing wavelengths. Therefore, the photocatalytic electrode has the ability for sewage water splitting and H2 generation with 14.6% efficiency. This high value was produced without using any additional electrolyte and infers that the electrode converts the sewage water into H2 and O2 with high efficiency in comparison with other previous literature [29,49,50,51,52,53].
There is an inverse relationship between the number of photons and IPCE. The optimum values occur at low wavelengths, in which the light has a high frequency for transferring most electrons to the conducting band, so the Jph value increases, and thereby the H2 production increases [54].
The water splitting reaction for the H2 generation process using the Cu/CuO photoelectrode under different temperatures is mentioned in Figure 6a. The Jph values increase from 4.7 to 8.8 mA·cm−2 with increasing the temperature from 30 to 60 °C, respectively. This increasing behavior of the Jph indicates the high mobility of the sewage water ions with increasing temperature, in which the high ionic mobility facilitates the H2 and O2 generation in both sides of the electrochemical cell. So increasing the Jph values with the temperature indicates an increase in the H2 generation rate [55,56].
The activation energy (Ea) for the H2 generation can be calculated under different temperatures using the Arrhenius equation (Equation (5)) [57]. This Ea depends on the rate of collisions (k is the rate constant) and temperature values in Kelvin (T) using the universal gas constant (R) and Arrhenius constant (A) as the standard constant. From the Ea value, the initiator temperature for starting the splitting reaction is determined [58,59,60]. Ea values are obtained from slope of ln Jph versus 1/T as shown in Figure 6c. The Ea value is 11.8 KJ mol−1 for the electrode. The small value of the water splitting reaction indicates the splitting occurs easily for the H2 and O2 evolution [61].
For calculating the enthalpy (ΔH*) and entropy (ΔS*), Equation (6) is applied (Eyring equation) [62,63]. This equation is similar to the Arrhenius equation, but it contains additional parameters and uses additional constants, in which kB is the Boltzman constant and h is the Planck constant. By applying this equation for the sewage splitting reaction of the Cu/CuO photoelectrode, ΔH* and ΔS* values are obtained as 9.519 kJ mol−1 and 180.4 JK−1·mol−1, respectively (Figure 6c).
Moreover, the H2 moles are calculated from the Faraday equation, Equation (7) [64,65], under time change (dt), using the Faraday constant (F). The estimated H2 mole for the Cu/CuO photoelectrode is 60 μmol h−1 cm−2. In addition, a comparison between the present study and the previously reported literature is added in Table 2.
k = Ae Ea / RT
k = T · kB h · e Δ S / R · e Δ H / RT
H 2   mole = 0 t J ph · dt / F
The mechanism of the CuO photocatalytic materials for the sewage water-splitting reaction is based on the effect of incident light on the motivation of the photoelectrons from the CuO NP material that resulted from the energy band changes (Figure 7). The photoelectrons generated under the effect of light incidence have two steps. First, electron-hole generation, in which the generated electrons leave the holes and transfer to the upper level. The second step is the localized surface plasmonic resonance (LSPR); this resonance process causes the energy transfer [66]. These two steps occurred easily due to the small CuO band gap of 1.39 eV in addition to the absence of depletion regions inside the Cu and CuO nanomaterials. Therefore, the results are the transfer of electrons from the Cu to CuO without any restrictions, and the continuous electrons flow is carried out without any restrictions [67]. These hot electrons cause the generation of Jph under the applied potential [68,69,70]. The experimental image of electrons transfer processes is represented in the optical analyses (Figure 3) and the electrochemical measurements under various effects such as light intensity and light wavelengths. The Cu metal substrate has plasmonic properties that cause the light capture and electron resonance process [52] that motivates the CuO nanomaterials and generates electrons over their surface [71,72,73]. These electrons are represented in Jph values and the H2 generation reaction rate [74,75].

4. Conclusions

This work provides promising results in support of H2 production from sewage water using a CuO NP photoelectrode. All the analyses were carried out for confirming the chemical structure, morphology, and optical properties of the prepared CuO NP. From the SEM, the prepared materials have nanoporous features that look like small islands with diameters of about 300–400 nm, with each island composed of a package of small nanoporous particles. XRD confirmed the monoclinic CuO crystalline phase with crystallite size of 89 nm. The obtained optical band gap value for CuO NP was 1.39 eV. The sewage water was used as a source of H2 gas produced by the Cu/CuO photoelectrode. The Jph value was changed from 2.17 to 4.7 mA·cm−2 upon increasing the light power density from 50 to 100 mW·cm−2, respectively. The IPCE was changed under the effect of monochromatic light, in which the optimum IPCE value was 14.6% at 390 nm. Moreover, the effect of on and off chopped current was studied that confirms the motivation of the photocatalyst under light incidence. The sewage water splitting thermodynamics were studied, in which the enthalpy (ΔH*) and entropy (ΔS*) values for the Cu/CuO electrode were 9.519 kJ·mol−1 and 180.4 JK−1·mol−1, respectively. A mechanism was proposed to explain the relationship between the incidence light and the Jph values of the H2 generation rate.

Author Contributions

Conceptualization, A.A.A.A., N.M.A.H., M.A., A.H. and M.R.; methodology, software, validation, formal analysis, investigation, resources, data curation, writing—original draft preparation, visualization, supervision, and project administration, S.H.M., B.H., A.A.A.A., M.R., M.S., A.H. and N.M.A.H.; synthetized photoelectrode, S.H.M., B.H., A.A.A.A., M.R. and B.H.; writing—review and editing, A.A.A.A., M.R., N.M.A.H., M.A., M.R., A.H. and M.S.; funding, N.M.A.H., A.H. and M.A. All authors have read and agreed to the published version of the manuscript.

Funding

The authors extend their appreciation to the Deanship of Scientific Research at Jouf University for funding this work through a research grant (no. DSR-2021-03-0318). Also, we would like to thank Kuwait University for their technical and financial support on this work.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Nishiyama, H.; Yamada, T.; Nakabayashi, M.; Maehara, Y.; Yamaguchi, M.; Kuromiya, Y.; Nagatsuma, Y.; Tokudome, H.; Akiyama, S.; Watanabe, T.; et al. Photocatalytic solar hydrogen production from water on a 100-m2 scale. Nature 2021, 598, 304–307. [Google Scholar] [CrossRef]
  2. Domen, K.; Hisatomi, T. Reaction systems for solar hydrogen production via water splitting with particulate semiconductor. Nat. Catal. 2019, 2, 387–399. [Google Scholar] [CrossRef]
  3. Pagliaro, M. Preparing for the future: Solar energy and bioeconomy in the United Arab Emirates. Energy Sci. Eng. 2019, 7, 1451–1457. [Google Scholar] [CrossRef] [Green Version]
  4. Takata, T. Photocatalytic water splitting with quantum efficiency of almost unity. Nature 2020, 581, 411–414. [Google Scholar] [CrossRef]
  5. Mohamed, F.; Rabia, M.; Shaban, M. Synthesis and characterization of biogenic iron oxides of different nanomorphologies from pomegranate peels for efficient solar hydrogen production. J. Mater. Res. Technol. 2020, 9, 4255–4271. [Google Scholar] [CrossRef]
  6. Shaban, M.; Ali, S.; Rabia, M. Design and application of nanoporous graphene oxide film for CO2, H2, and C2H2 gases sensing. J. Mater. Res. Technol. 2019, 8, 4510–4520. [Google Scholar] [CrossRef]
  7. Elsayed, A.M.; Rabia, M.; Shaban, M.; Aly, A.H.; Ahmed, A.M. Preparation of hexagonal nanoporous Al2O3/TiO2/TiN as a novel photodetector with high efficiency. Sci. Rep. 2021, 11, 17572. [Google Scholar]
  8. Kang, Z.; Cheng, Y.; Zheng, Z.; Cheng, F.; Chen, Z.; Li, L.; Tan, X.; Xiong, L.; Zhai, T.; Gao, Y. MoS2-Based photodetectors powered by asymmetric contact structure with large work function difference. Nano-Micro Lett. 2019, 11, 1–12. [Google Scholar] [CrossRef] [Green Version]
  9. Lee, J.H.; Lee, W.W.; Yang, D.W.; Chang, W.J.; Kwon, S.S.; Park, W. Il anomalous photovoltaic response of graphene-on-GaN schottky photodiodes. ACS Appl. Mater. Interfaces 2018, 10, 14170–14174. [Google Scholar] [CrossRef] [PubMed]
  10. Abdelazeez, A.A.A.; El-Fatah, G.A.; Shaban, M.; Ahmed, A.M.; Rabia, M. ITO/Poly-3-Methylaniline/Au Electrode for Electrochemical Water Splitting and Dye Removal. ECS J. Solid State Sci. Technol. 2021, 10, 123009. [Google Scholar] [CrossRef]
  11. Rabia, M.; Mohamed, H.S.H.; Shaban, M.; Taha, S. Preparation of polyaniline/PbS core-shell nano/microcomposite and its application for photocatalytic H2 electrogeneration from H2O. Sci. Rep. 2018, 8, 1107. [Google Scholar] [CrossRef] [Green Version]
  12. Mohamed, H.S.H.; Rabia, M.; Zhou, X.G.; Qin, X.S.; Khabiri, G.; Shaban, M.; Younus, H.A.; Taha, S.; Hu, Z.Y.; Liu, J.; et al. Phase-junction Ag/TiO2 nanocomposite as photocathode for H2 generation. J. Mater. Sci. Technol. 2021, 83, 179–187. [Google Scholar] [CrossRef]
  13. Nikolaidis, P.; Poullikkas, A. A comparative overview of hydrogen production processes. Renew. Sustain. Energy Rev. 2017, 67, 597–611. [Google Scholar] [CrossRef]
  14. Almohammedi, A.; Shaban, M.; Mostafa, H.; Rabia, M. Nanoporous TiN/TiO2/Alumina Membrane for Photoelectrochemical Hydrogen Production from Sewage Water. Nanomaterials 2021, 11, 2617. [Google Scholar] [CrossRef]
  15. Zhang, X.; Li, J.; Yang, W.; Leng, B.; Niu, P.; Jiang, X.; Liu, B. High-performance flexible ultraviolet photodetectors based on AZO/ZnO/PVK/PEDOT:PSS heterostructures integrated on human hair. ACS Appl. Mater. Interfaces 2019, 11, 24459–24467. [Google Scholar] [CrossRef] [PubMed]
  16. Kim, J.; Lee, H.C.; Kim, K.H.; Hwang, M.S.; Park, J.S.; Lee, J.M.; So, J.P.; Choi, J.H.; Kwon, S.H.; Barrelet, C.J.; et al. Photon-triggered nanowire transistors. Nat. Nanotechnol. 2017, 12, 963–968. [Google Scholar] [CrossRef]
  17. Zhang, X.; Liu, Q.; Liu, B.; Yang, W.; Li, J.; Niu, P.; Jiang, X. Giant UV photoresponse of a GaN nanowire photodetector through effective Pt nanoparticle coupling. J. Mater. Chem. C 2017, 5, 4319–4326. [Google Scholar] [CrossRef]
  18. Li, M.; Yu, M.; Su, D.; Zhang, J.; Jiang, S.; Wu, J.; Wang, Q.; Liu, S. Ultrahigh responsivity UV photodetector based on Cu nanostructure/ZnO QD hybrid architectures. Small 2019, 15, 1901606. [Google Scholar] [CrossRef] [PubMed]
  19. Schultze, J.W.; Lohrengel, M.M. Stability, reactivity and breakdown of passive films. Problems of recent and future research. Electrochim. Acta 2000, 45, 2499–2513. [Google Scholar] [CrossRef]
  20. Basnet, P.; Zhao, Y. Tuning the CuxO nanorod composition for efficient visible light induced photocatalysis. Catal. Sci. Technol. 2016, 6, 2228–2238. [Google Scholar] [CrossRef]
  21. Chiang, C.Y.; Shin, Y.; Aroh, K.; Ehrman, S. Copper oxide photocathodes prepared by a solution based process. Int. J. Hydrogen Energy 2012, 37, 8232–8239. [Google Scholar] [CrossRef]
  22. Chiang, C.-Y.; Chang, M.-H.; Liu, H.-S.; Tai, C.Y.; Ehrman, S. Process Intensification in the production of photocatalysts for solar hydrogen generation. Ind. Eng. Chem. Res. 2012, 51, 5207–5215. [Google Scholar] [CrossRef]
  23. Ragupathi, V.; Raja, M.A.; Panigrahi, P.; Ganapathi Subramaniam, N. CuO/g-C3N4 nanocomposite as promising photocatalyst for photoelectrochemical water splitting. Optik 2020, 208, 164569. [Google Scholar] [CrossRef]
  24. Thi Quyen, V.; Jitae, K.; Thi Huong, P.; Thi Thu Ha, L.; My Thanh, D.; Minh Viet, N.; Quang Thang, P. Copper doped titanium dioxide as a low-cost visible light photocatalyst for water splitting. Sol. Energy 2021, 218, 150–156. [Google Scholar] [CrossRef]
  25. Chen, J.; Shen, S.; Guo, P.; Wang, M.; Wu, P.; Wang, X.; Guo, L. In-situ reduction synthesis of nano-sized Cu2O particles modifying g-C3N4 for enhanced photocatalytic hydrogen production. Appl. Catal. B Environ. 2014, 152–153, 335–341. [Google Scholar] [CrossRef]
  26. Teixeira, G.F.; Silva Junior, E.; Vilela, R.; Zaghete, M.A.; Colmati, F. Perovskite Structure Associated with Precious Metals: Influence on Heterogenous Catalytic Process. Catalysts 2019, 9, 721. [Google Scholar] [CrossRef] [Green Version]
  27. Mishra, M.; Chun, D.M. α-Fe2O3 as a photocatalytic material: A review. Appl. Catal. A Gen. 2015, 498, 126–141. [Google Scholar]
  28. Acar, C.; Dincer, I.; Naterer, G.F. Review of photocatalytic water-splitting methods for sustainable hydrogen production. Int. J. Energy Res. 2016, 40, 1449–1473. [Google Scholar]
  29. Liu, G.; Karuturi, S.K.; Chen, H.; Wang, D.; Ager, J.W.; Simonov, A.N.; Tricoli, A. Enhancement of the photoelectrochemical water splitting by perovskite BiFeO3 via interfacial engineering. Sol. Energy 2020, 202, 198–203. [Google Scholar] [CrossRef]
  30. Wang, Z.; Cao, D.; Wen, L.; Xu, R.; Obergfell, M.; Mi, Y.; Zhan, Z.; Nasori, N.; Demsar, J.; Lei, Y. Manipulation of charge transfer and transport in plasmonic-ferroelectric hybrids for photoelectrochemical applications. Nat. Commun. 2016, 7, 1–8. [Google Scholar] [CrossRef] [Green Version]
  31. Davies, J.; du Preez, S.P.; Bessarabov, D.G. On-Demand Hydrogen Generation by the Hydrolysis of Ball-Milled Aluminum& ndash; Bismuth – Zinc Composites. Materials 2022, 15, 1197. [Google Scholar] [CrossRef] [PubMed]
  32. Chen, M.; Chou, S.-F.; Blaabjerg, F.; Davari, P. Overview of Power Electronic Converter Topologies Enabling Large-Scale Hydrogen Production via Water Electrolysis. Appl. Sci. 2022, 12, 1906. [Google Scholar] [CrossRef]
  33. Chiang, C.Y.; Aroh, K.; Franson, N.; Satsangi, V.R.; Dass, S.; Ehrman, S. Copper oxide nanoparticle made by flame spray pyrolysis for photoelectrochemical water splitting—Part II. Photoelectrochemical study. Int. J. Hydrogen Energy 2011, 36, 15519–15526. [Google Scholar] [CrossRef]
  34. Guo, X.; Diao, P.; Xu, D.; Huang, S.; Yang, Y.; Jin, T.; Wu, Q.; Xiang, M.; Zhang, M. CuO/Pd composite photocathodes for photoelectrochemical hydrogen evolution reaction. Int. J. Hydrogen Energy 2014, 39, 7686–7696. [Google Scholar] [CrossRef]
  35. Zúñiga, A.; Fonseca, L.; Souza, J.A.; Rivaldo-Gomez, C.; Pomar, C.D.; Criado, D. Anomalous ferromagnetic behavior and size effects in CuO nanowires. J. Magn. Magn. Mater. 2019, 471, 77–81. [Google Scholar] [CrossRef]
  36. Zhan, W.; Chen, Z.; Hu, J.; Chen, X. Vertical CuO nanowires array electrodes: Visible light sensitive photoelectrochemical biosensor of ethanol detection. Mater. Sci. Semicond. Process. 2018, 85, 90–97. [Google Scholar] [CrossRef]
  37. Mahmoodi, A.; Solaymani, S.; Amini, M.; Nezafat, N.B.; Ghoranneviss, M. Structural, morphological and antibacterial characterization of CuO nanowires. Silicon 2017, 10, 1427–1431. [Google Scholar] [CrossRef]
  38. Kim, J.-H.; Katoch, A.; Choi, S.-W.; Kim, S.S. Growth and sensing properties of networked p-CuO nanowires. Sensors Actuators B. Chem. 2015, 212, 190–195. [Google Scholar] [CrossRef]
  39. Shaban, M.; Rabia, M.; Fathallah, W.; El-Mawgoud, N.A.; Mahmoud, A.; Hussien, H.; Said, O. Preparation and characterization of polyaniline and Ag/polyaniline composite nanoporous particles and their antimicrobial activities. J. Polym. Environ. 2017, 26, 434–442. [Google Scholar] [CrossRef]
  40. Abukhadra, M.R.; Rabia, M.; Shaban, M.; Verpoort, F. Heulandite/polyaniline hybrid composite for efficient removal of acidic dye from water; kinetic, equilibrium studies and statistical optimization. Adv. Powder Technol. 2018, 29, 2501–2511. [Google Scholar] [CrossRef]
  41. Fadel, M.S.S.; Rabia, M.; Ezzat, S.; Mansour, N.; Saeed, E.; Sayyah, S.M. Effect of annealing temperature on VO2(M)/ITO film nanomaterials for thermochromic smart windows application and study its contact angle. J. Nanophotonics 2018, 12, 016009. [Google Scholar] [CrossRef]
  42. Ahmed, M.A.; Rabia, M.; Shaban, M. The structure and photoelectrochemical activity of Cr-doped PbS thin films grown by chemical bath deposition. RSC Adv. 2020, 10, 14458–14470. [Google Scholar] [CrossRef]
  43. Ravichandran, A.T.; Dhanabalan, K.; Vasuhi, A.; Chandramohan, R.; Mantha, S. Morphology, bandgap, and grain size tailoring in Cu2O thin film by SILAR method. IEEE Trans. Nanotechnol. 2015, 14, 108–112. [Google Scholar] [CrossRef]
  44. Yang, L.; Liu, Z. Study on light intensity in the process of photocatalytic degradation of indoor gaseous formaldehyde for saving energy. Energy Convers. Manag. 2007, 48, 882–889. [Google Scholar] [CrossRef]
  45. Baniasadi, E.; Dincer, I.; Naterer, G.F. Measured effects of light intensity and catalyst concentration on photocatalytic hydrogen and oxygen production with zinc sulfide suspensions. Int. J. Hydrogen Energy 2013, 38, 9158–9168. [Google Scholar] [CrossRef]
  46. Sayyah, S.M.; Shaban, M.; Rabia, M. A sensor of m-toluidine/m-cresol polymer film for detection of lead ions by potentiometric methods. Sens. Lett. 2016, 14, 522–529. [Google Scholar] [CrossRef]
  47. Sayyah, S.M.; Shaban, M.; Khaliel, A.B.; Rabia, M. Electropolymerization of M-Cresol on Platinum Electrode from Aqueous Acidic Solution. Int. J. Chem. 2014, 35, 2051–2732. [Google Scholar]
  48. Razek, S.A.; Popeil, M.R.; Wangoh, L.; Rana, J.; Suwandaratne, N.; Andrews, J.L.; Watson, D.F.; Banerjee, S.; Piper, L.F.J. Designing catalysts for water splitting based on electronic structure considerations. Electron. Struct. 2020, 2, 023001. [Google Scholar] [CrossRef]
  49. Freeman, E.; Kumar, S.; Thomas, S.R.; Pickering, H.; Fermin, D.J.; Eslava, S. PrFeO3 Photocathodes prepared through spray pyrolysis. ChemElectroChem 2020, 7, 1365–1372. [Google Scholar] [CrossRef]
  50. Rabia, M.; Shaban, M.; Jibali, B.M.; Abdelkhaliek, A.A. Effect of Annealing Temperature on the Photoactivity of ITO/VO2 (M)/Au Film Electrodes for Water Splitting. J. Nanosci. Nanotechnol. 2020, 20, 4120–4130. [Google Scholar] [CrossRef] [PubMed]
  51. Potashnikov, V.; Golub, A.; Brody, M.; Lugovoy, O. Decarbonizing Russia: Leapfrogging from Fossil Fuel to Hydrogen. Energies 2022, 15, 683. [Google Scholar] [CrossRef]
  52. Xiao, X.; Engelbrekt, C.; Zhang, M.; Li, Z.; Ulstrup, J.; Zhang, J.; Si, P. A straight forward approach to electrodeposit tungsten disulfide/poly(3,4-ethylenedioxythiophene) composites onto nanoporous gold for the hydrogen evolution reaction. Appl. Surf. Sci. 2017, 410, 308–314. [Google Scholar] [CrossRef]
  53. Modibane, K.D.; Waleng, N.J.; Ramohlola, K.E.; Maponya, T.C.; Monama, G.R.; Makgopa, K.; Hato, M.J. Poly(3-aminobenzoic acid) Decorated with Cobalt Zeolitic Benzimidazolate Framework for Electrochemical Production of Clean Hydrogen. Polymers 2020, 12, 1581. [Google Scholar] [CrossRef] [PubMed]
  54. Shaban, M.; Ahmed, A.M.; Abdel-Rahman, E.; Hamdy, H. Tunability and sensing properties of plasmonic/1d photonic crystal. Sci. Rep. 2017, 7, 1–10. [Google Scholar] [CrossRef] [Green Version]
  55. Sayyah, S.M.; Shaban, M.; Rabia, M. Electropolymerization of m-toluidin on platinum electrode from aqueous acidic solution and character of the obtained polymer. Adv. Polym. Technol. 2018, 37, 126–136. [Google Scholar] [CrossRef]
  56. Sayyah, E.S.M.; Shaban, M.; Rabia, M. A sensor of m-cresol nanopolymer/Pt-electrode film for detection of lead ions by potentiometric methods. Adv. Polym. Technol. 2018, 37, 1296–1304. [Google Scholar] [CrossRef]
  57. Arrhenius, S. Über die Dissociationswärme und den Einfluss der Temperatur auf den Dissociationsgrad der Elektrolyte. Z. Phys. Chem. 1889, 7, 1. [Google Scholar] [CrossRef] [Green Version]
  58. Awad, M.A.; Shaban, M.; Rabia, M. The efficiency of M (M = Li, Na, or Cs) doped CdS nanomaterials in optoelectronic applications. Int. J. Energy Res. 2022. [Google Scholar] [CrossRef]
  59. Rabia, M.; Mohamed, S.H.; Zhao, H.; Shaban, M.; Lei, Y.; Ahmed, A.M. TiO2/TiOxNY hollow mushrooms-like nanocomposite photoanode for hydrogen electrogeneration. J. Porous Mater. 2020, 27, 133–139. [Google Scholar] [CrossRef]
  60. Shaban, M.; Rabia, M.; Eldakrory, M.G.; Maree, R.M.; Ahmed, A.M. Efficient photoselectrochemical hydrogen production utilizing of APbI3 (A = Na, Cs, and Li) perovskites nanorods. Int. J. Energy Res. 2021, 45, 7436–7446. [Google Scholar] [CrossRef]
  61. Rabia, M.; Shaban, M.; Adel, A.; Abdel-Khaliek, A.A. Effect of plasmonic au nanoparticles on the photoactivity of polyaniline/indium tin oxide electrodes for water splitting. Environ. Prog. Sustain. Energy 2019, 38, 13171. [Google Scholar] [CrossRef]
  62. Eyring, H. The activated complex in chemical reactions. J. Chem. Phys. 1935, 3, 63–71. [Google Scholar] [CrossRef]
  63. Shaban, M.; Benghanem, M.; Almohammedi, A.; Rabia, M. Optimization of the Active Layer P3HT:PCBM for Organic Solar Cell. Coatings 2021, 11, 863. [Google Scholar] [CrossRef]
  64. Helmy, A.; Rabia, M.; Shaban, M.; Ashraf, A.M.; Ahmed, S.; Ahmed, A.M. Graphite/rolled graphene oxide/carbon nanotube photoelectrode for water splitting of exhaust car solution. Int. J. Energy Res. 2020, 44, 7687–7697. [Google Scholar] [CrossRef]
  65. Mohamed, H.S.H.; Rabia, M.; Shaban, M.; Taha, S. Controlled synthesis of CdS nanoflowers thin films for H2 electro-generation. Mater. Sci. Semicond. Process. 2020, 120, 105307. [Google Scholar] [CrossRef]
  66. Koya, A.N.; Zhu, X.; Ohannesian, N.; Yanik, A.A.; Alabastri, A.; Proietti Zaccaria, R.; Krahne, R.; Shih, W.C.; Garoli, D. Nanoporous metals: From plasmonic properties to applications in enhanced spectroscopy and photocatalysis. ACS Nano 2021, 15, 6038–6060. [Google Scholar] [CrossRef] [PubMed]
  67. Podder, S.; Pal, A.R. Plasmonic visible-NIR photodetector based on hot electrons extracted from nanostructured titanium nitride. J. Appl. Phys. 2019, 126, 083108. [Google Scholar] [CrossRef]
  68. Hussain, A.A.; Sharma, B.; Barman, T.; Pal, A.R. Self-powered broadband photodetector using plasmonic titanium nitride. ACS Appl. Mater. Interfaces 2016, 8, 4258–4265. [Google Scholar] [CrossRef] [PubMed]
  69. Hussain, A.A.; Pal, A.R.; Patil, D.S. An efficient fast response and high-gain solar-blind flexible ultraviolet photodetector employing hybrid geometry. Appl. Phys. Lett. 2014, 104, 193301. [Google Scholar] [CrossRef]
  70. Jiang, L.; Zhou, G.; Mi, J.; Wu, Z. Fabrication of visible-light-driven one-dimensional anatase TiO2/Ag heterojunction plasmonic photocatalyst. Catal. Commun. 2012, 24, 48–51. [Google Scholar] [CrossRef]
  71. Cao, Q.; Yu, J.; Cao, Y.; Delaunay, J.J.; Che, R. Unusual effects of vacuum annealing on large-area Ag3PO4 microcrystalline film photoanode boosting cocatalyst- and scavenger-free water splitting. J. Mater. 2021, 7, 929–939. [Google Scholar] [CrossRef]
  72. Zhu, L.Y.; Yuan, K.; Yang, J.G.; Ma, H.P.; Wang, T.; Ji, X.M.; Feng, J.J.; Devi, A.; Lu, H.L. Fabrication of heterostructured p-CuO/n-SnO2 core-shell nanowires for enhanced sensitive and selective formaldehyde detection. Sens. Actuators B Chem. 2019, 290, 233–241. [Google Scholar] [CrossRef]
  73. Cao, Q.; Hao, S.; Wu, Y.; Pei, K.; You, W.; Che, R. Interfacial charge redistribution in interconnected network of Ni2P–Co2P boosting electrocatalytic hydrogen evolution in both acidic and alkaline conditions. Chem. Eng. J. 2021, 424, 130444. [Google Scholar] [CrossRef]
  74. Sayyah, S.M.; Shaban, M.; Rabia, M. M-toluidine polymer film coated platinum electrode as a pH sensor by potentiometric methods. Sens. Lett. 2015, 13, 961–966. [Google Scholar] [CrossRef]
  75. Sayyah, S.M.; Shaban, M.; Rabia, M. A High-sensitivity potentiometric mercuric ion sensor based on m-toluidine films. IEEE Sens. J. 2016, 16, 1541–1548. [Google Scholar] [CrossRef]
  76. Huang, X.; Zhang, M.; Sun, R.; Long, G.; Liu, Y.; Zhao, W. Enhanced hydrogen evolution from CuOx-C/TiO2 with multiple electron transport pathways. PLoS ONE 2019, 14, 0215339. [Google Scholar] [CrossRef]
  77. Li, J.; Jin, X.; Li, R.; Zhao, Y.; Wang, X.; Liu, X.; Jiao, H. Copper oxide nanowires for efficient photoelectrochemical water splitting. Appl. Catal. B Environ. 2019, 240, 1–8. [Google Scholar] [CrossRef]
  78. Baran Aydin, E. Fabrication and characterization of CuO nanostructures: Applications in electrocatalytic hydrogen production. Çukurova Univ. J. Fac. Eng. Archit. 2020, 35, 127–138. [Google Scholar]
  79. Masudy-Panah, S.; Moakhar, R.S.; Chua, C.S.; Tan, H.R.; Wong, T.I.; Chi, D.; Dalapati, G.K. Nanocrystal Engineering of sputter-grown CuO photocathode for visible-light-driven electrochemical water splitting. ACS Appl. Mater. Interfaces 2016, 8, 1206–1213. [Google Scholar] [CrossRef] [PubMed]
  80. Uchiyama, H.; Isobe, K.; Kozuka, H. Preparation of porous CuO films from Cu(NO3)2 aqueous solutions containing poly(vinylpyrrolidone) and their photocathodic properties. RSC Adv. 2017, 7, 18014–18018. [Google Scholar] [CrossRef] [Green Version]
  81. Jin, L.; Al Otaibi, B.; Benetti, D.; Li, S.; Zhao, H.; Mi, Z.; Vomiero, A.; Rosei, F. Near-infrared colloidal quantum dots for efficient and durable photoelectrochemical solar-driven hydrogen production. Adv. Sci. 2016, 3, 1500345. [Google Scholar] [CrossRef] [Green Version]
  82. Ebaid, M.; Kang, J.-H.; Ryu, S.-W. Controlled synthesis of GaN-based nanowires for photoelectrochemical water splitting applications. Semicond. Sci. Technol. 2016, 32, 013001. [Google Scholar] [CrossRef]
  83. Li, Z.; Feng, S.; Liu, S.; Li, X.; Wang, L.; Lu, W. A three-dimensional interconnected hierarchical FeOOH/TiO2/ZnO nanostructural photoanode for enhancing the performance of photoelectrochemical water oxidation. Nanoscale 2015, 7, 19178–19183. [Google Scholar] [CrossRef] [PubMed]
  84. Sherman, B.D.; Ashford, D.L.; Lapides, A.M.; Sheridan, M.V.; Wee, K.-R.; Meyer, T.J. Light-Driven water splitting with a molecular electroassembly-based core/shell photoanode. J. Phys. Chem. Lett. 2015, 6, 3213–3217. [Google Scholar] [CrossRef]
  85. Shaban, M.; Rabia, M.; El-Sayed, A.M.A.; Ahmed, A.; Sayed, S. Photocatalytic properties of PbS/graphene oxide/polyaniline electrode for hydrogen generation. Sci. Rep. 2017, 7, 1–13. [Google Scholar] [CrossRef] [Green Version]
  86. Naldoni, A.; Guler, U.; Wang, Z.; Marelli, M.; Malara, F.; Meng, X.; Besteiro, L.V.; Govorov, A.O.; Kildishev, A.V.; Boltasseva, A.; et al. Broadband hot-electron collection for solar water splitting with plasmonic titanium nitride. Adv. Opt. Mater. 2017, 5, 1601031. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic diagram for photoelectrochemical H2 generation process from sewage water.
Figure 1. Schematic diagram for photoelectrochemical H2 generation process from sewage water.
Materials 15 01489 g001
Figure 2. (a,b) FESEM images, (c) XRD pattern, (d) EDAX, (e) elemental mapping for O (red) and Cu (green), and (f) optical reflectance and bandgap (insert) of the CuO NP thin film. (g,h) The cross section and surface roughness using the modeling program (ImageJ), respectively.
Figure 2. (a,b) FESEM images, (c) XRD pattern, (d) EDAX, (e) elemental mapping for O (red) and Cu (green), and (f) optical reflectance and bandgap (insert) of the CuO NP thin film. (g,h) The cross section and surface roughness using the modeling program (ImageJ), respectively.
Materials 15 01489 g002
Figure 3. The voltage–current relation for (a) Cu/CuO and (b,c) on/off chopping current (d) stability.
Figure 3. The voltage–current relation for (a) Cu/CuO and (b,c) on/off chopping current (d) stability.
Materials 15 01489 g003
Figure 4. (a,b) The effect of light intensity on Jph for the Cu/CuO photoelectrode.
Figure 4. (a,b) The effect of light intensity on Jph for the Cu/CuO photoelectrode.
Materials 15 01489 g004
Figure 5. (a) The potential Jph relation under different wavelengths (390 to 636 nm) illumination and (b) the IPCE for Cu/CuO electrode.
Figure 5. (a) The potential Jph relation under different wavelengths (390 to 636 nm) illumination and (b) the IPCE for Cu/CuO electrode.
Materials 15 01489 g005
Figure 6. (a,b) The temperature effect, (c) activation energy, and (d) heat of reaction for Cu/CuO electrodes.
Figure 6. (a,b) The temperature effect, (c) activation energy, and (d) heat of reaction for Cu/CuO electrodes.
Materials 15 01489 g006
Figure 7. Schematic diagram of the sewage water splitting mechanism.
Figure 7. Schematic diagram of the sewage water splitting mechanism.
Materials 15 01489 g007
Table 1. The sewerage water chemical composition used as the electrolyte for H2 production.
Table 1. The sewerage water chemical composition used as the electrolyte for H2 production.
Material or ElementConcentration (mg/L)
Phenols0.015
F1.0
Al3+3.0
NH35.0
Hg2+0.005
Pb2+0.5
Cd3+0.05
As3+0.05
Cr3+1.0
Cu2+1.5
Ni3+0.1
Fe3+1.5
Mn2+1.0
Zn2+5.0
Ag+0.1
Ba3+2.0
Co2+2.0
Other cations0.1
Pesticides0.2
CN−10.1
Industrial washing0.5
Coli groups4000/100 cm3
Table 2. Comparison between the present study and the previous reported literature; electrolyte used, Jph, and IPCE values.
Table 2. Comparison between the present study and the previous reported literature; electrolyte used, Jph, and IPCE values.
PhotoelectrodeElectrolyteJph (mA/cm2)Applied Voltage (V)IPCE% (390 nm)Light Source
g-C3N4-CuO [23]NaOH0.011.6-300 W xenon lamp
CuO-C/TiO2 [76]glycerol0.001−0.5-300 W xenon lamp
CuO nanowire [77]Na2SO41.5−0.5-simulated AM1.5 illumination
CuO nanostructure [78]KOH1−1.2-White light
CuO thin films [79]Na2SO42.503.1Solar simulator 1.5 global (AM 1.5G)
CuO nanocrystals [80]Na2SO41.1−0.58.7Xenon lamp light
TiO2/CdS/PbS [81]Na2S/Na2S2O320.24AM 1.5G illumination
GaN [82]HBr0.6+18Sunlight
ZnO/TiO2/FeOOH [83]Na2S2O31.590.8-A 150 W xenon lamp
SnO2/TiO2 [84]Na2S2O30.40.6-1 Sun (100 mW cm−2)
Au/PbS/Ro-GO/PANI [85]Na2S2O31.1+110400 W xenon lamp
TiN-TiO2 [86]NaOH3.0 × 10−40.20.03Solar simulator (150 mW cm−2)
BiFeO3 [29]NaOH0.11.60.211 sun (AM 1.5G solar spectrum)
ITO/VO2 [50]Na2S2O31.5 +14400 W metal halid
PrFeO [49]Na2SO40.130−0.6-Simulated sunlight
Poly(3-aminobenzoic acid) frame [53]H2SO41.21.6-150 W xenon lamp
Cu/CuO (Present work)Sewage water4.7+114.6Simulated sunlight
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hadia, N.M.A.; Abdelazeez, A.A.A.; Alzaid, M.; Shaban, M.; Mohamed, S.H.; Hoex, B.; Hajjiah, A.; Rabia, M. Converting Sewage Water into H2 Fuel Gas Using Cu/CuO Nanoporous Photocatalytic Electrodes. Materials 2022, 15, 1489. https://doi.org/10.3390/ma15041489

AMA Style

Hadia NMA, Abdelazeez AAA, Alzaid M, Shaban M, Mohamed SH, Hoex B, Hajjiah A, Rabia M. Converting Sewage Water into H2 Fuel Gas Using Cu/CuO Nanoporous Photocatalytic Electrodes. Materials. 2022; 15(4):1489. https://doi.org/10.3390/ma15041489

Chicago/Turabian Style

Hadia, N. M. A., Ahmed Adel A. Abdelazeez, Meshal Alzaid, Mohamed Shaban, S. H. Mohamed, Bram Hoex, Ali Hajjiah, and Mohamed Rabia. 2022. "Converting Sewage Water into H2 Fuel Gas Using Cu/CuO Nanoporous Photocatalytic Electrodes" Materials 15, no. 4: 1489. https://doi.org/10.3390/ma15041489

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