DOI:
10.1039/D2NJ03116G
(Paper)
New J. Chem., 2022,
46, 15495-15503
A carbon nanowire-promoted Cu2O/TiO2 nanocomposite for enhanced photoelectrochemical performance
Received
23rd June 2022
, Accepted 11th July 2022
First published on 12th July 2022
Abstract
The use of semiconductors as photoelectrochemical electrodes and photocatalysts has been studied extensively over the last few decades; however, it is challenging to design one single material meeting a wide range of requirements for applications such as photostability, wide light absorbance and earth abundance. In this study, we started with earth-abundant materials, namely TiO2, Cu2O and carbon, to design a nanocomposite with TiO2 nanofibers (NFs) and Cu2O nanocubes (NCs) connected by carbon nanowires (NWs). The nanocomposite (C/TiO2/Cu2O) demonstrated excellent photo(electro)activity under simulated visible light as well as enhanced durability. Morphologies imaged by electron microscopes showed that TiO2 NFs and Cu2O NCs were well connected, while thin and long carbon NWs were present throughout the samples and connected the NFs and NCs to form a compact composite. Electrochemical analysis revealed that the band alignment between TiO2 and Cu2O was typical for a type-II heterojunction, which is favourable for the separation of photogenerated charge carriers. Moreover, carbon NWs act as bridges, thus facilitating more efficient charge transfer. Overall, this research showed the design and preparation of a nanostructured composite with improved charge transfer and photostability for solar photoelectrochemical applications.
1. Introduction
Technologies based on photon and semiconductor interactions, e.g. photoelectrochemical and photocatalysis water splitting, have become attractive research areas in recent decades because they can offer a sustainable roadmap towards clean fuels and the environment. An important factor for advancing the research is to obtain catalytic/electrode materials with suitable properties, such as wide light absorbance, photostability, high efficiency, environment friendliness and earth abundance. An extensive range of materials have been studied over the last few decades,1–10 but each has its own limitations.11–15
Titanium dioxide (TiO2) is a well-known stable n-type semiconductor used over the last few decades with a large band gap of 3.2 eV;11 thus, it is mainly active in the ultraviolet (UV) region, which is only 4% of the solar spectrum. Efforts have been devoted to expanding the spectral response of TiO2 in the visible light region using various band engineering methods, such as surface modification, metal doping, creating heterojunction, co-catalyst loading and coupling with narrow band gap semiconductors.16
A highly attractive narrow band gap semiconductor, Cu2O, is found in abundant, low-cost and non-toxic p-type semiconductor with a band gap between 2.0 and 2.2 eV.17 However, Cu2O is prone to photocorrosion, which is attributed to the oxidation and reduction potential of Cu2O lying within the band gap.18
TiO2–Cu2O as a binary heterojunction/composite has demonstrated synergistic photoelectrochemical and photocatalytic performance based on efficiency and stability.19 However, there are inevitable defects or poor connections at the interface between the two different components, which act as charge carrier sinks and hinder efficient charge carrier migration. A strategy to improve the interface and provide better connections between the two different components is highly desirable to fully unleash the synergy between TiO2 and Cu2O, thus providing an ideal catalyst for photoelectrochemical and photocatalytic applications.
There are attempts to utilise carbon nanomaterials in composite photoelectrochemical electrode materials and photocatalysts as charge transfer promoters and structure stabilisers. For example, carbon nanowires (NWs) are known to have excellent mechanical and electron transport properties.20
In addition to the physical and chemical properties of TiO2 and Cu2O, their structural properties (e.g. morphology, surface area and crystallinity) also play vital roles in their photoelectrochemical and photocatalytic properties. In our previous study, it is already revealed that high crystalline TiO2 nanofibers (NFs) and Cu2O nanocubes (NCs) can offer optimised charge transfer efficiency and photostability compared to other structures.19,21–23
In this study, we designed a nanocomposite with TiO2 NFs and Cu2O NCs, which was further connected and enhanced by carbon NWs. The synthesis procedure was simple and required inexpensive solvents and operating conditions. The photoelectrochemical (PEC) performance of the ternary C-NW/TiO2 NF/Cu2O composite showed improved charge carrier separation, superior PEC activity and enhanced stability.
2. Experimental
2.1. Synthesis of C-NW/TiO2 NF
A single crystalline 1D TiO2 NF was fabricated as described in our previous study via topotactic transformation by ion-exchange and dehydration.22 Subsequent C-NW/TiO2 NF was prepared by a hydrothermal synthesis using gelatine as the carbon precursor as it is a cheap polymer obtained from animal skin, bones and tissues.24 It is known to have strong coating interactions with metal nanoparticles.25 Gelatine was dissolved in a mixture of TiO2 NFs in 20 ml water and stirred at 40 °C. The sample was placed in a stainless-steel autoclave and heated at 250 °C for 4 h. After the hydrothermal treatment, the sample was centrifugated at 7000 rpm for 10 min with several washes of water and dried in the oven. Gelatine is abundant in amino and carboxyl groups, so it can be hydrothermally treated with only pure water. Hence, it is an inexpensive water-soluble polymer. An advantage of this procedure is that no strong acid or surface passivation reagent or post-treatment is needed to fabricate C-NW/TiO2 NFs.26
2.2. Synthesis of C-NW/TiO2 NF/Cu2O
C-NW/TiO2 NF/Cu2O was synthesised by solution phase chemistry under ambient conditions, as described previously.16 The TiO2 to Cu2O ratio was targeted at 1
:
1 by controlling the dosages of corresponding precursors in the synthesis process. In a typical process, PEG was ultrasonicated in 100 ml of water for 30 min. C-NW/TiO2 NF was mixed with 0.2 M CuCl2 and 0.01 M PEG-600 before ultrasonication for 10 min. This mixture was then heated to 50 °C, forming a blue solution. 2 M NaOH was poured into the blue solution and 0.2 M N2H4·H2O was added dropwise in the stirred mixture to generate Cu(I) ions and further heated for 10 min. The resulting mixture was purged in N2 to form C-NW/TiO2 NF/Cu2O, followed by centrifugation at 7000 rpm with several washes with H2O and ethanol to remove any residual PEG.
2.3. Characterisation methods
TEM images were taken at 200 kV on a JEM-2100Plus microscope (Warwick University, UK). Samples were prepared by ultrasonication and drop-casted on a Cu grid coated with a carbon film. Powder X-ray diffraction (XRD) was operated on a Bruker-AXS D8 ADVANCE diffractometer at 40 kV, 40 mA and Cu Kα radiation (λ = 0.15418 nm) between 2θ 10 and 80° in 0.02° steps. Brunauer–Emmett–Teller (BET) surface areas were calculated at 77 K by N2 physisorption using a Quantachrome NOVA 4000e porosimeter on degassed samples at 120 °C for 4 h. Diffuse reflectance UV-Vis Spectra (DRUVS) were recorded on a Thermo Scientific Evo220 Spectrometer using KBr as a standard in an absorbance range between 200 and 700 nm. The ratio of TiO2 to Cu2O in TiO2 NF/Cu2O and C-NW/TiO2 NF/Cu2O samples was quantified by ICP-OES (Optima2000DV, USA).
2.4. Photoelectrochemical methods
A homogenous colloid was formed within 30 min of sonication of the as-prepared samples (5 mg), and Nafion (10 μL, 5 wt%) was dispersed in a water/ethanol mixture (1 ml, 3
:
1 v/v). 5 μL colloid was deposited on a glassy carbon electrode (3 mm dia.) used as the working electrode. A Pt wire was used as the counter electrode, and Hg/Hg2SO4 was used as the reference electrode in a three-electrode photoelectrochemical cell. The electrolyte (0.5 M Na2SO4) was purged in N2 for 30 min. A 200 W Hg–Xe arc lamp (Oriel Instruments 66002, with UV cutoff filter, λ > 420 nm) was used as the light source. The light intensity was adjusted and kept at 100 mW cm−2 for all experiments. Measurements were performed using Autolab potentiostat with Nova software. Mott–Schottky plots were recorded under DC polarisation at a potential range of −1 to 0 V with a potential step of 10 mV at a frequency of 1000 Hz. Nyquist plots were recorded under an AC signal of 10 mV at a frequency range of 100 kHz–0.1 Hz.
3. Results and discussion
TEM and SEM images in Fig. 1 show the morphology and size of the individual components and C-NW/TiO2 NF/Cu2O as a nanocomposite. Fig. 1a demonstrates the 1D TiO2 NF with lengths between 500 nm and 5 μm and diameters between 100 and 200 nm. Fig. 1b illustrates the Cu2O nanocubes with sizes ranging between 150 and 250 nm. After the hydrothermal synthesis, amino acids were carbonised to form carbon NWs, and thus the ternary Cu2O, C-NW and TiO2 nanocomposite was formed, as shown in Fig. 1c and d. In these two images, cubic Cu2O, TiO2 NF and C-NW were all clearly visible. TiO2 NFs were intimately in contact with Cu2O nanocubes, with C-NWs occupying the spaces and further promoting the Cu2O–TiO2 connections.
 |
| Fig. 1 TEM images of (a) TiO2 NF, (b) Cu2O nanocubes, (c) C-NW/TiO2 NF/Cu2O, and (d) SEM images of C-NW/TiO2 NF/Cu2O. | |
XRD patterns confirmed the characteristics of Cu2O and TiO2 peaks in Fig. 2a, indicating that the heterojunction nanocomposite consists of Cu2O and TiO2, which is in good agreement with the TEM and SEM images. The diffraction peaks for TiO2 NF were well defined and corresponded to the crystal planes of (101), (004), (200), (105), (211) and (204) at 2θ = 25.3°, 37.8°, 48.0°, 53.8°, 55.4° and 62.5°, respectively, which confirmed anatase TiO2 NF (tetragonal, JCPDS 21-1272).27 The XRD patterns illustrated the formation of crystalline Cu2O by 2θ peaks at 29.6°, 36.4°, 42.3°, 61.4°, 73.5° and 77.4°, which are associated with characteristic (110), (111), (200), (220), (311), and (222) peaks of Cu2O, respectively (cubic, JCPDS 73-0687).28 The ratio of TiO2 to Cu2O in TiO2 NF/Cu2O and C-NW/TiO2 NF/Cu2O samples was quantified as 0.9
:
1 by ICP-OES. There were weak traces of CuO in both Cu2O and TiO2 NF/Cu2O, which is in agreement with JCPDS 45-093729 but not observed in the C-NW/TiO2 NF/Cu2O sample. There is a weak but a broad pattern at 25°, suggesting either a low concentration of C-NWs or the amorphous carbon.
 |
| Fig. 2 (a) XRD patterns of (i) Cu2O, (ii) TiO2 NF, (iii) TiO2 NF/Cu2O and (iv) C-NW/TiO2 NF/Cu2O. (b) TGA and (c) Raman spectrum of C-NW/TiO2 NF/Cu2O. (d) N2 adsorption–desorption isotherms of TiO2 NF, Cu2O nanocubes, TiO2 NF/Cu2O, C-NW/TiO2 NF/Cu2O. | |
The carbon content in C-NW/TiO2 NF/Cu2O was investigated by thermogravimetric analysis (TGA), as shown in Fig. 2b. There are three significant weight losses at 100 °C, 200 °C and 400 °C, respectively. The initial loss for the samples at 100 °C was due to the adsorbed moisture (2.1%), while the 2.9% weight loss until 200 °C was attributed to the removal of residual PEG, which was the structuring agent used to fabricate Cu2O nanocubes. The weight loss above 200 °C was 4.3%, which accounted for the total loss of C-NWs in the heterojunction. Different C-NW contents (i.e., in the range from 2% to 10%) were prepared. Within this range, there was no clear difference in various photoelectrochemical performances. This 4.3% sample was presented as a representative. Fig. 2c illustrates the nature of carbon in C-NW/TiO2 NF/Cu2O using Raman spectroscopy. The Raman spectrum shows two peaks of 1370 cm−1 and 1593 cm−1 for the D and G bands, respectively, which is characteristic of amorphous carbon.
Surface areas varied between the different samples (Fig. 2d), with Cu2O nanocubes having the lowest BET surface area of 22.9 m2 g−1. TiO2 NF exhibited the highest BET surface area of 93.5 m2 g−1. Binary TiO2 NF/Cu2O had a surface area of 53.8 m2 g−1 reasonably between the two individual components. C-NW/TiO2 NF/Cu2O had a higher BET surface area than TiO2 NF/Cu2O, which was due to the integration of C-NWs as it has a high surface area. This may have further implications for photoelectrochemical performance because C-NWs may act as an electron anchor by improving electron mobility and transportation in the heterojunction.30,31
The optical properties of different samples were analysed and presented in Fig. 3, whereby the absorbance spectra ranged from 200 to 600 nm (Fig. 3a), consistent with previously reported.32
 |
| Fig. 3 (a) DRUVS and (b) Tauc plots of TiO2 NF, Cu2O, TiO2 NF/Cu2O and C-NW/TiO2 NF/Cu2O. | |
The optical band gaps EBG were calculated from the Tauc plots (Fig. 3b) using eqn (1).
where
C is the proportionality constant and ‘
α’ is the absorption coefficient determined from the Kubelka–Munk formula (
eqn (2)).
|  | (2) |
The direct band gaps of TiO
2 NF, Cu
2O, TiO
2 NF/Cu
2O and C-NW/TiO
2 NF/Cu
2O were 3.28, 2.36, 2.38 and 2.21 eV, respectively, as presented in
Table 1. The tested samples containing Cu
2O absorbed light in the visible region at the wavelength of 650 nm. Two absorbance peaks are present at 300 nm and 480 nm, confirming both components were present in the system, which is in agreement with a previous report.
33 C-NW/TiO
2 NF/Cu
2O showed a red shift to TiO
2 NF/Cu
2O, which indicates C-NWs have strong visible light absorption properties. It may also promote interfacial contact between C-NWs and the two semiconductors.
Table 1 Photophysical properties of TiO2 NF, Cu2O, TiO2 NF/Cu2O and C-NW/TiO2 NF/Cu2O
Sample |
BET surface areaa (m2 g−1) |
Band gapb (eV) |
N2 porosimetry.
DRUVS.
|
TiO2 NF |
93.5 |
3.28 |
Cu2O |
22.9 |
2.36 |
TiO2 NF/Cu2O |
53.8 |
2.38 |
C-NW/TiO2 NF/Cu2O |
66.9 |
2.21 |
The effect of C-NW enhancement on the interfacial charge transfer between the TiO2 NF–Cu2O heterojunction was investigated using PEC measurements under chopped visible light illumination at 0 V vs. RHE. PEC measurements were studied in a three-electrode system, with Pt wire as the counter electrode, Hg/Hg2SO4 as the reference electrode, and different samples were drop-casted onto a glassy carbon electrode as the working electrode. Fig. 4a shows C-NW/TiO2 NF/Cu2O > TiO2 NF/Cu2O > Cu2O > TiO2 NF in terms of photoresponse. In order to enable comparison with other reported and relevant TiO2-based materials, the widely used benchmark P25 TiO2 was also tested, showing photoresponse similar to TiO2 NF (data did not show as it overlapped with that of TiO2 NF), which was consistent with our previous study.19 The low photocurrent density of TiO2 NF was attributed to poor visible light response, and Cu2O had photostability issues during charge transfer. The photocurrent density of C-NW/TiO2 NF/Cu2O was approximately 6 times higher than TiO2 NF/Cu2O and above the factor of 8 times higher than both Cu2O and TiO2 NF. The C-NW promoted the heterojunction nanocomposite produced a higher photocurrent response, indicating more efficient charge transfer, where C-NWs can act as a carrier mobility layer at the TiO2 NF/Cu2O interface.34 Cu2O was difficult to use in practical applications as it has low photostability; therefore, the impact of C-NWs on catalyst stability (Fig. 4b) was studied. The photocurrent generated by Cu2O decreased at a faster rate than C-NW/TiO2/Cu2O over a time period of 24 hours, which suggests that the addition of TiO2 NF and C-NWs increased photostability. The advantage of C-NW as a conductive network was more profound when we consider that the samples were dropcasted on a glassy carbon electrode. The sample particles formed a dense layer with the C-NW conductive network embedded, which could not only promote the charge mobility but also enhance the Cu2O stability.
 |
| Fig. 4 (a) Transient photocurrent and (b) photostability (0 V vs. RHE) of TiO2 NF, Cu2O, TiO2 NF/Cu2O and C-NW/TiO2 NF/Cu2O under visible light irradiation. (c) I–V curve and (d) chopped I–V curves of TiO2 NF, Cu2O, TiO2 NF/Cu2O and C-NW/TiO2 NF/Cu2O under visible light irradiation at a scan rate of 10 mV s−1, (reaction conditions: 200 W Hg–Xe arc lamp, 0.5 M Na2SO4 electrolyte). | |
Bare TiO2 NF showed a low photocurrent density vs. voltage curves (Fig. 4c), but Cu2O nanocubes exhibited a higher photocurrent density under visible light due to its narrow band gap. With the introduction of the heterojunction nanocomposite, TiO2 NF/Cu2O, the photocurrent density increased to 60.3 μA cm−2 at 1.23 V vs. RHE. The integration of C-NWs yielded a significant increase in the photocurrent density of 139.8 μA cm−2, which was almost 4 times higher than TiO2 NF. The sharp increase in the photocurrent density indicates high charge transfer and separation efficiency with C-NWs in C-NW/TiO2 NF/Cu2O.35 The results were similar in the chopped photocurrent density vs. voltage curve, as shown in Fig. 4d, where C-NW/TiO2 NF/Cu2O reached photocurrent densities of 122.4 μA cm−2 at −0.3 V vs. RHE, which was 2 times higher than TiO2 NF/Cu2O.
The interfacial properties of C-NW/TiO2 NF/Cu2O were further studied by electrochemical impedance spectroscopy (EIS) in the form of a Nyquist plot. A Nyquist plot was used to investigate the characteristics of the charge transfer process, whereby the semicircle diameter indicates the charge transfer resistance. In Fig. 5a, the semicircle diameter of Cu2O and TiO2 NF is slightly larger than that of TiO2 NF/Cu2O, which shows they have similar charge transfer resistance. C-NW/TiO2 NF/Cu2O exhibited the lowest charge transfer resistance, which is attributed to the integration of C-NWs. This indicates that C-NWs reduced the resistance to charge carrier transport due to the efficient separation of charge carriers in the nanocomposite.35Fig. 5b and c show the Mott–Schottky plots of TiO2 NF, C-NW/TiO2 NF and Cu2O, which were utilised to calculate the flat band potential and majority carrier densities of each semiconductor in the heterojunction. The positive and negative slopes are indicative of n-type and p-type semiconductors, respectively. This type of p–n heterojunction is a favourable structure for enhanced charge separation.36 The flat band potentials are shown by the x-intercept of the linear section, which are −0.76 V, −0.72 V and 0.54 V, vs. RHE for TiO2 NF, C-NW/TiO2 NF and Cu2O, respectively. The anodic shift of 40 mV between TiO2 NF and C-NW/TiO2 NF is in agreement with previous reports.37
 |
| Fig. 5 (a) EIS (Nyquist plot at 0 V vs. RHE) of TiO2 NF, Cu2O, TiO2 NF/Cu2O and C-NW/TiO2 NF/Cu2O under visible light irradiation; Mott–Schottky plot of TiO2 NF (b) and Cu2O (c); (200 W Hg-Xe arc lamp, 0.5 M Na2SO4 electrolyte); (d) Schematic for the energy band structure of C-NWs/TiO2/Cu2O nanocomposite. | |
The gradient of the linear portion of the plot was used to calculate the majority carrier density (NA) calculated from eqn (3).
|  | (3) |
where
C is the capacitance,
ε is the dielectric constant of TiO
2 and Cu
2O, which is 170
38 and 7.60,
39 respectively,
ε0 is the permittivity of free space,
V. is the applied potential,
Efb is the flat band potential,
e is the electron charge,
kB is the Boltzmann's constant and
T is the temperature. The
NA values of Cu
2O, TiO
2 NF and TiO
2 NF/C-NWs were 5.36 × 10
18 cm
−3, 1.06 × 10
19 cm
−3 and 1.30 × 10
19 cm
−3, respectively. This showed that 1D TiO
2 NF had more charge carrier activity, but more studies are needed to gain conclusive results.
The valence band edge of Cu2O, EV, was calculated from eqn (4):
|  | (4) |
where the effective density of states,

. The effective hole mass is denoted as
mh = 0.58
me for Cu
2O,
40 where
me is the mass of a free electron.
NV for Cu
2O was 1.10 × 10
19 cm
−3. The valence band edge (
EV) was 0.56 V for Cu
2O. Thus, as Cu
2O has a band gap of 2.36, the valence and conduction band positions were at 0.56 V and −1.80 V
vs. RHE, respectively. Assuming the distance between the flat band potential of TiO
2 NF and the lowest potential of the conduction band (CB) is negligible,
40 the CB of TiO
2 NF was −0.76 V
vs. RHE. Thus, the band gap of TiO
2 NF was 3.28, so the valence band and conduction band were positioned at 2.52 V and −0.76 V
vs. RHE, respectively.
A schematic of the energy band structure is proposed in Fig. 5d. The enhanced photoelectrochemical performance of C-NW/TiO2 NF/Cu2O nanocomposite was attributed to a few reasons. As Cu2O has a narrow bandgap energy of 2.36 eV, it can be excited by visible light. The conduction band of TiO2 lies at a more positive potential than that of Cu2O, and the electrons from Cu2O can inject into the conduction band of TiO2. This prevents the recombination of photogenerated electrons and holes in Cu2O.41 Accumulated electrons in the conduction band of TiO2 can then transfer to the glassy carbon electrode.42 The addition of C-NWs further broadens the solar absorption to the visible light region. Also, the impedance spectra show that the C-NW/TiO2 NF/Cu2O nanocomposite has improved charge carrier density and least charge transfer resistance compared to bare TiO2 NF and Cu2O. This suggests that the efficient charge transfer was further assisted by the increased conductivity with C-NWs. The addition of C-NWs improved the charge carrier density; so it is proposed that the C-NWs promote electron transfer between CB of Cu2O and CB of TiO2, enhancing the separation of photogenerated electron-hole pairs. Overall, the introduction of C-NWs in the nanocomposite of C-NWs/TiO2/Cu2O enhances the PEC performance, increases charge transport, and the nanocomposite increases photostability and extends the light absorption region.
4. Conclusions
In summary, we fabricated a C-NW/TiO2 NF/Cu2O heterojunction nanocomposite via a hydrothermal treatment and subsequent solution phase method, which provided an intimate contact between 1D TiO2 NF and Cu2O nanocubes surrounded by C-NWs. The carbon content in the prepared samples was confirmed by both TGA and Raman spectroscopy, and DRUVS indicated that C-NWs have strong absorption properties as there was superior absorption at higher wavelengths. C-NW/TiO2 NF/Cu2O recorded enhanced PEC activity with higher photocurrent density, and impedance spectra illustrated less charge transfer resistance than bare TiO2 NF and Cu2O. After calculating the charge carrier densities of TiO2 NF and C-NW/TiO2 NF, it was shown that the addition of C-NWs improved the charge carrier density by promoting electron transfer. Therefore, C-NW/TiO2 NF/Cu2O provided superior light absorption properties, efficient charge transfer, separation of electron–hole pairs and the integration of C-NWs, which promote further electron transport within the heterojunction nanocomposite. All these properties contribute towards fabricating a nanocomposite with superior PEC activity. This work is an example of designing and fabricating effective heterojunction nanocomposite materials for solar water splitting.
Conflicts of interest
There are no conflicts of interest to declare.
Acknowledgements
This research was supported by the European Commission H2020 Marie S Curie Research and Innovation Staff Exchange (RISE) award (Grant No. 871998).
References
- S. Li, C. Wang, M. Cai, F. Yang, Y. Liu, J. Chen, P. Zhang, X. Li and X. Chen, Facile fabrication of TaON/Bi2MoO6 core–shell S-scheme heterojunction nanofibers for boosting visible-light catalytic levofloxacin degradation and Cr(VI) reduction, Chem. Eng. J., 2022, 428, 131158, DOI:10.1016/j.cej.2021.131158.
- S. Li, J. Chen, S. Hu, H. Wang, W. Jiang and X. Chen, Facile construction of novel Bi2WO6/Ta3N5 Z-scheme heterojunction nanofibers for efficient degradation of harmful pharmaceutical pollutants, Chem. Eng. J., 2020, 402, 126165, DOI:10.1016/j.cej.2020.126165.
- C. Wang, S. Li, M. Cai, R. Yan, K. Dong, J. Zhang and Y. Liu, Rationally designed tetra (4-carboxyphenyl) porphyrin/graphene quantum dots/bismuth molybdate Z-scheme heterojunction for tetracycline degradation and Cr(VI) reduction: Performance, mechanism, intermediate toxicity appraisement, J. Colloid Interface Sci., 2022, 619, 307–321, DOI:10.1016/j.jcis.2022.03.075.
- S. Li, C. Wang, Y. Liu, M. Cai, Y. Wang, H. Zhang, Y. Guo, W. Zhao, Z. Wang and X. Chen, Photocatalytic degradation of tetracycline antibiotic by a novel Bi2Sn2O7/Bi2MoO6 S-scheme heterojunction: Performance, mechanism insight and toxicity assessment, Chem. Eng. J., 2022, 429, 132519, DOI:10.1016/j.cej.2021.132519.
- S. Li, C. Wang, M. Cai, Y. Liu, K. Dong and J. Zhang, Designing oxygen vacancy mediated bismuth molybdate (Bi2MoO6)/N-rich carbon nitride (C3N5) S-scheme heterojunctions for boosted photocatalytic removal of tetracycline antibiotic and Cr(VI): Intermediate toxicity and mechanism insight, J. Colloid Interface Sci., 2022, 624, 219–232, DOI:10.1016/j.jcis.2022.05.151.
- S. Li, M. Cai, Y. Liu, J. Zhang, C. Wang, S. Zang, Y. Li, P. Zhang and X. Li, In situ construction of a C3N5 nanosheet/Bi2WO6 nanodot S-scheme heterojunction with enhanced structural defects for the efficient photocatalytic removal of tetracycline and Cr(VI), Inorg. Chem. Front., 2022, 9, 2479–2497, 10.1039/D2QI00317A.
- C. Wang, M. Cai, Y. Liu, F. Yang, H. Zhang, J. Liu and S. Li, Facile construction of novel organic–inorganic tetra (4-carboxyphenyl) porphyrin/Bi2MoO6 heterojunction for tetracycline degradation: Performance, degradation pathways, intermediate toxicity analysis and mechanism insight, J. Colloid Interface Sci., 2022, 605, 727–740, DOI:10.1016/j.jcis.2021.07.137.
- J. Luo, P. Lin, P. Zheng, X. Zhou, X. Ning, L. Zhan, Z. Wu, X. Liu and X. Zhou, In suit constructing S-scheme FeOOH/MgIn2S4 heterojunction with boosted interfacial charge separation and redox activity for efficiently eliminating antibiotic pollutant, Chemosphere, 2022, 298, 134297, DOI:10.1016/j.chemosphere.2022.134297.
- J. Luo, X. Zhou, F. Yang, X. Ning, L. Zhan, Z. Wu and X. Zhou, Generating a captivating S-scheme CuBi2O4/CoV2O6 heterojunction with boosted charge spatial separation for efficiently removing tetracycline antibiotic from wastewater, J. Clean. Prod., 2022, 357, 131992, DOI:10.1016/j.jclepro.2022.131992.
- J. Luo, J. Chen, X. Chen, X. Ning, L. Zhan and X. Zhou, Construction of cerium oxide nanoparticles immobilized on the surface of zinc vanadate nanoflowers for accelerated photocatalytic degradation of tetracycline under visible light irradiation, J. Colloid Interface Sci., 2021, 587, 831–844, DOI:10.1016/j.jcis.2020.11.044.
- A. L. Linsebigler, G. Lu and J. T. Yates, Photocatalysis on TiO2 Surfaces: Principles, Mechanisms, and Selected Results, Chem. Rev., 1995, 95, 735–758, DOI:10.1021/cr00035a013.
- Ü. Özgür, Y. I. Alivov, C. Liu, A. Teke, M. A. Reshchikov, S. Doğan, V. Avrutin, S.-J. Cho and H. Morkoç, A comprehensive review of ZnO materials and devices, J. Appl. Phys., 2005, 98, 041301, DOI:10.1063/1.1992666.
- K. Sivula, F. Le Formal and M. Grätzel, Solar Water Splitting: Progress Using Hematite (α-Fe2O3) Photoelectrodes, ChemSusChem, 2011, 4, 432–449, DOI:10.1002/cssc.201000416.
- A. Kudo, K. Omori and H. Kato, A Novel Aqueous Process for Preparation of Crystal Form-Controlled and Highly Crystalline BiVO4 Powder from Layered Vanadates at Room Temperature and Its Photocatalytic and Photophysical Properties, J. Am. Chem. Soc., 1999, 121, 11459–11467, DOI:10.1021/ja992541y.
- M. Ni, M. K. H. Leung, D. Y. C. Leung and K. Sumathy, A review and recent developments in photocatalytic water-splitting using TiO2 for hydrogen production, Renew. Sustain. Energy Rev., 2007, 11, 401–425, DOI:10.1016/j.rser.2005.01.009.
- R. Abe, Recent progress on photocatalytic and photoelectrochemical water splitting under visible light irradiation, J. Photochem. Photobiol., C, 2010, 11, 179–209, DOI:10.1016/j.jphotochemrev.2011.02.003.
- M. Hara, T. Kondo, M. Komoda, S. Ikeda, J. N. Kondo, K. Domen, M. Hara, K. Shinohara and A. Tanaka, Cu2O as a photocatalyst for overall water splitting under visible light irradiation, Chem. Commun., 1998, 357–358, 10.1039/a707440i.
- S. J. A. Moniz, S. A. Shevlin, D. J. Martin, Z.-X. Guo and J. Tang, Visible-light driven heterojunction photocatalysts for water splitting – a critical review, Energy Environ. Sci., 2015, 8, 731–759, 10.1039/C4EE03271C.
- K. Sekar, C. Chuaicham, B. Vellaichamy, W. Li, W. Zhuang, X. Lu, B. Ohtani and K. Sasaki, Cubic Cu2O nanoparticles decorated on TiO2 nanofiber heterostructure as an excellent synergistic photocatalyst for H2 production and sulfamethoxazole degradation, Appl. Catal. B Environ., 2021, 294, 120221, DOI:10.1016/j.apcatb.2021.120221.
- K. Woan, G. Pyrgiotakis and W. Sigmund, Photocatalytic Carbon-Nanotube-TiO2 Composites, Adv. Mater., 2009, 21, 2233–2239, DOI:10.1002/adma.200802738.
- W. Li, Y. Bai, C. Liu, Z. Yang, X. Feng, X. Lu, N. K. van der Laak and K.-Y. Chan, Highly Thermal Stable and Highly Crystalline Anatase TiO2 for Photocatalysis, Environ. Sci. Technol., 2009, 43, 5423–5428, DOI:10.1021/es8037005.
- W. Li, Y. Bai, W. Liu, C. Liu, Z. Yang, X. Feng, X. Lu and K.-Y. Chan, Single-crystalline and reactive facets exposed anatase TiO2 nanofibers with enhanced photocatalytic properties, J. Mater. Chem., 2011, 21, 6718, 10.1039/c1jm10115c.
- S. Karthikeyan, S. Kumar, L. J. Durndell, M. A. Isaacs, C. M. A. Parlett, B. Coulson, R. E. Douthwaite, Z. Jiang, K. Wilson and A. F. Lee, Size-Dependent Visible Light Photocatalytic Performance of Cu2O Nanocubes, ChemCatChem, 2018, 10, 3554–3563, DOI:10.1002/cctc.201800439.
- M. Amjad, M. Iqbal, A. Faisal, A. M. Junjua, I. Hussain, S. Z. Hussain, H. A. Ghramh, K. A. Khan and H. A. Janjua, Hydrothermal synthesis of carbon nanodots from bovine gelatin and PHM3 microalgae strain for anticancer and bioimaging applications, Nanoscale Adv., 2019, 1, 2924–2936, 10.1039/C9NA00164F.
- Q. Cui, F. He, X. Wang, B. Xia and L. Li, Gold Nanoflower@Gelatin Core–Shell Nanoparticles Loaded with Conjugated Polymer Applied for Cellular Imaging, ACS Appl. Mater. Interfaces, 2013, 5, 213–219, DOI:10.1021/am302589g.
- Q. Liang, W. Ma, Y. Shi, Z. Li and X. Yang, Easy synthesis of highly fluorescent carbon quantum dots from gelatin and their luminescent properties and applications, Carbon N. Y, 2013, 60, 421–428, DOI:10.1016/j.carbon.2013.04.055.
- H. G. Yang, G. Liu, S. Z. Qiao, C. H. Sun, Y. G. Jin, S. C. Smith, J. Zou, H. M. Cheng and G. Q. Lu, Solvothermal Synthesis and Photoreactivity of Anatase TiO2 Nanosheets with Dominant {001} Facets, J. Am. Chem. Soc., 2009, 131, 4078–4083, DOI:10.1021/ja808790p.
- S. Sun, X. Zhang, X. Song, S. Liang, L. Wang and Z. Yang, Bottom-up assembly of hierarchical Cu2O nanospheres: controllable synthesis, formation mechanism and enhanced photochemical activities, CrystEngComm, 2012, 14, 3545, 10.1039/c2ce25071c.
- S. Zhang, H. Liu, C. Sun, P. Liu, L. Li, Z. Yang, X. Feng, F. Huo and X. Lu, CuO/Cu2O porous composites: shape and composition controllable fabrication inherited from metal organic frameworks and further application in CO oxidation, J. Mater. Chem. A., 2015, 3, 5294–5298, 10.1039/C5TA00249D.
-
M. Hiramatsu and M. Hori, Carbon Nanowalls: Synthesis and Emerging Applications, Springer Vienna, Vienna, 2010 DOI:10.1007/978-3-211-99718-5.
- F. Meng, J. Li, S. K. Cushing, J. Bright, M. Zhi, J. D. Rowley, Z. Hong, A. Manivannan, A. D. Bristow and N. Wu, Photocatalytic Water Oxidation by Hematite/Reduced Graphene Oxide Composites, ACS Catal., 2013, 3, 746–751, DOI:10.1021/cs300740e.
- J. Zhang, W. Liu, X. Wang, X. Wang, B. Hu and H. Liu, Enhanced decoloration activity by Cu2O@TiO2 nanobelts heterostructures via a strong adsorption-weak photodegradation process, Appl. Surf. Sci., 2013, 282, 84–91, DOI:10.1016/j.apsusc.2013.05.054.
- Y. Bessekhouad, D. Robert and J.-V. Weber, Photocatalytic activity of Cu2O/TiO2, Bi2O3/TiO2 and ZnMn2O4/TiO2 heterojunctions, Catal. Today, 2005, 101, 315–321, DOI:10.1016/j.cattod.2005.03.038.
- S. G. Babu, R. Vinoth, D. Praveen Kumar, M. V. Shankar, H.-L. Chou, K. Vinodgopal and B. Neppolian, Influence of electron storing, transferring and shuttling assets of reduced graphene oxide at the interfacial copper doped TiO2 p–n heterojunction for increased hydrogen production, Nanoscale, 2015, 7, 7849–7857, 10.1039/C5NR00504C.
- J. Hou, C. Yang, H. Cheng, S. Jiao, O. Takeda and H. Zhu, High-performance p-Cu2O/n-TaON heterojunction nanorod photoanodes passivated with an ultrathin carbon sheath for photoelectrochemical water splitting, Energy Environ. Sci., 2014, 7, 3758–3768, 10.1039/C4EE02403F.
- S. Li, C. Wang, Y. Liu, B. Xue, W. Jiang, Y. Liu, L. Mo and X. Chen, Photocatalytic degradation of antibiotics using a novel Ag/Ag2S/Bi2MoO6 plasmonic p–n heterojunction photocatalyst: Mineralization activity, degradation pathways and boosted charge separation mechanism, Chem. Eng. J., 2021, 415, 128991, DOI:10.1016/j.cej.2021.128991.
- S. Sakthivel and H. Kisch, Daylight Photocatalysis by Carbon-Modified Titanium Dioxide, Angew. Chem., Int. Ed., 2003, 42, 4908–4911, DOI:10.1002/anie.200351577.
- R. A. Parker, Static Dielectric Constant of Rutile (TiO2), 1.6—106 K, Phys. Rev., 1961, 124, 1719–1722, DOI:10.1103/PhysRev.124.1719.
- Z. Zhang and P. Wang, Highly stable copper oxide composite as an effective photocathode for water splitting via a facile electrochemical synthesis strategy, J. Mater. Chem., 2012, 22, 2456–2464, 10.1039/C1JM14478B.
- A. Paracchino, J. C. Brauer, J.-E. Moser, E. Thimsen and M. Graetzel, Synthesis and Characterization of High-Photoactivity Electrodeposited Cu2O Solar Absorber by Photoelectrochemistry and Ultrafast Spectroscopy, J. Phys. Chem. C., 2012, 116, 7341–7350, DOI:10.1021/jp301176y.
- M. Wang, L. Sun, Z. Lin, J. Cai, K. Xie and C. Lin, p–n Heterojunction photoelectrodes composed of Cu2O-loaded TiO2 nanotube arrays with enhanced photoelectrochemical and photoelectrocatalytic activities, Energy Environ. Sci., 2013, 6, 1211, 10.1039/c3ee24162a.
- Y. Hou, X. Y. Li, Q. D. Zhao, X. Quan and G. H. Chen, Fabrication of Cu2O/TiO2 nanotube heterojunction arrays and investigation of its photoelectrochemical behavior, Appl. Phys. Lett., 2009, 95, 093108, DOI:10.1063/1.3224181.
|
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2022 |
Click here to see how this site uses Cookies. View our privacy policy here.