Next Article in Journal
Catalytic Behavior of Au Confined in Ionic Liquid Film: A Kinetics Study for the Hydrochlorination of Acetylene
Next Article in Special Issue
Recent Advancements in Photocatalysis Coupling by External Physical Fields
Previous Article in Journal
Trimeric Ruthenium Cluster-Derived Ru Nanoparticles Dispersed in MIL-101(Cr) for Catalytic Transfer Hydrogenation
Previous Article in Special Issue
Microwave-Assisted Photocatalytic Degradation of Organic Pollutants via CNTs/TiO2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photocatalytic CO2 Conversion Using Anodic TiO2 Nanotube-CuxO Composites

by
Timofey P. Savchuk
1,2,3,*,
Ekaterina V. Kytina
1,
Elizaveta A. Konstantinova
1,*,
Vladimir G. Kytin
1,
Olga Pinchuk
2,
Andrey K. Tarhanov
2,
Vladimir B. Zaitsev
1 and
Tomasz Maniecki
3
1
Physics Department M.V., Lomonosov Moscow State University, Leninskie Gory 1-2, 119991 Moscow, Russia
2
Institute of Advanced Materials and Technologies, National Research University of Electronic Technology—MIET, Bld. 1, Shokin Square, Zelenograd, 124498 Moscow, Russia
3
Institute of General and Ecological Chemistry, Lodz University of Technology, Zeromskiego 116, 90-924 Lodz, Poland
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(9), 1011; https://doi.org/10.3390/catal12091011
Submission received: 2 August 2022 / Revised: 27 August 2022 / Accepted: 30 August 2022 / Published: 7 September 2022
(This article belongs to the Special Issue 10th Anniversary of Catalysts—Feature Papers in Photocatalysis)

Abstract

:
Nanosized titanium dioxide (TiO2) is currently being actively studied by the global scientific community, since it has a number of properties that are important from a practical point of view. One of these properties is a large specific surface, which makes this material promising for use in photocatalysts, sensors, solar cells, etc. In this work, we prepared photocatalysts based on TiO2 nanotubes for converting carbon dioxide (CO2) into energy-intensive hydrocarbon compounds. Efficient gas-phase CO2 conversion in the prepared single-walled TiO2 nanotube-CuxO composites was investigated. Parameters of defects (radicals) in composites were studied. Methanol and methane were detected during the CO2 photoreduction process. In single-walled TiO2 nanotubes, only Ti3+/oxygen vacancy defects were detected. The Cu2+ centers and O2 radicals were found in TiO2 nanotube-CuxO composites using the EPR technique. It has been established that copper oxide nanoparticles are present in the TiO2 nanotube-CuxO composites in the form of the CuO phase. A phase transformation of CuO to Cu2O takes place during illumination, as has been shown by EPR spectroscopy. It is shown that defects accumulate photoinduced charge carriers. The mechanism of methane and methanol formation is discussed. The results obtained are completely original and show high promise for the use of TiO2-CuxO nanotube composites as photocatalysts for CO2 conversion into hydrocarbon fuel precursors.

1. Introduction

An increase in the concentration of atmospheric CO2 is considered one of the main sources of irreversible and dangerous climate change [1,2]. Therefore, conversion of CO2 into useful products is a globally important objective. Different approaches are being developed for CO2 conversion [3]. One promising approach is photocatalytic reduction of CO2 [4]. The issue is that most of the known photocatalysts are too inefficient [5]. Among others, photocatalysts based on nanostructured TiO2 have such important advantages as chemical stability and photocorrosion resistance, and suitable electron energy spectrum [6,7]. TiO2 is also non-toxic material. Titania nanotube arrays (TiO2 NTs) have large specific surface area. Due to their geometry and surface structure, they can promote faster diffusion of photocatalytic reaction products, directional charge transfer, and lower recombination rates of photoinduced charge carriers in comparison to TiO2 nanoparticles [8,9,10]. Highly ordered TiO2 NTs arrays can be formed by electrochemical oxidation of Ti [6,11,12]. TiO2 NTs formed electrochemically in ethylene-glycol-based electrolyte contain an inner layer (IL) enriched with organic products, and an outer layer (OL) of mostly pure TiO2 [13,14]. To improve catalytic activity and obtain crystalline TiO2 structure, as-formed TiO2 NTs arrays are annealed in air. After annealing, the inner layer of a nanotube has a mesoporous structure with a mixture of anatase and rutile nanocrystals with an average size of ~7 nm with the presence of carbon species [15,16]. The outer layer of a nanotube consists of practically stoichiometric anatase [15]. One of the issues of nanostructured TiO2 catalysts is low absorption of visible light due to a wide band gap. Several ways have been proposed to improve the photocatalytic activity of TiO2-based materials under visible-light illumination [6]. Creation of nano-heterojunctions with CuxO species is an effective method [17,18,19,20,21,22,23].
Photocatalytic reduction of CO2 in water and alkaline solutions was shown in [24] using TiO2-CuxO composites with different Cu content deposited onto a molecular sieve. It was demonstrated that these composites enhanced CO2 photoreduction with selective formation of oxalic acid. Furthermore, methanol and acetic acid were observed on the molecular sieve combinate system. In [25], computational modeling and experimental studies were conducted to elucidate the role of Cu sites in promoting CO2 reduction on TiO2. The nanocomposites demonstrated significantly higher activity than bare TiO2 in photocatalysis, and the activity was found to be related to the chemical state of surface Cu+ species. According to modeling studies, the highly dispersed Cu sites likely contributed to the improved photocatalysis by stabilizing surface adsorption of CO2 on TiO2. The experimental and modeling studies further confirmed the direct involvement of surface Cu sites in CO generation via CO2 reduction. Additionally, using TiO2-CuxO heterojunctions can promote photocatalytic decomposition reactions of phenol and acetone [26] and other organic compounds [27,28]. In [29], it was shown that an optimum CuO and TiO2 ratio was necessary to achieve the highest efficiency of methanol conversion to methyl formate. In addition, in [30], the authors describe the TiO2-protecting effect of Cu2O from photocorrosion during CO2 reduction in a saturated solution.
The review [31] has shown that, due to the smaller band gap of metal oxides, such as CuxO, the light absorption capacity of TiO2 can be improved in heterojunctions. Methanol production was reported in CO2 photoreduction process when TiO2 is doped with CuO that provides the highest yield. Cu2O is named in [31] as very promising and the most frequently used dopant candidate for the photoreduction of CO2. It has a band gap ranging from 2 to 2.2 eV that absorbs visible light effectively.
From our point of view, the enhancement of photocatalytic activity in TiO2-CuxO heterojunctions compared to nanostructured TiO2 is due to effective absorption of visible light in CuxO nanoparticles, and charge separation at the heterojunction. The latter reduces the recombination rate of photoexcited electrons and holes.
Activity and selectivity of a TiO2-CuxO catalyst depends on the method of its synthesis [17,18,19,20,21,22,23,32]. Moreover, the efficiency and selectivity of CO2 conversion by TiO2-CuxO catalysts varies during the conversion process [17,18,19,20,21,22,23,32]. One of the possible origins is the reduction of CuO to Cu2O. Additionally, the properties of TiO2-CuxO heterojunctions depend on the catalyst’s microstructure. The efficiency of methanol production on TiO2-NTs-CuxO photocatalysts with different sizes and compositions of copper oxide particles has been studied in [32]. The output of methanol was significantly different for catalysts with different x and size of CuxO particles. The results were explained by the dependence of electron energy spectrum on copper oxide phase composition and by the dependence of methanol production rate on the size, shape and distribution of CuxO particles [32].
To our knowledge, there is a very limited number of articles reporting the study of the photocatalytic activity of TiO2 NTs arrays obtained in ethylene-glycol-based electrolyte and TiO2 NTs-CuxO arrays in the CO2 conversion process in a gas phase in the presence of water vapor [7,9,33]. Moreover, the effect of multilayered structures and the role of inner layers was not discussed much in these publications.
In our previous publication, in [34] we reported a comparative study of the photocatalytic properties of TiO2 multi-walled and single-walled nanotube arrays for photoinduced CO2 conversion in the gas phase, their structures and their defect properties. Electron paramagnetic resonance (EPR) spectroscopy was applied to observe the structural defects in the outer and inner layers of TiO2 nanotubes and analyze their effect on the photocatalytic properties of the nanotubes [35]. The present work aims to study the structural, electronic and photocatalytic properties of TiO2 NTs/CuxO composites obtained from TiO2 NTs arrays without IL by CuxO deposition applying successive ionic layer adsorption and reaction (SILAR). This section may be divided by subheadings. It should provide a concise and precise description of the experimental results, their interpretation and conclusions that can be drawn.

2. Results and Discussion

2.1. Samples Morphology and Composition

Morphology of SW TiO2 NTs/CuxO has been studied using a scanning electron microscope equipped with energy-dispersion X-ray spectroscopy. SEM images of investigated samples are presented in Figure 1.
As can be seen from SEM images, the size of deposited CuxO particles increases with amount of deposition cycles. It is also noticeable that deposition in the described conditions leads to formation of islands and non-homogeneous coverage of the surface of SW TiO2 NTs arrays by CuxO nanoparticles.
The size range of deposited nanoparticles were 9–21 nm, 11–36 nm and 14–40 nm for SW TiO2 NTs/CuxO-10/30/60 samples, respectively. With an increase in the amount of deposition cycles, nanoparticles form agglomerates while the sizes of nanoparticles change slightly and do not depend significantly on the amount of cycles. A characteristic feature of the SW TiO2 NTs/CuxO-60 sample is the presence of nanoparticles with the shape of needles or sheets with the length from 100 nm to 500 nm and 30 to 40 nm thickness. It is well noticeable that CuxO deposition inside the tube does not take place, i.e., only the top surfaces of the SW TiO2 NTs arrays are modified during deposition.
As it is seen from the energy dispersive X-ray microanalysis (Table 1), the amount of copper at the SW TiO2 NTs arrays’ surface varies non-linearly depending on the amount of deposition cycles, and reaches about 3.3 at. % after 60 deposition cycles.
The obtained SW TiO2 NTs and SW TiO2 NTs/CuxO-60 samples were investigated by X-ray diffraction analysis (Figure S1). The copper oxide phase was not recorded by XRD, because of the amount of the CuxO is too small to be registered by this method.

2.2. Optical Properties

The diffuse reflectance spectra depend on the amount of copper oxide, as presented in Figure 2.
According to calculations performed using the Kubelka-Munk theory, the optical band gap is similar for all samples under study: 3.1 ± 0.1 eV for SW TiO2 NTs samples and 3.3 ± 0.1 eV for NTs/CuxO samples. All NTs/CuxO samples showed higher light absorption in visible areas compared to pure SW TiO2 NTs. The samples of SW TiO2 NTs/CuxO-30 possess the most efficient visible light absorption from all samples under study.

2.3. Photocatalytic Conversion of CO2

Prepared samples were investigated in situ during CO2 conversion in gas phase in the presence of water vapor. The products of photoelectrochemical reactions were detected by chromatograph equipped with a plasma ionization detector (PID). The outcome of reaction products was calculated per unit of specific surface area for each sample 7.7 × 10−3 m2/cm2. The specific surface area has been calculated in the frame of a simple geometric model using scanning electron microscopy data (Figure S2); the methodology was described in [34].
Three LEDs operating at 370 nm main wavelength were used for illumination. Illumination intensity at the sample was about 23 mW/cm2. The measurement process was carried out in four stages: (1) pumping of helium through the reactor during 10 h; (2) cleaning of the sample surface from organic impurities in helium flux under LED illumination; (3) photoinduced conversion of CO2 (helium flux replaced by CO2 flux); (4) registration of relaxation after turning off illumination. Gas flux was set to 1.1 mL/min. Results are presented in Figure 3. Methane and methanol were detected as main reaction products. Acetaldehyde, ethanol, formic acid and acetone were detected in helium flux under illumination, but their amount did not increase after CO2 supplement. The kinetics of acetaldehyde formation is presented in the Supplementary Material, Figure S3.
Only two compounds increased in concentration in the presence of CO2 flux: methane and methanol. This is evidence of photoinduced CO2 conversion. One can see that maximum methanol outcome was observed for the SW TiO2 NTs sample while minimum methane output was observed for this sample. Deposition of copper oxide shifts the selectivity of the conversion process to methane production. Output of methane reaches its maximum for the SW TiO2 NTs/CuxO-30 sample. Increasing the number of deposition cycles from 10 to 60 led to an increase in methanol output. However, the rate of methanol production for the samples with deposited copper oxide remains smaller than for the SW TiO2 NTs sample.
Variation of CO2 conversion selectivity can be explained by the difference of methane and methanol formation potentials and changing of conduction band potentials of TiO2 in contact with CuO. Potentials of reactions of methane (1) and methanol (2) formations are equal to −0.24 V and −0.38 V versus normal hydrogen electrode (NHE), respectively [36].
CO 2 + 8 H + + 8 e CH 4 + H 2 O E Redox = 0.24 V   vs .   NHE
CO 2 + 6 H + + 6 e CH 3 OH + H 2 O E Redox = 0.38 V   vs .   NHE
According to photoelectrochemical data, the conduction band bottom potential in TiO2 is more negative compared to methanol reaction potential [37]. Therefore, reactions (1) and (2) can take place at the surface of TiO2. Redistribution of charge carriers between two semiconductors occurs after deposition of copper oxide on the surface of anodic titania nanotube arrays. This leads to the bending of energy bands in semiconductors (Figure 4). The band edges positions of copper oxides are indicated according data in the literature [18,33,38].
As a result of the band-bending, the conduction band potential in TiO2 becomes more positive than the reduction potential for the reaction (2). The latter can block the reaction. Illumination, in turn, shifts energy band potentials in TiO2 to more negative values due to the migration of photoelectrons from CuO to TiO2. This can lead to an increased probability of the reaction occurring (2). Variation (increase) of an amount of copper oxide leads in turn to the enhancement of the concentration of photogenerated charge carriers and thus to more significant flattening of energy band in TiO2.
This mechanism makes the shift of CO2 photoconversion selectivity towards methane production, caused by deposition of copper oxide on TiO2 NTs arrays, understandable. One can suggest that the conduction band potential of TiO2 in NTs/CuxO-10 sample is appropriate for CO2 photoconversion to methane but not to methanol. For the NTs/CuxO-30 sample, the conduction band potential is also not negative enough for the reaction (2). However, methane output increases due to larger photoelectron concentration separated at the heterojunction of two semiconductors. Deposition of a larger amount of copper oxide (NTs/CuxO-60 sample) makes band flattening under illumination large enough for the reaction (2). However, efficiency of methanol production for this sample is smaller than for non-modified sample SW TiO2 NTs. This fact can be linked to the impossibility of water decomposition on the surface of copper oxide due to insufficient positive potential of the top of valence band. As a consequence, hydrogen ions and hydroxyl ions can be formed only on the surface of TiO2. The free surface of TiO2 of NTs decreases with an increase in the amount of copper oxide deposition cycles resulting in a reduction in the rate of CO2 photoconversion to methanol.

2.4. Electron Paramagnetic Resonance

The nanocomposites obtained were further investigated by electron paramagnetic resonance spectroscopy (EPR) since defects (spin centers) play an important role in photocatalytic processes [6]. The main type of defect in initial SW TiO2 NTs is Ti3+/oxygen vacancy (g1 = 1.9961, g2 = 1.9697). Figure 5a shows the EPR spectra of SW TiO2 NTs in the dark and under illumination. It can be seen (Figure 5a) that, upon photoexcitation, the intensity of the EPR signal (IEPR) increases slightly.
Figure 5b shows an example of the EPR spectrum of the SW TiO2 NTs/CuxO-30 nanocomposites. This spectrum is a superposition of several EPR signals. First, a strong EPR signal from copper ions Cu2+ (g = 2.1612) [39] is detected (Figure 5b), which indicates the presence of the CuO phase. We also observed on the right side of the EPR spectrum the signal from O2 radicals (g1 = 2.029, g2 = 2.009, g3 = 2.003). The appearance of O2 radicals can be explained by the adsorption of oxygen molecules on oxygen vacancies on the surface of TiO2 and, probably, on the surface of copper oxide nanoparticles, followed by the capture of electrons from the conduction band. Upon illumination, the EPR signal intensity from Cu2+ decreases while EPR signal intensity from O2 radicals increases (Figure 5b). We suppose that paramagnetic defects Cu2+ pass into a non-paramagnetic state (Cu2+ + e → Cu+), which leads to a diminution of corresponding EPR signal (Figure 5b). At the same time, the capture of a photoinduced electron by an oxygen molecule leads to an increase in the EPR signal from O2 radicals: O2 + e → O2. The generation of Cu+ ions under illumination indicates the formation of Cu2O phase.
Notice that we did not observe the EPR signal in SW TiO2 NTs/CuxO nanocomposites from the Ti3+/oxygen vacancy centers, probably against the background of strong EPR signals from copper ions and oxygen radicals. Under illumination, they are also not recorded, as these defects probably capture the photoexcited hole and pass into a non-paramagnetic state: Ti3+ + h → Ti4+. Therefore, in SW TiO2 NTs/CuxO nanocomposites the defects accumulate the charge necessary for redox reactions. We also investigated the behavior of spin centers after illumination. Figure 5b shows EPR spectrum of SW TiO2 NTs/CuxO-30 one hour after illumination (spectrum 3). It can be seen that the EPR signal intensity from Cu2+ increases a little but does not reach the value before illumination, while EPR signal intensity from O2 radicals strongly decreases to the value before illumination. Therefore, a phase transformation of CuO to Cu2O takes place during illumination. A reduction of O2 radicals after illumination can be due to spending radicals in redox reactions.

2.5. Secondary Ion Mass Spectroscopy

To understand the nature of copper oxide alternation on TiO2 surface, two parts of the same sample were studied before and after illumination by means of secondary ion mass spectroscopy (SIMS).
It has been found that emission of CuO and CuO2− (normalized to total amount of detected ions) is much higher from the sample before illumination compared to emission after illumination. Moreover, there is a difference in the ratio of amount of emitted CuO2− and CuO ions before and after illumination. This difference can point to the different stoichiometry of copper oxides on the surface of the samples before and after illumination. One can suggest that the amount of oxygen bound to copper is larger before illumination. The data obtained provide evidence to suggest that copper oxide on the surface of the samples is most likely in the CuO phase before illumination. However, after illumination, copper oxide transforms to a less oxidized Cu2O phase, which is consistent with the EPR data. We suppose that copper oxide operates as a source of photogenerated electrons. Their quantity controls potentials of energy bands in TiO2.

2.6. Role of Cu2O in Photocatalytic Conversion of CO2

Taking into account transformation of CuO to Cu2O phase, one can present an energy diagram of the structure as an energy diagram of heterostructure consisting of three semiconductors. Transformation of CuO to Cu2O can take place at the TiO2/CuO boundary and at CuO free surface (Figure 6). The change of copper oxide phase will not lead to significant shift of Fermi level position due to the small amount of Cu2O formed compared to the amount of titania. For both possible positions of Cu2O phase shown in Figure N, the amount of photoelectrons transferred to TiO2 will be reduced due to their additional recombination at the CuO/Cu2O boundary or accumulation in the CuO phase followed by CuO transformation to Cu2O and to Cu [32]. Therefore, elevating of Fermi level caused by illumination will be smaller in the structure containing Cu2O than in the structure with CuO phase only. Thus, the probability of methanol formation should be reduced due to CuO to Cu2O phase transformation. The probability of methane formation can also be reduced but less significantly due to the less negative potential of methane formation.
Based on the data obtained, one can suggest that the efficiency and selectivity of CO2 conversion in the Cu2O-CuO-TiO2 system is determined by the number of photoelectrons transferred to TiO2. This number determines the rise of Fermi level in TiO2 under illumination and its position with respect to potentials for CO2 to methane or methanol conversion. At high concentrations of photoelectrons, the position of the Fermi level under illumination is high enough for effective conversion of CO2 to methanol (2). At lower concentrations of photoelectrons, the position of the Fermi level is too low for effective methanol production. In this case, methane production (1) dominates.

3. Materials and Methods

Synthesis of single-wall (SW) nanotubes of anodic titania is described in [34]. SW TiO2 NTs/CuxO nanocomposites were prepared by deposition of copper oxide on the surface of titania nanotubes arrays using successive ionic layer adsorption and reaction (SILAR). An aqueous solution of CuCl2·2H2O served as ionic source. An aqueous solution of 25% NH4OH (Russian Specifications GOST 3760-79, Sigma Tec, Khimki, Russia) was added to increase the pH to 10. Mixture of ethanol with deionized water in 1:3 proportion heated up to 70 °C was used as the source of anions. The SILAR method was carried out in three stages. In the first stage, the sample was inserted into an aqueous solution of copper chloride containing [Cu(NH3)4]2+ ions for 30 s. In the second stage, the sample was inserted into a mixture of ethanol with deionized water for 7 s. In the third stage, the sample washed in deionized water for 30 s.
Variation of the amount of copper oxide CuxO deposited on SW TiO2 NTs was performed by variation of cycles of ionic layer formation (10, 30 and 60 cycles were carried out for samples SW TiO2 NTs/CuxO-10, SW TiO2 NTs/CuxO-30 and SW TiO2 NTs/CuxO-60, respectively). After deposition, the samples were heated up to 300 °C at a rate of 30 °C/min and kept at 300 °C in air for 60 min.
The surface morphologies and compositions were studied using a Helios G4CX (Thermo Fisher Scientific, Waltham, MA, USA) scanning electron microscope equipped with an EDS attachment (EDAX Octane Elite Super).
Spectra of diffuse light reflection from SW TiO2 NTs and SW TiO2 NTs/CuxO samples were recorded on LS-55 Perkin Elmer spectrometer (Waltham, MA, USA) which allows for the registration of the diffuse light scattering from the sample surface in a spectral range from 200 to 800 nm with spectral slit widths from 2.5 to 20 nm.
Electron paramagnetic resonance (EPR) spectra were recorded on a Bruker ELEXSYS E500 EPR spectrometer (Billerica, MA, USA) (X-band). The samples were illuminated directly in the cavity of the EPR spectrometer with the light of a BRUKER ELEXSYS ER 202 UV high-pressure mercury lamp (50 W). The photoexcitation intensity of the samples was 40 mW/cm2.
Secondary ion mass spectroscopy was produced by TOF-SIMS IV manufactured by IONTOF GmbH, Muenster, Germany, and clusters of bismuth ion (Bi3+) were used.
The details of the methodology of the photocatalytic conversion of CO2 were presented earlier in the work [34]. The photocatalytic CO2 conversion was provided under the following conditions: reactor temperature 30 °C, relative humidity 5%, total CO2 (ISO 14175-C1) gas flow 1.2 mL/min. The analysis of the gas products was carried out using an HP PLOT/Q capillary column Hewlett Packard HP 5890 Series II Gas Chromatograph-FID (Agilent Technologies, Santa Clara, CA, USA).

4. Conclusions

Single-walled TiO2 nanotube-CuxO composites for converting carbon dioxide (CO2) into hydrocarbon compounds were prepared. Methanol and methane were detected during the CO2 photoreduction process. The production rate of gas-phase CO2 conversion and photoinduced transformations of defects in the prepared photocatalysts were investigated. Only Ti3+/oxygen vacancy defects were detected by EPR spectroscopy in initial TiO2 nanotubes. Cu2+ centers and O2 radicals were detected in the TiO2 nanotube–CuxO composites. Under illumination, the capture of a photoinduced electron by oxygen molecules led to an increase in the EPR signal from O2 radicals. It was shown by EPR technique that CuxO nanoparticles are present in the TiO2 nanotube-CuxO composites in the form of the CuO phase. Under illumination, a phase transformation of CuO to Cu2O was observed. These results are consistent with data obtained by secondary ion mass spectroscopy. The shift of CO2 photoconversion selectivity was observed towards methane production due to the deposition of copper oxide on the surface of TiO2 nanotubes. The mechanism of methane and methanol formation in the TiO2 nanotubes with different copper oxide content is discussed. The obtained results are completely novel and show a high level of importance for the development of energy-efficient photocatalysts for CO2 conversion into hydrocarbon fuel precursors.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12091011/s1. Figure S1: XRD spectra of the SW TiO2 NTs and SW TiO2 NTs/CuxO-60 samples. Figure S2: Geometric scheme of the nanotube array, Figure S3: Kinetics of acetaldehyde formation.

Author Contributions

Conceptualization, T.P.S., E.A.K. and T.M.; data curation, E.A.K. and V.G.K.; formal analysis, T.P.S., E.V.K. and V.B.Z.; investigation, T.P.S., E.V.K., O.P. and V.B.Z.; methodology, T.P.S., E.V.K., O.P. and T.M.; project administration, E.A.K.; resources, A.K.T. and T.M.; software, A.K.T.; supervision, E.A.K. and T.M.; validation, V.G.K.; visualization, T.P.S. and V.G.K.; writing—original draft, T.P.S. and E.A.K.; writing—review & editing, E.V.K., E.A.K. and V.G.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded a grant from Russian Science Foundation No. 21-19-00494.

Acknowledgments

The work was supported by a grant from Russian Science Foundation No 21-19-00494, https://rscf.ru/en/project/21-19-00494/ (accessed on 29 August 2022). The EPR experiments were performed using the facilities of the Collective Use Center at the Moscow State University (including the Bruker ER 4112 HV temperature control system of the Moscow State University Development Program). We thank the team of the Institute of General and Ecological Chemistry of Lodz University of Technology for assistance in research and comprehensive assistance.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Smith, P.; Davis, S.J.; Creutzig, F.; Fuss, S.; Minx, J.; Gabrielle, B.; Kato, E.; Jackson, R.B.; Cowie, A.; Kriegler, E.; et al. Biophysical and economic limits to negative CO2 emissions. Nat. Clim. Change 2016, 6, 42–50. [Google Scholar] [CrossRef]
  2. Vijaya, V.; Raman, S.; Iniyan, S.; Goic, R. A review of climate change, mitigation and adaptation. Renew. Sustain. Energy Rev. 2012, 16, 878–897. [Google Scholar]
  3. Saravanan, A.; Senthil Kumar, P.; Dai-Viet, N.V.; Jeevanantham, S.; Bhuvaneswari, V.; Anantha, N.V.; Yaashikaa, P.R.; Swetha, S.; Reshma, B. A comprehensive review on different approaches for CO2 utilization and conversion pathways. Chem. Eng. Sci. 2021, 236, 116515. [Google Scholar] [CrossRef]
  4. Kovacič, Ž.; Likozar, B.; Hus, M. Photocatalytic CO2 Reduction: A Review of Ab Initio Mechanism, Kinetics, and Multiscale Modeling Simulations. ACS Catal. 2021, 10, 14984–15007. [Google Scholar] [CrossRef]
  5. Sorcar, S.; Hwang, Y.; Lee, J.; Kim, H.; Grimes, K.M.; Grimes, C.A.; Jung, J.-W.; Cho, C.-H.; Majima, T.; Hoffmann, M.R.; et al. CO2, water, and sunlight to hydrocarbon fuels: A sustained sunlight to fuel (Joule-to-Joule) photoconversion efficiency of 1%. Energy Environ. Sci. 2019, 12, 2685–2696. [Google Scholar] [CrossRef]
  6. Rawool, S.A.; Yadav, K.K.; Polshettiwar, V. Defective TiO2 for photocatalytic CO2 conversion to fuels and chemicals. Chem. Sci. 2021, 12, 4267–4269. [Google Scholar] [CrossRef] [PubMed]
  7. Zubair, M.; Kim, H.; Razzaq, A.; Grimes, C.A.; In, S.-I. Solar spectrum photocatalytic conversion of CO2 to CH4 utilizing TiO2 nanotube arrays embedded with graphene quantum dots. J. CO2 Util. 2018, 26, 70–79. [Google Scholar] [CrossRef]
  8. Liu, Z.; Zhang, X.; Nishimoto, S.; Murakami, T.; Fujishima, A. Efficient Photocatalytic Degradation of Gaseous Acetaldehyde by Highly Ordered TiO2 Nanotube Arrays. Environ. Sci. Technol. 2008, 42, 8547–8551. [Google Scholar] [CrossRef]
  9. Low, J.; Qiu, S.; Xu, D.; Jiang, C.; Cheng, B. Direct evidence and enhancement of surface plasmon resonance effect on Ag-loaded TiO2 nanotube arrays for photocatalytic CO2 reduction. Appl. Surf. Sci. 2018, 434, 423–432. [Google Scholar] [CrossRef]
  10. Varghese, O.K.; Paulose, M.; LaTempa, T.J.; Grimes, C.A. High-Rate Solar Photocatalytic Conversion of CO2 and Water Vapor to Hydrocarbon Fuels. Nano Lett. 2009, 9, 731–737. [Google Scholar] [CrossRef]
  11. Liu, N.; Mirabolghasemi, H.; Lee, K.; Albu, S.P.; Tighineanu, A.; Altomare, M.; Schmuki, P. Anodic TiO2 nanotubes: Double walled vs. single walled. Faraday Discuss. 2013, 164, 107. [Google Scholar] [CrossRef] [PubMed]
  12. Belov, A.N.; Gavrilin, I.M.; Gavrilov, S.A.; Dronov, A.A.; Labunov, V. Effect of the activity of fluorine-containing electrolytes on reaching the maximum thickness of porous anodic titanium oxide. Semiconductors 2013, 47, 1707–1710. [Google Scholar] [CrossRef]
  13. So, S.; Riboni, F.; IHwang Paul, D.; Hammond, J.; Tomanec, O.; Zboril, R.; Sadoway, D.R.; Schmuki, P. The double-walled nature of TiO2 nanotubes and formation of tube-in-tube structures—A characterization of different tube morphologies. Electrochim. Acta 2017, 231, 721–731. [Google Scholar] [CrossRef]
  14. Dronov, A.; Gavrilin, I.; Kirilenko, E.; Dronova, D.; Gavrilov, S. Investigation of anodic TiO2 nanotube composition with high spatial resolution AES and ToF SIMS. Appl. Surf. Sci. 2018, 434, 148–154. [Google Scholar] [CrossRef]
  15. Gavrilin, I.; Dronov, A.; Volkov, R.; Savchuk, T.; Dronova, D.; Borgardt, N.; Pavlikov, A.; Gavrilov, S.; Gromov, D. Differences in the local structure and composition of anodic TiO2 nanotubes annealed in vacuum and air. Appl. Surf. Sci. 2020, 516, 146120. [Google Scholar] [CrossRef]
  16. Motola, M.; Sopha, H.; Krbal, M.; Hromádko, L.; Zmrhalová, Z.O.; Plesch, G.; Macak, J.M. Comparison of photoelectrochemical performance of anodic single- and double-walled TiO2 nanotube layers. Electrochem. Commun. 2018, 97, 1–5. [Google Scholar] [CrossRef]
  17. Xi, H.; Xu, Y.; Zou, W.; Ji, J.; Cai, Y.; Wan, H.; Dong, L. Enhanced methanol selectivity of CuxO/TiO2 photocatalytic CO2 reduction: Synergistic mechanism of surface hydroxyl and low-valence copper species. J. CO2 Util. 2022, 55, 101825. [Google Scholar] [CrossRef]
  18. Park, S.-M.; Razzaq, A.; Park, Y.H.; Sorcar, S.; Park, Y.; Grimes, C.A.; In, S.-I. Hybrid CuxO−TiO2 Heterostructured Composites for Photocatalytic CO2 Reduction into Methane Using Solar Irradiation: Sunlight into Fuel. ACS Omega 2016, 1, 868–875. [Google Scholar] [CrossRef]
  19. Miyauchi, M.; Sunada, K.; Hashimoto, K. Antiviral Effect of Visible Light-Sensitive CuxO/TiO2 Photocatalyst. Catalysts 2020, 10, 1093. [Google Scholar] [CrossRef]
  20. Bharad, P.A.; Nikam, A.V.; Thomas, F.; Gopinath, C.S. CuOx-TiO2 Composites: Electronically Integrated Nanocomposites for Solar Hydrogen Generation. Chem. Sel. 2018, 3, 12022–12030. [Google Scholar]
  21. Wang, K.; Bielan, Z.; Endo-Kimura, M.; Janczarek, M.; Zhang, D.; Kowalski, D.; Zielińnska-Jurek, A.; Markowska-Szczupak, A.; Ohtani, B.; Kowalska, E. On the mechanism of photocatalytic reactions on CuxO@TiO2 core–shell photocatalysts. J. Mater. Chem. A 2021, 9, 10135–10145. [Google Scholar] [CrossRef]
  22. Saedy, S.; Hiemstra, N.; Benz, D.; Bui, H.V.; Nolan, M.; van Ommen, J.R. Dual promotional effect of CuxO clusters grown with atomic layer deposition on TiO2 for photocatalytic hydrogen production. Catal. Sci. Technol. 2022, 22, 4511–4523. [Google Scholar] [CrossRef]
  23. Li, H.; Zhou, J.; Feng, B. Facile synthesis of CuxO (x = 1, 2)/TiO2 nanotube arrays as an efficient visible-light driven photocatalysts. J. Porous Mater. 2017, 24, 97–102. [Google Scholar] [CrossRef]
  24. Srivinas, B.; Shubhamangala, B.; Lalitha, K.; Kumar Reddy, P.A.; Kumari, V.D.; Subrahmanyam, M.; De, B.R. Photocatalytic Reduction of CO2 over Cu-TiO2/ Molecular Sieve 5A Composite. Photochem. Photobiol. 2011, 87, 995–1001. [Google Scholar]
  25. Liu, C.; Iyemperumal, S.K.; Deskins, N.A.; Li, G. Photocatalytic CO2 Reduction by Highly Dispersed Cu Sites on TiO2. J. Photonics Energy 2016, 7, 012004. [Google Scholar] [CrossRef]
  26. Méndez-Medrano, M.G.; Kowalska, E.; Ohtani, B.; Bahena Uribe, D.; Colbeau-Justin, C.; Rau, S.; Rodríguez-López, J.L.; Remita, H. Heterojunction of CuO Nanoclusters with TiO2 for Photo-Oxidation of Organic Compounds and for Hydrogen Production. J. Chem. Phys. 2020, 153, 034705. [Google Scholar] [CrossRef] [PubMed]
  27. Janczarek, M.; Kowalska, E. On the Origin of Enhanced Photocatalytic Activity of Copper-Modified Titania in the Oxidative Reaction Systems. Catalysts 2017, 7, 317. [Google Scholar] [CrossRef]
  28. Hamad, H.; Elsenety, M.M.; Sadik, W.; El-Demerdash, A.G.; Nashed, A.; Mostafa, A.; Elyamny, S. The Superior Photocatalytic Performance and DFT Insights of S-Scheme CuO@TiO2 Heterojunction Composites for Simultaneous Degradation of Organics. Sci. Rep. 2022, 12, 317. [Google Scholar] [CrossRef]
  29. Shi, Q.; Ping, G.; Wang, X.; Xu, H.; Li, J.; Cui, J.; Abroshan, H.; Ding, H.; Li, G. CuO/TiO2 Heterojunction Composites: An Efficient Photocatalyst for Selective Oxidation of Methanol to Methyl Formate. J. Mater. Chem. A 2019, 7, 2253–2260. [Google Scholar] [CrossRef]
  30. Aguirre, M.E.; Zhou, R.; Eugene, A.J.; Guzman, M.I.; Grela, M.A. Cu2O/TiO2 Heterostructures for CO2 Reduction through a Direct Z-Scheme: Protecting Cu2O from Photocorrosion. Appl. Catal. B Environ. 2017, 217, 485–493. [Google Scholar] [CrossRef]
  31. Rehman, Z.U.; Bilal, M.; Hou, J.; Butt, F.K.; Ahmad, J.; Ali, S.; Hussain, A. Photocatalytic CO2 Reduction Using TiO2-Based Photocatalysts and TiO2 Z-Scheme Heterojunction Composites: A Review. Molecules 2022, 27, 2069. [Google Scholar] [CrossRef] [PubMed]
  32. De Almeida, J.; Pacheco, M.S.; de Brito, J.F.; de Arruda Rodrigues, C. Contribution of CuxO Distribution, Shape and Ratio on TiO2 Nanotubes to Improve Methanol Production from CO2 Photoelectroreduction. J. Solid State Electrochem. 2020, 24, 3013–3028. [Google Scholar] [CrossRef]
  33. Li, Y.; Zhang, W.; Shen, X.; Peng, P.; Xiong, L.; Yu, Y. Octahedral Cu2O-modified TiO2 nanotube arrays for efficient photocatalytic reduction of CO2. Chin. J. Catal. 2015, 36, 2229–2236. [Google Scholar] [CrossRef]
  34. Savchuk, T.; Gavrilin, I.; Konstantinova, E.; Dronov, A.; Volkov, R.; Borgardt, N.; Maniecki, T.; Gavrilov, S.; Zaitsev, V. Anodic TiO2 nanotube arrays for photocatalytic CO2 conversion: Comparative photocatalysis and EPR study. Nanotechnology 2022, 33, 055706. [Google Scholar] [CrossRef] [PubMed]
  35. Sviridova, T.V.; Sadovskaya LYu Konstantinova, E.A.; Belyasova, N.A.; Kokorin, A.I.; Sviridov, D.V. Photoaccumulating TiO2–MoO3, TiO2–V2O5, and TiO2–WO3 Heterostructures for Self-Sterilizing Systems with the Prolonged Bactericidal Activity. Catal. Lett. 2019, 149, 1147–1153. [Google Scholar] [CrossRef]
  36. Abdullah, H.; Khan, M.M.R.; Ong, H.R.; Yaakob, Z. Modified TiO2 Photocatalyst for CO2 Photocatalytic Reduction: An Overview. J. CO2 Util. 2017, 22, 15–32. [Google Scholar] [CrossRef]
  37. Ola, O.; Maroto-Valer, M.M. Review of Material Design and Reactor Engineering on TiO2 Photocatalysis for CO2 Reduction. J. Photochem. Photobiol. C Photochem. Rev. 2015, 24, 16–42. [Google Scholar] [CrossRef]
  38. Tan, Y.; Wei, Y.; Xu, Y.; Liang, K.; Jiao, Y.; Zhang, S. A Novel Nano-Pore Stainless Steel Film Modified by Cu@CuO/Cu2O Semiconductor Heterojunction for Photoelectrocatalytic Activity and Semiconductor Characteristics. ECS J. Solid State Sci. Technol. 2021, 10, 063002. [Google Scholar] [CrossRef]
  39. Martín-Gomez, J.; Hidalgo-Carrillo, J.; Montes, V.; Estevez-Toledano, R.C.; Escamilla, J.C.; Marinas, A.; Urbano, F.J. EPR and CV studies cast further light on the origin of the enhanced hydrogen production through glycerol photoreforming on CuO:TiO2 physical mixtures. J. Environ. Chem. Eng. 2021, 9, 105336. [Google Scholar] [CrossRef]
Figure 1. SEM images of SW TiO2 NTs array surface before and after copper oxide CuxO deposition.
Figure 1. SEM images of SW TiO2 NTs array surface before and after copper oxide CuxO deposition.
Catalysts 12 01011 g001
Figure 2. Normalized Spectra of Diffuse Reflection of Light from the SW TiO2 NTs/CuxO-10 (1), SW TiO2 NTs/CuxO-30 (2) and SW TiO2 NTs/CuxO-60 (3).
Figure 2. Normalized Spectra of Diffuse Reflection of Light from the SW TiO2 NTs/CuxO-10 (1), SW TiO2 NTs/CuxO-30 (2) and SW TiO2 NTs/CuxO-60 (3).
Catalysts 12 01011 g002
Figure 3. Kinetics of CO2 photoinduced conversion in gas phase under illumination UV-A (370 nm): to methane (a), to methanol (b).
Figure 3. Kinetics of CO2 photoinduced conversion in gas phase under illumination UV-A (370 nm): to methane (a), to methanol (b).
Catalysts 12 01011 g003
Figure 4. Energy band bending model in NTs/CuxO composites.
Figure 4. Energy band bending model in NTs/CuxO composites.
Catalysts 12 01011 g004
Figure 5. (a) EPR spectra of SW TiO2 NTs in dark (1) and under illumination (2). (b) EPR spectra of SW TiO2 NTs/CuxO-30 in dark (1), under illumination (2) and in dark after illumination (3).
Figure 5. (a) EPR spectra of SW TiO2 NTs in dark (1) and under illumination (2). (b) EPR spectra of SW TiO2 NTs/CuxO-30 in dark (1), under illumination (2) and in dark after illumination (3).
Catalysts 12 01011 g005
Figure 6. Energy band bending model in NTs/CuO-Cu2O composites.
Figure 6. Energy band bending model in NTs/CuO-Cu2O composites.
Catalysts 12 01011 g006
Table 1. EDX microanalysis data for SW TiO2 NTs arrays with different copper content.
Table 1. EDX microanalysis data for SW TiO2 NTs arrays with different copper content.
SW TiO2 NTsSW TiO2 NTs/
CuxO-10
SW TiO2 NTs/
CuxO-30
SW TiO2 NTs/
CuxO-60
Ti, at. %31.932.934.031.4
O, at. %55.758.156.658.5
F, at. %0.30.70.00.4
C, at. %4.46.43.86.0
Cu, at. %0.02.23.03.3
Al, at. %0.50.30.20.3
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Savchuk, T.P.; Kytina, E.V.; Konstantinova, E.A.; Kytin, V.G.; Pinchuk, O.; Tarhanov, A.K.; Zaitsev, V.B.; Maniecki, T. Photocatalytic CO2 Conversion Using Anodic TiO2 Nanotube-CuxO Composites. Catalysts 2022, 12, 1011. https://doi.org/10.3390/catal12091011

AMA Style

Savchuk TP, Kytina EV, Konstantinova EA, Kytin VG, Pinchuk O, Tarhanov AK, Zaitsev VB, Maniecki T. Photocatalytic CO2 Conversion Using Anodic TiO2 Nanotube-CuxO Composites. Catalysts. 2022; 12(9):1011. https://doi.org/10.3390/catal12091011

Chicago/Turabian Style

Savchuk, Timofey P., Ekaterina V. Kytina, Elizaveta A. Konstantinova, Vladimir G. Kytin, Olga Pinchuk, Andrey K. Tarhanov, Vladimir B. Zaitsev, and Tomasz Maniecki. 2022. "Photocatalytic CO2 Conversion Using Anodic TiO2 Nanotube-CuxO Composites" Catalysts 12, no. 9: 1011. https://doi.org/10.3390/catal12091011

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