Next Article in Journal
The Promoting Effect of Ce on the Performance of Au/CexZr1−xO2 for γ-Valerolactone Production from Biomass-Based Levulinic Acid and Formic Acid
Next Article in Special Issue
Highly Active TiO2 Microspheres Formation in the Presence of Ethylammonium Nitrate Ionic Liquid
Previous Article in Journal
A Novel High-Activity Zn-Co Catalyst for Acetylene Acetoxylation
Previous Article in Special Issue
Electrochemically Obtained TiO2/CuxOy Nanotube Arrays Presenting a Photocatalytic Response in Processes of Pollutants Degradation and Bacteria Inactivation in Aqueous Phase
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synergistic Effect of Cu2O and Urea as Modifiers of TiO2 for Enhanced Visible Light Activity

1
Institute for Catalysis, Hokkaido University, N21, W10, Sapporo 001-0021, Japan
2
Department of Chemical Technology, Gdansk University of Technology, Narutowicza Str. 11/12, 80-233 Gdansk, Poland
*
Author to whom correspondence should be addressed.
Catalysts 2018, 8(6), 240; https://doi.org/10.3390/catal8060240
Submission received: 14 March 2018 / Revised: 1 June 2018 / Accepted: 1 June 2018 / Published: 6 June 2018
(This article belongs to the Special Issue Nanomaterials for Environmental Purification and Energy Conversion)

Abstract

:
Low cost compounds, i.e., Cu2O and urea, were used as TiO2 modifiers to introduce visible light activity. Simple and cheap methods were applied to synthesize an efficient and stable nanocomposite photocatalytic material. First, the core-shell structure TiO2–polytriazine derivatives were prepared. Thereafter, Cu2O was added as the second semiconductor to form a dual heterojunction system. Enhanced visible light activity was found for the above-mentioned nanocomposite, confirming a synergistic effect of Cu2O and urea (via polytriazine derivatives on titania surface). Two possible mechanisms of visible light activity of the considered material were proposed regarding the type II heterojunction and Z-scheme through the essential improvement of the charge separation effect.

1. Introduction

Titanium dioxide (TiO2, titania) is widely recognized as an efficient, stable, and green photocatalytic material (long term stability, chemical inertness and corrosion resistance). Therefore, its application potential in photocatalysis is still growing and presently focuses on emergency areas, such as environmental remediation (water treatment and air purification), renewable energy processes (i.e., photocurrent generation, water splitting for hydrogen production, conversion of CO2 to hydrocarbons) and self-cleaning surfaces [1,2,3,4,5]. However, the application of titania is still limited to regions with a high intensity of solar radiation due to its wide bandgap (ca. 3.0 to 3.2 eV). The following strategies of titania doping, modification, semiconductor coupling, and dye sensitization can be applied to incorporate visible light absorption to TiO2 [6,7,8]. The nature of electron transfer between TiO2 and other materials has been intensively studied and recognized as the origin of the high performances of TiO2-based composites [9]. The photocatalytic activity of TiO2 systems depend on the following properties, such as particle size, surface area, crystal phase, morphology, uncoordinated surface sites, defects in the lattice, and degree of crystallinity. Design of TiO2 composite structures based on the heterojunction between titania and other semiconducting materials can improve many of these properties. Moreover, this strategy can create and tune other properties, such as mid-band-gap electronic states, which can be responsible for the intensification of charge separation or incorporation of a red shift to the absorption spectrum [7,9].
Copper oxides (Cu2O, CuO) have been intensively studied as titania modifiers due to their intrinsic p-type configuration. Cu2O and CuO are inexpensive semiconductors with band gap energies of 2.1–2.2 eV and 1.2–1.7 eV, respectively. This fact makes both materials promising for research directed to solar energy utilization [10,11,12]. If the electronic properties of both oxides are compared, CuO has a significantly smaller band gap than Cu2O, and thus can absorb more photons, but the positions of CB and VB for CuO are insufficient to catalyze the generation of hydroxyl and superoxide radicals, which are the primary initiators for the photocatalytic oxidation of organic compounds [13,14]. Therefore, the photocatalytic oxidation of organic compounds over Cu2O/TiO2 composites has been of particular interest, especially to introduce visible light activity to TiO2-based systems [15,16,17,18,19]. For example, Liu et al. prepared Cu2O/TiO2 composites where titania was in the form of nanosheets with exposed {001} facets. They reported visible light photocatalytic activity for Cu2O/TiO2 nanosheets three times higher than that for nitrogen-doped titania nanosheets. Owing to the type II heterojunction between Cu2O and TiO2, an efficient charge separation is observed and visible light induced electron transfer from Cu2O to TiO2 occurs, resulting in visible light photocatalytic performance of the system [17]. Furthermore, Cu2O possesses promising application potential because of its very good antipathogenic properties, even better than that of metallic copper [20,21,22,23].
Apart from p-type semiconducting metal oxides as composite junctions for titania, a promising option to prepare efficient and low cost visible light-active material based on TiO2 is the application of organic compounds like urea as a modifier. Urea-derived titania materials were initially recognized as nitrogen “doped” TiO2 photocatalysts [24,25,26,27]. Subsequently, the presence of nitridic and amidic species or nitrogen species with several oxidation states of nitrogen were suggested [25,26]. Finally, Mitoraj and Kisch proposed another explanation for the nature of urea-modification of titania including both nitrogen and carbon originating from urea as the elements building the chemical structure responsible for visible light-sensitization of TiO2 [28,29,30]. They reported that thermal processing of urea with titania at 400 °C produced poly(amino-tri-s-triazine) derivatives (shell) covalently attached to the semiconductor (core), i.e., a unique example of inorganic with an organic (polytriazine as a crystalline layer) semiconductor connected through the Ti–N–C bond. Titania acts as a catalyst in this urea transformation. Condensation between the triazine amino and titania hydroxy groups forms Ti–N bonds. Considered chemical structures arise from condensed aromatic s-triazine compounds containing melem and melon (trimer of melem) units, which form a visible light absorbing semiconducting organic layer coupled with titania. Compared to unmodified TiO2, the prepared photocatalyst may exhibit a band-gap narrowing. It has been proposed that the absorption shoulder in the visible region corresponding to this material is a charge-transfer band enabling an optical transfer from polytriazine component to titania [29]. In contrast, it is known that prolonged heating of urea at 550 °C produces polycondensed s-triazines with a graphitic structure (g-C3N4), which was recognized as a separate metal-free polymeric n-type semiconductor (band gap energy equal to 2.7 eV) with the possibility to create heterojunctions with other semiconductors, e.g., titania [31,32,33,34]. Therefore, thermal treatment of a TiO2/urea composite at higher temperature results in the formation of a TiO2/g-C3N4 heterojunction, whereas at lower temperatures, novel electronic states responsible for visible light activity located near valence band of titania originating from the presence of poly(amino-s-triazine) shell are formed [6].
Herein, we demonstrate that titania with a polytriazine layer originated from urea, and modified with Cu2O (Cu2O/PTr–TiO2) becomes an efficient visible light photocatalyst, significantly more active than that of single modified titania, i.e., PTr–TiO2 or Cu2O/TiO2. The preparation method is economically and practically attractive by using low cost components such as urea and Cu2O and simple synthetic operations, i.e., lower temperatures of preparation in comparison with g-C3N4 synthesis (400 v. 550 °C). The important issue is the probable proposition of the explanation of the synergistic behavior of two different modifiers with corresponding visible light activation mechanisms. There are only two studies describing combinations of g-C3N4 with Cu2O [35] and additionally with TiO2 [36] showing improvements of visible light photocatalytic activity in comparison to single components (However, dyes have been used as test compounds, and thus origin of visible activity could not be unequivocally decided, i.e., dye sensitization could not be excluded [37,38]). Moreover, polycondensed s-triazines of graphitic structure were used, but not the non-graphitic forms obtained at lower temperature as proposed by Mitoraj and Kisch [29]. To the best of our knowledge, this report is the first study for this type of photocatalytic material.

2. Results and Discussion

2.1. Preparation Conditions and Visible Light Photocatalytic Efficiency

Commercially available P25 titania was used as a base to prepare the final photocatalytic material (Cu2O/PTr–TiO2). The first step of preparation was to obtain PTr–TiO2 with urea as a modifier (preparation details in Materials and Methods). “PTr” symbolizes the polytriazine shell. Different ratios between urea and P25 were investigated: 0.5:1, 1:1, 2:1 and 3:1 to find the material with the best photocatalytic efficiency under visible light irradiation. Samples with different shades of yellow color were obtained (more intensive yellow represents a higher amount of urea). 2-propanol oxidation to acetone was selected as the reaction system to test visible light photocatalytic activity (λ > 455 nm). The sample prepared at 1:1 urea–P25 ratio possessed the highest photocatalytic activity (measured as the produced acetone amount) and was chosen for further research and marked as PTr–TiO2 (see Figure 1). It was found that higher amounts of urea were not advantageous for the photocatalytic activity of the final product probably due to the detrimental influence of some not converted products of urea thermal transformation (excess of urea) on the semiconducting character of the polytriazine shell structure, as previously reported [29].
The second step in the photocatalyst preparation was the physical mixing of PTr–TiO2 with Cu2O by using different contents of Cu2O. Figure 2 shows the evidence of the advantageous role of the addition of cuprous oxide to the PTr–TiO2 sample. The optimum Cu2O content was found to be in the range of ca. 5 wt %, whereas 10 wt % and larger amounts were detrimental for photocatalytic efficiency. Detrimental influence of modifiers at their larger contents on photocatalytic activity is not surprising, and has already been reported for various systems. There are two main reasons for this behavior, i.e., (i) a shielding effect as Cu2O blocks PTr–TiO2 for photon absorption; and (ii) competitive adsorption as the oxygen and/or organic compounds (here 2-propanol) could not adsorbed directly on the titania surface, occupied by its modifier. Based on the obtained results, 5 wt % was chosen the Cu2O content in the Cu2O/PTr–TiO2 sample.

2.2. Characterization of Samples

P25 is a well-known mixed-phase titania photocatalyst containing two crystalline phases: anatase (86.4%) and rutile (13.6%), and amorphous titania (exact composition using NiO as the internal standard was estimated previously showing the anatase/rutile/amorphous ratio to be 76–80/13–15/6–11 (P25 is not a uniform sample) [39]) with a 21 nm crystallite size of anatase (determined from XRD) and 59.1 m2/g specific surface area. After the thermal treatment with urea, the properties of the PTr–TiO2 sample (1:1 urea–P25 ratio) changed slightly to a 23 nm crystallite size of anatase and 62.3 m2 g−1 specific surface area, but the crystalline phase content remained unchanged. It is proposed that an increase in crystallite size should result from amorphous titania conversion to anatase during calcination, whereas an increase in specific surface area comes from the adsorbed organic layer of poly(amino-tri-s-triazine) derivatives. Cu2O, which was used as a component of the Cu2O–PTr–TiO2 mixture, was characterized by a 65 nm crystallite size and 23 m2 g−1 specific surface area. Therefore, the results stated that the PTr–TiO2 crystallites were almost three times smaller than the Cu2O ones. These values were confirmed by XRD analysis of the Cu2O/PTr–TiO2 sample. Figure 3 shows the XRD diffractogram of Cu2O/PTr–TiO2 with peaks corresponding to anatase, rutile, and Cu2O. An aromatic system of carbon nitrides should appear at around 27.4°, but a strong peak of rutile overlaps it. It was reported that in the case of a small fraction of polycondensed polytriazines, this peak would not be detected [29].
The diffuse reflectance spectra of the PTr–TiO2 and Cu2O/PTr–TiO2 samples (Figure 4) show the strong red shift originated from the absorption properties of the polytriazine layer on P25. The visible light absorption effect was stronger than that for similar samples reported by Mitoraj and Kisch [28,30]. However, a different type of titania was used as the base for synthesis in that study. Therefore, it was proposed that the content of the surface hydroxyl groups on the surfaces of the various TiO2 samples were responsible for better/worse ability of polytriazine layer formation (the issue of finishing the process of formation polytriazine layer by the reaction of amino groups of the relevant intermediates with titania surface OH groups [30]). To check this hypothesis, reference experiments were performed for own-prepared decahedral anatase particles (DAP) [40] modified with urea using the same conditions as for P25. Interestingly, it was observed that the DAP sample after this modification practically did not change, i.e., the color was still white and the visible light activity was not observed. Accordingly, it is proposed that the highly crystalline titania samples with low surface area (DAP as an example) possess too low number of surface OH groups to successfully introduce the polytriazine layer on their surfaces. It is also possible that surface defects, e.g., oxygen vacancy, play a crucial role in polytriazine layer formation since the DAP (well-crystallized faceted titania) possess an extremely low content of such defects (clarification of this phenomenon for other titania samples is under study). Mitoraj and Kisch proposed that the visible light absorption of urea-modified samples with a yellow color originated from the presence of poly(tri-s-triazine) centered intra-bandgap levels which may form (depending on PTr concentration) a narrow energy band in titania [29]. Additionally, in the case of the Cu2O/PTr–TiO2 photocatalyst, more intensive light absorption in the range 450–600 nm, as the consequence of the inherent light absorption properties of Cu2O, was observed (Figure 3). The Cu2O/PTr–TiO2 sample had a reddish-yellow color.
XPS analysis of PTr–TiO2 and Cu2O/PTr–TiO2 samples showed the presence of nitrogen 1s with binding energies of 399.1 and 400.5 eV, which indicates the presence of C=N–C and (–C=N–)x bonds, respectively, being in agreement with previously reported values [29,41,42]. XPS results showing the fractions of oxidation states of titanium, oxygen and copper are shown in Table 1. For samples modified with urea (PTr–TiO2 and Cu2O/PTr–TiO2), the content of oxygen related to hydroxyl groups bound to titanium and carbon was lower than for bare TiO2 due to the fact that surface hydroxyl groups participate in poly-s-triazine layer formation. In contrast, the content of Ti3+ increased after modification, which could confirm the participation of surface defects in the formation of PTr–TiO2 (as discussed above for modified faceted titania (DAP)). It is proposed that thermal treatment could increase the formation of Ti3+. Therefore, it is also possible that a decrease in the content of hydroxyl groups after titania modification could result from their replacement by poly-s-triazine, similar to titania surface modification with nanoparticles of noble metals [43]. Moreover, samples with the poly-s-triazine layer also had a higher carbon content (higher C/Ti ratio). These results confirmed the presence of an organic layer on the titania surface consisting of poly-s-triazine derivatives, as reported previously [29].
To confirm the presence of the organic layer and possible formation of the Cu2O/PTr–TiO2 heterojunction, scanning transmission electron microscopy (STEM) observation was carried out. It was found that Cu2O existed in both crystalline and amorphous forms (large aggregates of very fine NPs). Moreover, two kinds of crystalline structures were noticed, i.e., large crystals of 100–150 nm and fine faceted nanocrystals of 10–30 nm (cubic and decahedral). For the Cu2O/PTr-TiO2 sample, titania particles were covered with a 5–10-nm layer of polytriazine located nearby larger Cu2O crystallites, as clearly shown in Figure 5 (respective SEM and TEM modes of the same image). Moreover, fine amorphous Cu2O interconnecting PTr–TiO2 particles was also observed. Interestingly, it was found that long-term electron beam during STEM observations could destroy the polytriazine layer around the titania since new nanostructures were formed (gravitationally-formed honey-like tails).

2.3. Enhanced Visible Light Photocatalytic Activity as a Synergistic Effect of Two Modifiers

Figure 6 shows the results of the photocatalytic activity tests of five photocatalysts: TiO2, Cu2O, Cu2O/TiO2, PTr-TiO2, and Cu2O/PTr–TiO2 in the visible light-induced 2-propanol oxidation. The highest activity was observed for Cu2O/PTr-TiO2 with 1.2 μmol of acetone formation after 3 h of irradiation. The Cu2O/TiO2 and PTr–TiO2 samples were characterized by significantly lower photoactivity: 0.18 and 0.72 μmol of acetone, respectively. The unmodified photocatalysts TiO2 (P25) and Cu2O were almost inactive in this reaction system. The improvement of visible light photocatalytic performance for the Cu2O/PTr–TiO2 sample in comparison to the Cu2O/TiO2 and PTr–TiO2 photocatalysts is evidence of the synergistic effect of urea (poly(tri-s-triazine)) and Cu2O modifiers for enhancing the visible light photocatalytic activity of titania. The photocatalytic activity test for the Cu2O/PTr–TiO2 sample was extended to 8 h to check the basic stability of the material. After 8 h of irradiation, a linear course of acetone production was still observed.
A lack of visible light activity for a single Cu2O, in spite of its good visible light absorption properties, can be explained by a strong photocorrosion effect and fast charge recombination [44,45]. It was reported that the coupling of cuprous oxide with titania reduced these detrimental effects by the heterojunction mechanism (type II) between the p- and n-type semiconductors introducing visible light activity and enhancing stability [15,16,17,18,19,46].
The PTr–TiO2 photocatalyst possesses significantly higher visible light activity for 2-propanol oxidation than Cu2O/TiO2. The core-shell structure of titania–poly(tri-s-triazine) derivatives changes photoabsorption and electronic properties in comparison to the unmodified titania. Figure 7 shows a way to determine band gap energy and absorption onset for the PTr–TiO2 sample. The band gap narrowing occurred from 3.17 eV for bare TiO2 (P25) to 2.88 eV for PTr-TiO2. The absorption onset is useful to determine the location of surface states in PTr–TiO2 originating from the presence of an organic modifier (shaded area in the Figure 8 and Figure 9) [6].
Figure 8 and Figure 9 illustrate the propositions of explanation for enhanced visible light photocatalytic activity of Cu2O/PTr–TiO2, a photocatalytic system with two main composites. In the first mechanistic variant (Figure 7), two semiconductors: p-type Cu2O and n-type titania (P25) with an organic sensitizer created a type II heterojunction. This system provided the optimum band positions for efficient charge carrier separation. Visible light photoexcited electrons were transferred from CB(Cu2O) to CB(PTr–TiO2) and this transfer could occur due to the favorable energetics of the relative positions of both CBs, whereas holes were simultaneously transferred from VB(PTr–TiO2) to VB(Cu2O). The main consequence of such phenomenon is the separation of the photogenerated electrons and holes reducing the probability of recombination and increasing the lifetimes of the charge carriers [46]. Cu2O has a valence band potential of 1.07 V [47]. To form hydroxyl radicals, the valence band potential of Cu2O should be more positive than the following values at pH = 7 [7]:
h+ + H2O → OH + H+, E0 = 2.73 V (vs. NHE)
h+ + OH → OH, E0 = 1.90 V (vs. NHE)
Therefore, the formation of hydroxyl radicals may be thermodynamically unfavorable. It can be concluded that holes accumulated would be mostly consumed through the direct oxidation of 2-propanol. Electrons accumulated on CB(PTr–TiO2) play a key role in the formation of reactive oxygen species. The properties of the PTr–TiO2 component of the considered semiconductor composition by introduction of visible light activity into titania enabled the formation of this advantageous heterojunction functioning under visible light irradiation.
Another mechanistic variant describing enhanced visible light activated Cu2O/PTr–TiO2 system relies on the Z-scheme concept [48,49]. As shown in Figure 9, photogenerated electrons in PTr–TiO2 with a lower reduction ability recombined with the photogenerated holes in Cu2O with lower oxidation ability. Therefore, electrons accumulated on CB(Cu2O) with a high reduction ability of holes accumulated on VB(PTr–TiO2) with a high oxidation ability can be maintained. The occurrence of the second mechanistic variant might be even more possible due to the existence of optimal Cu2O content (Figure 2) and the low photocatalytic activity of the Cu2O/TiO2 system.
In the case of both mechanisms, the efficient charge separation induced by visible light irradiation was the main reason of the exceptional photocatalytic activity of the proposed system and provides the explanation of the synergistic role of the two considered types of titania modifiers. Clarification of which mechanistic variant is responsible for enhanced photocatalytic properties of the Cu2O/PTr–TiO2 system in the visible light is necessary. Further studies including action spectrum analysis, photoactivity tests in the presence of scavengers, and detailed characterization of these materials (and other photocatalysts prepared with different titania photocatalysts, i.e., varied by surface properties and/or content of oxygen defects), such as the estimation of the quasi–Fermi level, are along this line.

3. Materials and Methods

3.1. Preparation of Cu2O/PTr–TiO2

P25 (AEROXIDE® TiO2 P25, Nippon Aerosil, Tokyo, Japan), urea (Wako Pure Chemicals, Osaka, Japan) and Cu2O (Wako Pure Chemicals) were used for the study without purification. PTr–TiO2 samples were prepared by the method based on the processes reported elsewhere [26,29]. Typically, 400 g of P25 powder and different amounts of urea corresponding to 0.5:1, 1:1, 2:1 and 3:1 (w/w) ratios were ground in an agate mortar, followed by calcination in air at 400 °C for 30 min. Powders were placed in the open test-tube with 15 cm of length. The resulting powders were washed five times with water to remove the excess of urea decomposition products and finally dried under air at 70 °C. In the second step, PTr–TiO2 sample and Cu2O powder (in different amounts: 1, 2, 5, 7, 10 wt %) were mixed thoroughly with 5 min of grinding.

3.2. Sample Characterization

The UV–Vis diffuse reflectance spectra (DRS) were recorded on JASCO V-670 (JASCO, Tokyo, Japan) equipped with PIN-757 integrating sphere using BaSO4 as a reference. Gas-adsorption measurements of prepared titania samples were performed on a Yuasa Ionics Autosorb, 6AG (Yuasa Ionics, Osaka, Japan) surface area and pore size analyzer. Specific surface area (SSA) was calculated from nitrogen adsorption at 77 K using the Brunauer–Emmett–Teller equation. X-ray diffraction patterns (XRD) were collected using an X-ray diffractometer (Rigaku intelligent XRD SmartLab with a Cu target, Rigaku, Tokyo, Japan). X-ray photoelectron spectra (XPS) were recorded using a JEOL JPC-9010MC (JEOL, Tokyo, Japan) spectrometer with a MgKa X-ray source. Samples were also characterized by scanning transmission electron microscopy (STEM, HITACHI HD-2000, HITACHI, Tokyo, Japan).

3.3. Photocatalytic Reaction

The photocatalyst (50 mg) was suspended in an aqueous solution of 2-propanol (5 vol %, 5 mL) and photoirradiated (120 W-xenon lamp) with a Y48 cut-off filter mounted in the irradiation window, therefore the light of wavelengths >450 nm reached the suspension, which was under continuous magnetic stirring (1000 rpm) in a thermostated bath. Generated acetone was detected using GC–FID (Shimadzu GC-14B (Shimadzu, Kyoto, Japan) equipped with a flame ionization detector). Before the injection of the liquid sample to GC, the photocatalyst was separated using a filter (Whatman Mini-UniPrep, PVDF, Whatman, Maidstone, UK).

4. Conclusions

The results presented in this study clearly revealed that the application of both low cost- modifiers for titania: urea (first step of preparation) and Cu2O (second step), significantly enhanced the visible light photocatalytic properties of TiO2 in comparison to the single materials of urea-modified TiO2 and Cu2O-modified titania. Therefore, it is possible to describe this phenomenon as a synergistic effect of urea (more detailed: the presence of polytriazine layer on titania originated from the thermal treatment of urea) and Cu2O. The type II heterojunction or Z-scheme systems formed by two semiconductors, p-type Cu2O and n-type PTr–TiO2, can be responsible for the improvement of photocatalytic activity through the intensification of visible light-induced charge separation and subsequent reduction of the electron-hole recombination effect. A preparation of composite photocatalysts based on titania and low-cost materials as modifiers, such as a precursor of the organic sensitizer of titania and metal oxide, to prepare the efficient and stable photocatalytic system operating in the visible light within heterojunction principles is a very promising direction for a wider practical application of photocatalysis given the preference for concomitantly cheap and efficient solutions.

Author Contributions

M.J. conceived, designed, performed the experiments and characterizations, interpreted the data, and wrote the paper. E.K. performed STEM, interpreted the data, and corrected the manuscript. K.W. performed STEM experiments. All authors read and approved the final manuscript.

Acknowledgments

M.J. acknowledges Hokkaido University for guest lecturer position (2016–2017).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bahnemann, D.W. Photocatalytic water treatment: Solar energy applications. Solar Energy 2004, 77, 445–459. [Google Scholar] [CrossRef]
  2. Chen, X.B.; Shen, S.H.; Guo, L.J.; Mao, S.S. Semiconductor-based photocatalytic hydrogen generation. Chem. Rev. 2010, 110, 6503–6570. [Google Scholar] [CrossRef] [PubMed]
  3. Li, K.; Peng, B.S.; Peng, T.Y. Recent advances in heterogeneous photocatalytic CO2 conversion to solar fuels. ACS Catal. 2016, 6, 7485–7527. [Google Scholar] [CrossRef]
  4. Pietron, J.J.; DeSario, P.A. Review of roles for photonic crystals in solar fuels photocatalysis. J. Photonics Energy 2016, 7, 012007. [Google Scholar] [CrossRef] [Green Version]
  5. Ragesh, P.; Ganesh, V.A.; Nair, S.V.; Nair, A.S. A review on ‘self-cleaning and multifunctional materials’. J. Mater. Chem. A 2014, 2, 14773–14797. [Google Scholar] [CrossRef]
  6. Kisch, H. Semiconductor photocatalysis—Mechanistic and synthetic aspects. Angew. Chem. Int. Ed. 2013, 52, 812–847. [Google Scholar] [CrossRef] [PubMed]
  7. Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.; Horiuchi, Y.; Anpo, M.; Bahnemann, D.W. Understanding TiO2 photocatalysis: Mechanisms and materials. Chem. Rev. 2014, 114, 9919–9986. [Google Scholar] [CrossRef] [PubMed]
  8. Pelaez, M.; Nolan, N.T.; Pillai, S.C.; Seery, M.K.; Falaras, P.; Kontos, A.G.; Dunlop, P.S.M.; Hamilton, J.W.J.; Byrne, J.A.; O’Shea, K.; et al. A review on the visible light active titanium dioxide photocatalysts for environmental applications. Appl. Catal. B Environ. 2012, 125, 331–349. [Google Scholar] [CrossRef] [Green Version]
  9. Dahl, M.; Liu, Y.; Yin, Y. Composite titanium dioxide nanomaterials. Chem. Rev. 2014, 114, 9853–9889. [Google Scholar] [CrossRef] [PubMed]
  10. Bhanushali, S.; Ghosh, P.; Ganesh, A.; Cheng, W.L. 1D copper nanostructures: Progress, challenges and opportunities. Small 2015, 11, 1232–1252. [Google Scholar] [CrossRef] [PubMed]
  11. Clarizia, L.; Spasiano, D.; Di Somma, I.; Marotta, R.; Andreozzi, R.; Dionysiou, D.D. Copper modified-TiO2 catalysts for hydrogen generation through photoreforming of organics. A short review. Int. J. Hydrogen Energy 2014, 39, 16812–16831. [Google Scholar] [CrossRef]
  12. 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]
  13. Nguyen, M.A.; Bedford, N.M.; Ren, Y.; Zahran, E.M.; Goodin, R.C.; Chagani, F.F.; Bachas, L.G.; Knecht, M.R. Direct Synthetic Control over the Size, Composition, and Photocatalytic Activity of Octahedral Copper Oxide Materials: Correlation Between Surface Structure and Catalytic Functionality. ACS Appl. Mater. Interfaces 2015, 7, 13238–13250. [Google Scholar] [CrossRef] [PubMed]
  14. Deng, X.L.; Wang, C.G.; Shao, M.H.; Xu, X.J.; Huang, J.Z. Low-temperature solution synthesis of CuO/Cu2O nanostructures for enhanced photocatalytic activity with added H2O2: Synergistic effect and mechanism insight. RSC Adv. 2017, 7, 4329–4338. [Google Scholar] [CrossRef]
  15. Yang, L.; Luo, S.; Li, Y.; Xiao, Y.; Kang, Q.; Cai, Q. High efficient photocatalytic degradation of p-nitrophenol on a unique Cu2O/TiO2 p-n heterojunction network catalyst. Environ. Sci. Technol. 2010, 44, 7641–7646. [Google Scholar] [CrossRef] [PubMed]
  16. Chu, S.; Zheng, X.M.; Kong, F.; Wu, G.H.; Luo, L.L.; Guo, Y.; Liu, H.L.; Wang, Y.; Yu, H.X.; Zou, Z.G. Architecture of Cu2O@TiO2 core-shell heterojunction and photodegradation for 4-nitrophenol under simulated sunlight irradiation. Mater. Chem. Phys. 2011, 129, 1184–1188. [Google Scholar] [CrossRef]
  17. Liu, L.; Gu, X.; Sun, C.; Li, H.; Deng, Y.; Gao, F.; Dong, L. In situ loading of ultra-small Cu2O particles on TiO2 nanosheets to enhance the visible-light photoactivity. Nanoscale 2012, 4, 6351–6359. [Google Scholar] [CrossRef] [PubMed]
  18. Liu, L.M.; Yang, W.Y.; Li, Q.; Gao, S.A.; Shang, J.K. Synthesis of Cu2O nanospheres decorated with TiO2 nanoislands, their enhanced photoactivity and stability under visible light illumination, and their post-illumination catalytic memory. ACS Appl. Mater. Interfaces 2014, 6, 5629–5639. [Google Scholar] [CrossRef] [PubMed]
  19. Xiong, L.B.; Yang, F.; Yan, L.L.; Yan, N.N.; Yang, X.; Qiu, M.Q.; Yu, Y. Bifunctional photocatalysis of TiO2/Cu2O composite under visible light: Ti3+ in organic pollutant degradation and water splitting. J. Phys. Chem. Solids. 2011, 72, 1104–1109. [Google Scholar] [CrossRef]
  20. Qiu, X.; Miyauchi, M.; Sunada, K.; Minoshima, M.; Liu, M.; Lu, Y.; Li, D.; Shimodaira, Y.; Hosogi, Y.; Kuroda, Y.; et al. Hybrid CuxO/TiO2 nanocomposites as risk-reduction materials in indoor environments. ACS Nano 2012, 6, 1609–1618. [Google Scholar] [CrossRef] [PubMed]
  21. Hans, M.; Erbe, A.; Mathews, S.; Chen, Y.; Solioz, M.; Mucklich, F. Role of copper oxides in contact killing of bacteria. Langmuir 2013, 29, 16160–16166. [Google Scholar] [CrossRef] [PubMed]
  22. Duan, W.; Zheng, M.; Li, R.; Wang, Y. Morphology transformation of Cu2O by adding TEOA and their antibacterial activity. J. Nanopart. Res. 2016, 18, 342. [Google Scholar] [CrossRef]
  23. Lee, Y.J.; Kim, S.; Park, S.H.; Park, H.; Huh, Y.D. Morphology-dependent antibacterial activities of Cu2O. Mater. Lett. 2011, 65, 818–820. [Google Scholar] [CrossRef]
  24. Nosaka, Y.; Matsushita, M.; Nishino, J.; Nosaka, A.Y. Nitrogen-doped titanium dioxide photocatalysts for visible response prepared by using organic compounds. Sci. Technol. Adv. Mater. 2005, 6, 143–148. [Google Scholar] [CrossRef] [Green Version]
  25. Bacsa, R.; Kiwi, J.; Ohno, T.; Albers, P.; Nadtochenko, V. Preparation, testing and characterization of doped TiO2 active in the peroxidation of biomolecules under visible light. J. Phys. Chem. B 2005, 109, 5994–6003. [Google Scholar] [CrossRef] [PubMed]
  26. Kisch, H.; Sakthivel, S.; Janczarek, M.; Mitoraj, D. A low-band gap, nitrogen-modified titania visible-light photocatalyst. J. Phys. Chem. C 2007, 111, 11445–11449. [Google Scholar] [CrossRef]
  27. Beranek, R.; Neumann, B.; Sakthivel, S.; Janczarek, M.; Dittrich, T.; Tributsch, H.; Kisch, H. Exploring the electronic structure of nitrogen-modified TiO2 photocatalysts through photocurrent and surface photovoltage studies. Chem. Phys. 2007, 339, 11–19. [Google Scholar] [CrossRef]
  28. Mitoraj, D.; Kisch, H. The nature of nitrogen-modified titanium dioxide photocatalysts active in visible light. Angew. Chem. Int. Ed. 2008, 47, 9975–9978. [Google Scholar] [CrossRef] [PubMed]
  29. Mitoraj, D.; Kisch, H. On the mechanism of urea-Induced titania modification. Chem. Eur. J. 2010, 16, 261–269. [Google Scholar] [CrossRef] [PubMed]
  30. Mitoraj, D.; Kisch, H. Surface modified titania visible light photocatalyst powders. Solid State Phenom. 2010, 162, 49–75. [Google Scholar] [CrossRef]
  31. Dong, G.; Zhang, Y.; Pan, Q.; Qiu, J. A fantastic graphitic carbon nitride (g-C3N4) material: Electronic structure, photocatalytic and photoelectronic properties. J. Photochem. Photobiol. C Photochem. Rev. 2014, 20, 33–50. [Google Scholar] [CrossRef]
  32. Liu, J.; Zhang, T.; Wang, Z.; Dawson, G.; Chen, W. Simple pyrolysis of urea into graphitic carbon nitride with recyclable adsorption and photocatalytic activity. J. Mater. Chem. 2011, 21, 14398–14401. [Google Scholar] [CrossRef]
  33. Lee, S.C.; Lintang, H.O.; Yuliati, L. A urea precursor to synthesize carbon nitride with mesoporosity for enhanced activity in the photocatalytic removal of phenol. Chem. Asian J. 2012, 7, 2139–2144. [Google Scholar] [CrossRef] [PubMed]
  34. Senthil, R.A.; Theerthagiri, J.; Selvi, A.; Madhavan, J. Synthesis and characterization of low-cost g-C3N4/TiO2 composite with enhanced photocatalytic performance under visible-light rradiation. Opt. Mater. 2017, 64, 533–539. [Google Scholar] [CrossRef]
  35. Peng, B.; Zhang, S.; Yang, S.; Wang, H.; Yu, H.; Zhang, S.; Peng, F. Synthesis and characterization of g-C3N4/Cu2O composite catalyst with enhanced photocatalytic activity under visible light irradiation. Mater. Res. Bull. 2014, 56, 19–24. [Google Scholar] [CrossRef]
  36. Min, Z.; Wang, X.; Li, Y.; Jiang, J.; Li, J.; Qian, D.; Li, J. A highly efficient visible-light-responding Cu2O-TiO2/g-C3N4 photocatalyst for instantaneous discolorations of organic dyes. Mater. Lett. 2017, 193, 18–21. [Google Scholar] [CrossRef]
  37. Kisch, H.; Macyk, W. Visible-light photocatalysis by modified titania. Chem. Phys. Chem. 2002, 3, 399–400. [Google Scholar] [CrossRef]
  38. Yan, X.; Ohno, T.; Nishijima, K.; Abe, R.; Ohtani, B. Is methylene blue an appropriate substrate for a photocatalytic activity test? A study with visible-light responsive titania. Chem. Phys. Lett. 2006, 429, 606–610. [Google Scholar] [CrossRef] [Green Version]
  39. Wang, K.; Wei, Z.; Ohtani, B.; Kowalska, E. Interparticle electron transfer in methanol dehydrogenation on platinum-loaded titania particles prepared from P25. Catal. Today 2018, 303, 327–333. [Google Scholar] [CrossRef]
  40. Janczarek, M.; Kowalska, E.; Ohtani, B. Decahedral-shaped anatase titania photocatalyst particles: Synthesis in a newly developed coaxial-flow gas-phase reactor. Chem. Eng. J. 2016, 289, 502–512. [Google Scholar] [CrossRef] [Green Version]
  41. Dementjev, A.P.; de Graaf, A.; van den Sanden, M.C.M.; Maslakov, K.I.; Naumkin, A.V.; Serov, A.A. X-Ray photoelectron spectroscopy reference data for identification of the C3N4 phase in carbon–nitrogen films. Diam. Relat. Mater. 2000, 9, 1904–1907. [Google Scholar] [CrossRef]
  42. Guo, X.; Xie, Y.; Wang, X.; Zhang, S.; Hou, T.; Lv, S. Synthesis of carbon nitride nanotubes with the C3N4 stoichiometry via a benzene-thermal process at low temperatures. Chem. Commun. 2004, 26–27. [Google Scholar] [CrossRef] [PubMed]
  43. Janczarek, M.; Wei, Z.; Endo, M.; Ohtani, B.; Kowalska, E. Silver- and copper-modified decahedral anatase titania particles as visible light-responsive plasmonic photocatalyst. J. Photonics Energy 2017, 7, 012008. [Google Scholar] [CrossRef]
  44. Bessekhouad, Y.; Robert, D.; Weber, J.-V. Photocatalytic activity of Cu2O/TiO2, Bi2O3/TiO2 and ZnMn2O4/TiO2 heterojunctions. Catal. Today 2005, 101, 315–321. [Google Scholar] [CrossRef]
  45. Huang, L.; Peng, F.; Yu, H.; Wang, H. Preparation of cuprous oxides with different sizes and their behaviors of adsorption, visible-light driven photocatalysis and photocorrosion. Solid State Sci. 2009, 11, 129–138. [Google Scholar] [CrossRef]
  46. Marschall, R. Semiconductor Composites: Strategies for Enhancing Charge Carrier Separation to Improve Photocatalytic Activity. Adv. Funct. Mater. 2014, 24, 2421–2440. [Google Scholar] [CrossRef]
  47. Luna, A.L.; Valenzuela, M.A.; Colbeau-Justin, C.; Vazquez, P.; Rodriguez, J.; Avendano, J.R.; Alfaro, S.; Tirado, S.; Garduno, A.; De la Rosa, J.M. Photocatalytic degradation of gallic acid over CuO–TiO2 composites under UV/Vis LEDs irradiation. Appl. Catal. A Gen. 2016, 521, 140–148. [Google Scholar] [CrossRef]
  48. Li, H.; Tu, W.; Zhou, Y.; Zou, Z. Z-scheme photocatalytic systems for promoting photocatalytic performance: Recent progress and future challenges. Adv. Sci. 2016, 3, 1500389. [Google Scholar] [CrossRef] [PubMed]
  49. Low, J.; Jiang, C.; Cheng, B.; Wageh, S.; Al-Ghamdi, A.A.; Yu, J. A review of direct Z-scheme photocatalysts. Small Methods 2017, 1, 1700080. [Google Scholar] [CrossRef]
Figure 1. Visible light photocatalytic activity of the samples prepared with corresponding urea/P25 ratio.
Figure 1. Visible light photocatalytic activity of the samples prepared with corresponding urea/P25 ratio.
Catalysts 08 00240 g001
Figure 2. Visible light photocatalytic activity of the samples prepared with different Cu2O content.
Figure 2. Visible light photocatalytic activity of the samples prepared with different Cu2O content.
Catalysts 08 00240 g002
Figure 3. XRD diffractogram of Cu2O/PTr–TiO2 sample.
Figure 3. XRD diffractogram of Cu2O/PTr–TiO2 sample.
Catalysts 08 00240 g003
Figure 4. Diffuse reflectance spectra of TiO2, PTr–TiO2, Cu2O/PTr–TiO2. The Kubelka–Munk function F(R) is equivalent to absorbance.
Figure 4. Diffuse reflectance spectra of TiO2, PTr–TiO2, Cu2O/PTr–TiO2. The Kubelka–Munk function F(R) is equivalent to absorbance.
Catalysts 08 00240 g004
Figure 5. STEM images of Cu2O/PTr–TiO2 sample in SEM (a) and TEM (b) modes.
Figure 5. STEM images of Cu2O/PTr–TiO2 sample in SEM (a) and TEM (b) modes.
Catalysts 08 00240 g005
Figure 6. Visible light-induced (λ > 455 nm) 2-propanol oxidation in the presence of (■) TiO2, () Cu2O, () Cu2O/TiO2, () PTr-TiO2, and () Cu2O/PTr-TiO2.
Figure 6. Visible light-induced (λ > 455 nm) 2-propanol oxidation in the presence of (■) TiO2, () Cu2O, () Cu2O/TiO2, () PTr-TiO2, and () Cu2O/PTr-TiO2.
Catalysts 08 00240 g006
Figure 7. Plot of transformed Kubelka–Munk function vs. energy of light for the PTr–TiO2 sample. Determination of band gap energy (a) and absorption onset (b).
Figure 7. Plot of transformed Kubelka–Munk function vs. energy of light for the PTr–TiO2 sample. Determination of band gap energy (a) and absorption onset (b).
Catalysts 08 00240 g007
Figure 8. Energy diagram for the Cu2O/PTr–TiO2 photocatalytic system working under visible light, illustrating the coupling of two semiconductors as a type II heterojunction.
Figure 8. Energy diagram for the Cu2O/PTr–TiO2 photocatalytic system working under visible light, illustrating the coupling of two semiconductors as a type II heterojunction.
Catalysts 08 00240 g008
Figure 9. Energy diagram for Cu2O/PTr–TiO2 photocatalytic system working under visible light, illustrating the coupling of two semiconductors as a Z-scheme system.
Figure 9. Energy diagram for Cu2O/PTr–TiO2 photocatalytic system working under visible light, illustrating the coupling of two semiconductors as a Z-scheme system.
Catalysts 08 00240 g009
Table 1. XPS analysis for TiO2, PTr–TiO2 and Cu2O/PTR–TiO2 samples including fraction of oxidation states of Ti, O, and Cu from the deconvolution of XPS peaks of Ti 2p3/2, O 1s and Cu 2p3/2.
Table 1. XPS analysis for TiO2, PTr–TiO2 and Cu2O/PTR–TiO2 samples including fraction of oxidation states of Ti, O, and Cu from the deconvolution of XPS peaks of Ti 2p3/2, O 1s and Cu 2p3/2.
SamplesTi 2p3/2 (%)O 1s (%)RatioValent State (%)
Ti4+Ti3+TiO2 aTi-OH bTi-OH cO/TiC/TiCu2+Cu+Cu(0)
TiO298.51.557.926.715.42.63.5
PTr-TiO297.22.862.323.714.02.54.1
Cu2O/PTr-TiO297.03.061.824.114.12.54.30.999.1
a Oxygen in the TiO2 crystal lattice; b Ti–(OH)–Ti, Ti2O3, C=O; c Ti–OH, C–OH.

Share and Cite

MDPI and ACS Style

Janczarek, M.; Wang, K.; Kowalska, E. Synergistic Effect of Cu2O and Urea as Modifiers of TiO2 for Enhanced Visible Light Activity. Catalysts 2018, 8, 240. https://doi.org/10.3390/catal8060240

AMA Style

Janczarek M, Wang K, Kowalska E. Synergistic Effect of Cu2O and Urea as Modifiers of TiO2 for Enhanced Visible Light Activity. Catalysts. 2018; 8(6):240. https://doi.org/10.3390/catal8060240

Chicago/Turabian Style

Janczarek, Marcin, Kunlei Wang, and Ewa Kowalska. 2018. "Synergistic Effect of Cu2O and Urea as Modifiers of TiO2 for Enhanced Visible Light Activity" Catalysts 8, no. 6: 240. https://doi.org/10.3390/catal8060240

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