Adnan
Younis
*,
Dewei
Chu
*,
Yusuf Valentino
Kaneti
and
Sean
Li
School of Materials Science and Engineering, University of New South Wales, Sydney, 2052, NSW, Australia. E-mail: a.younis@unsw.edu.au; d.chu@unsw.edu.au; Fax: +61 (0)2 9385 6565; Tel: +61 (0)2 9385 0763
First published on 16th November 2015
For oxide semiconductors, the morphology, particle size and oxygen vacancies are usually considered as key influential parameters for photocatalytic degradation of organic pollutants/dyes. It is widely accepted that cation doping not only modifies their phase and microstructures but also introduces variations in oxygen vacancy concentration. Herein, we report the fabrication of sub-10 nm sized pure and indium doped CeO2 nanocrystals (NCs) via a facile, green hydrothermal method for the investigation of photocatalytic activities. X-ray diffraction and transmission electron microscopy were employed to examine the crystal phase and morphology of the as-prepared nanocrystals. Raman and X-ray photoelectron spectroscopy techniques were implemented to investigate the presence and variations in oxygen vacancy concentration in un-doped and indium doped CeO2 nanocrystals. The photocatalytic activity results revealed that 10 at% doping is the optimal indium doping level to demonstrate superior dye removal efficiency (∼40%) over un-doped and doped CeO2 NCs. Moreover, the 10% In-doped CeO2 nanocrystals expressed excellent cycling stability and superior photocatalytic performance toward other dye pollutants. Finally, on the basis of our findings, a possible photocatalytic mechanism in which indium doping can generate more surface oxygen vacancies in the ceria lattice which delay the electron–hole recombination rates, thus increasing the lifetime of electron–hole separation for enhanced photocatalytic performances was proposed.
Until now, the majority of photocatalytic studies have been mainly focused on titanium dioxide (TiO2) with comparatively few reports on other oxide semiconductors. To this end, the quest of a diverse set of high-performance photocatalysts is beneficial for both fundamental and applied photocatalyses. In particular, CeO2 (ceria) is a unique catalyst which can take and release oxygen through reversible shifts between Ce3+ and Ce4+. In fact, the actual importance of oxygen uptake/release is determined not only by the OSC, but also by the rates of the redox cycles. Usually, the oxidation rate of cerium dominates over its reduction rate; therefore various efforts have been made to promote its reducibility by adjusting the nature and density of oxygen vacancies.3
The performance of CeO2 nanoparticles/nanocrystals strongly depends on the morphology and the particle size distribution. Usually, the CeO2 (100) crystal face has the highest surface energy and reactivity when compared to its low-index crystal planes (111) and (110). This difference in surface energies originates from the instability of the top-layer oxygen and is located at the bridging positions between two cerium ions. Therefore, CeO2 morphologies exposed by the (100) face usually demonstrate superior catalytic and oxygen storage capabilities.4,5 Although, the (100) face has superior photocatalytic properties by having higher surface energy, the long term stability of the (100) surface remains a challenge. Therefore, it would be of great interest if the photocatalytic performance of the most stable (111) plane of CeO2 nanocrystals can be improved. On the other hand, the exact surface structures of nano-scaled CeO2 still remain less explored and there are only a few reports on the well-defined (111) CeO2 surface for excellent photocatalytic applications. Doping transitional metal ions into CeO2 has been considered as an effective approach to modify its physical and chemical properties. In particular, indium (In) is considered as an efficient metal to produce a large number of electrons due to vacant d-orbitals and has the ability to hinder photo-generated charges. In3+ also has a comparable ionic radius with Ce3+/4+, so their doping levels in ceria can be easily controlled. Furthermore, the doping of indium can introduce more oxygen vacancies into CeO2, which can act as reaction sites to improve photocatalytic properties. The doping of indium can also narrow the bandgap of CeO2, which is beneficial for visible light photocatalysis. Owing to the aforementioned advantages, it would be of great interest to explore the role of In incorporation into the CeO2 matrix to study the variations in oxygen vacancy concentration and eventually their photocatalytic activities.
Herein, we have been able to prepare high-surface-area, thermally stable and monodisperse nanostructured CeO2 nanocrystals with varying In doping concentrations (0 at%, 5 at%, 10 at% and 15 at%) by self-assembly of individual nanoparticles in a liquid crystal phase. The effect of In doping on the crystal structure, morphologies and oxygen vacancy concentration was investigated. Moreover, we researched the photocatalytic activities of all the products under the irradiation of ultraviolet light. To the best of our knowledge, the present report is a first attempt towards the study of indium (In) doping on the catalytic activities of high surface area occupied CeO2 nanocrystals.
Transmission electron microscopy (TEM) was utilized to determine the morphologies of the as-prepared samples. From Fig. 2a–d, it can be seen that the CeO2 nanocrystals are well separated by forming a monodisperse array. All the samples preserved the nano-cubic type morphology with a slight variation in d-spacing with increasing In doping. The lattice fringes in the insets of Fig. 2a–d are clearly visible with a d-spacing of 0.314, 0.3114, 0.309 and 0.3084 nm for CS-0, CS-05, CS-10 and CS-15, respectively. These d-spacings are attributed to the (111) planes of ceria (JCPDS 34-0394) which match well with the XRD results.
It is generally believed that the interface and the grain boundaries are the active regions in CeO2 based materials for the formation of Ce3+ and oxygen vacancies.9 So, the local atomic structure and grain orientation near the grain boundaries determine the possibilities of defect formation near the grain boundary region. Careful examination of the HRTEM images revealed that the shape of CeO2 nanocrystals transformed from nearly perfect cubic to irregular/de-shaped cubic by increasing the In content which may generate line defects, containing oxygen vacancies as one of the point defects. This is an assumption that defects may be located in these regions. To understand the defect induced changes in the local structure of CeO2 nanocrystals and to characterize oxygen vacancies, Raman spectroscopy was carried out and the corresponding spectra are shown in Fig. 3. Usually bulk ceria has an intense Raman band at 464 cm−1, which is attributed to the Raman active vibrational mode (F2g) of the fluorite-type structure.10 All the samples show a strong band around this region, confirming the fluorite structure in all the un-doped and doped CeO2 samples. Another band at ∼595 cm−1 was found which can be related to the oxygen vacancies and/or point defects due to the presence of Ce3+ in the CeO2 lattice and defects caused by small size effects/point defects.11 By increasing the In concentration, the intensity of this band increases which may be due to the formation of Ce3+ ions in order to accommodate In3+ ions in the Ce host.12 Another band around ∼565 cm−1 was also found whose intensity increases with increasing In doping (CS-05, CS-10 and CS-15) which can be attributed to the extrinsic oxygen vacancies generated by the substitution of tetravalent Ce4+ with trivalent In3+.13 A number of previous studies14,15 have attributed the band gap narrowing of CeO2 as a result of additional states introduced within the band gap of CeO2. These additional states may have originated from the increased concentration of Ce3+ induced by the defect levels generated in the CeO2 lattice and oxygen vacancy defect states. On the basis of Raman spectroscopy, it can be concluded that indium doping can introduce more defects/oxygen vacancies into the CeO2 lattice. Therefore, it is suspected that the introduction of oxygen vacancies or their increased density at the surface of individual CeO2 nanocrystals can inhibit the recombination of electron–hole pairs to improve their performance as efficient photocatalysts.
The oxidation/valence state of the elements and the surface composition of CS-00 and CS-10 samples have been analysed by the XPS technique. The Ce 3d core-level X-ray photoelectron spectra consisting of peaks corresponding to the Ce3+ and Ce4+ states have been Gaussian fitted as shown in Fig. 4a and b. The percentage composition of Ce3+ and Ce4+ has been calculated using the following equations:
Ce3+ = v0 + v′ + u0 + u′ | (1) |
Ce4+ = v + v′′ + v′′′ + u + u′′ + u′′′ | (2) |
(3) |
Fig. 4 XPS core-level spectrum of (a, b) Ce 3d of CeO2 and In-doped CeO2. (c) O 1s of CeO2. (d) O 1s of In-doped CeO2 nanocrystals. |
To further strengthen the aforementioned arguments about oxygen vacancies, a series of surface depth profiling XPS experiments were conducted and the concentration of Ce3+ on the surface and at various depths below the surface was recorded. For these measurements, all the samples (CS-0, CS-05, CS-10 and CS-15) were dispersed in ethanol followed by drop-coating on silicon wafers to form their subsequent thin films. To conduct the surface profiling experiments, the thin film samples were etched by using an argon (Ar) ion beam with an accelerating voltage of 3 keV. The etching duration was varied from 0 seconds to 90 seconds, 180 seconds and finally to 270 seconds. The variations in the Ce3+/Ce4+ ratio at various sample depths extracted from XPS data were calculated and are plotted in Fig. S1.†
From Fig. S1,† it is clear that the concentration of Ce3+/Ce4+ (oxygen vacancies) on the surface of all the samples was much higher as compared to their bulk counterpart. A significant decrease in the concentration of oxygen vacancies was observed by increasing the etching time. This trend provides a strong evidence to support the presence of a higher concentration of oxygen vacancies on the surface of CeO2 samples. Furthermore, their concentration is relatively much lower in the bulk part of the sample. The Ce3+/Ce4+ ratio on the surface of all the samples (with no etching) is plotted separately in the inset of Fig. S1† for better understanding. The sample CS-15 shows the highest concentration of oxygen vacancies as compared to other samples and its Ce3+/Ce4+ ratio is almost double that of the CS-0 sample. Therefore on the basis of these results, it can be concluded that the incorporation of indium into the CeO2 lattice can promote the formation of more oxygen vacancies on the surface of CeO2.
It is a known fact that the electron–hole pair separation is closely related to the availability of oxygen vacancies in the catalyst. The In doping in CeO2 nanocrystals can induce/generate more oxygen vacancies (as confirmed from Raman spectroscopy), so the probability of the cations sitting in close proximity to each other will increase. Therefore, all the existing mobile oxygen vacancies come together to form deep traps and as a consequence, small micro-domains may be generated due to the ordering/re-arrangement of the isolated cations and oxygen vacancies within the lattice.18,19 Although the increased concentration of deep traps favours a fast electron–hole recombination rate, however, beyond a certain level, their presence can suppress the photocatalytic efficiency13 (as shown in Fig. 5a for the case of CS-15). The poor photocatalytic performance of sample CS-15 may also be due to the presence of high In doping which may act as charge recombination centres resulting in the decrease of the photocatalytic activity. Therefore, the achievement of an optimum level of oxygen vacancies by incorporating In into the Ce lattice is mandatory to achieve an enhanced photocatalytic performance. Therefore, an optimum level of oxygen vacancies by incorporating In into the Ce lattice is mandatory to achieve better photocatalytic performance. On the basis of the present findings, 10% indium doping in the ceria lattice is the optimum level, which provides the optimum number of oxygen vacancies for achieving the highest photocatalytic activity.
The effectiveness of CeO2 nanocrystals (CS-0, CS-05, CS-10 and CS-15) as a test model substrate in determining the photocatalytic degradation efficiencies of the cationic dye (MB) was also tested and the results are shown in Fig. 5b. Interestingly, the dye degradation pattern of MB remains the same as for MO and the catalyst CS-10 showed the highest photocatalytic activity among all the catalysts. However, the overall catalytic efficiency for all the catalysts was lower compared to the degradation of anionic dyes. The surface state of the CeO2 could depend on the pH of the medium. If the pH of the solution is greater than the isoelectric point of ceria (6.4–6.5),20 the surface becomes negatively charged and when it is lower than the pH of the solution, the surface becomes positively charged. Also, if the doped material has a higher isoelectric point (indium oxide isoelectric point is (8.5–8.7)21) than the pH value of the dye, the surface is assumed to be positively charged. Therefore, the anionic dyes have a greater affinity toward the catalyst, which may be a potential reason for the higher photodegradation efficiency of the CeO2 based nanocrystals toward anionic dyes.
Based on these results, it can be concluded that CeO2 nanocrystals are a good catalyst towards the photodecomposition of both anionic and cationic dyes, but anionic dyes have a greater affinity toward the catalyst, resulting in a much higher photodegradation.
The kinetics of MO and MB decolourisation at room temperature are presented in Fig. 5c, and are found to follow pseudo-first order reaction. The decomposition rates of MO and MB over CS-10 samples can be calculated as 0.00267 min−1 and 0.0134 min−1, respectively, signifying that the photodegradation process of MO is much faster than that of MB. It is commonly believed that at high temperatures, the oxygen vacancies may become mobile and the probability to form deep trap sites is very low. Therefore, the degradation of MO with CS-10 was carried out at elevated temperatures as shown in Fig. 5d. It can be clearly observed that at high temperatures, the catalyst, CS-10 degraded the dye faster than at room temperature. The enhanced photocatalytic activity at high temperatures can also be attributed to the enhanced oxygen ion mobility and possible lattice (phonon) scattering effects. At high temperatures, the ions/atoms start to vibrate strongly thereby increasing the scattering probability of electrons/holes within the lattice, hence leading to the recombination of photo-generated carriers.
By exposing ceria nanocrystals to UV radiation, electrons are excited from the O2p valence band to an empty or partially empty conduction band (Ce4f), forming the electron–hole pair (Ce4+(e−) and O2−(h+) pair).14 The photo-generated holes get trapped by oxygen ions, whereas electrons were locally surrounded by Ce ions. The photo-generated electrons reduce adsorbed electron acceptors (e.g., O2) by forming superoxide radicals, and the holes oxidize organic substances adsorbed on the catalyst or react with water forming hydroxyl radicals, which can also be involved in the decomposition of the organic substrates.22 For elevated temperatures, the effect of increased lattice oxygen ion mobility surpasses the lattice vibration effect which is beneficial for the photo-generated electron and hole pairs, and as a consequence, superior photocatalytic activity was observed.
To gain more insights into the thermo-photocatalytic performance of CS-10, the catalytic activities of all the samples in the absence of light (but at high temperature 100 °C) were investigated. The thermocatalytic performances of all the samples were very low as compared to the results obtained under UV irradiation as shown in Fig. S2.†
The enhanced photocatalytic activity at high temperatures (under UV light) can be attributed to the enhanced oxygen ion mobility and possible lattice (phonon) scattering effects. At high temperatures, the ions/atoms start to vibrate strongly thereby increasing the scattering probability of electrons/holes within the lattice. As a result, the scattering of electrons/holes increases within the lattice leading to the recombination of photogenerated carriers.13,23 This is further supported by a previous report on Y-doped CeO2 where the increase in the photocatalytic activity at elevated temperatures was also attributed to the presence of higher amounts of oxygen vacancies and the enhanced oxygen ion mobility.13 Another study conducted by Li et al.23 had also indicated an increase in the separation of charge carriers at higher temperatures. At elevated temperatures, ions/atoms tend to vibrate strongly, leading to more phonons which subsequently increase the scattering probabilities of ions/electrons with phonons. As, In-doped CeO2 samples exhibit a higher amount of oxygen vacancies (based on the Raman and XPS data), the oxygen ion mobility in these samples was also improved by increasing temperature.24 The improved mobility of the lattice oxygen ions rather than lattice vibrations leads to the improved separation of photogenerated electrons and holes, which ultimately enhances the photocatalytic activity of the In-doped samples. Furthermore, the rate constants for the photocatalytic activity of CS-10 at room temperature and 100 °C for all three different conditions (in the absence of light, with visible light radiation and UV light radiation) were calculated and are plotted in Fig. S3.†
To assess the efficiency of the CS-10 samples for repetitive use, the degradation of MO was carried out repeatedly for five cycles as shown in Fig. 6. It is clearly observed that the nature of degradation remains unaltered and the inherent efficiency of CS-10 persisted (less than 5% decrease from its initial activity during the photodegradation process) without self-degradation.
In order to calculate the band-gap energies of the as-prepared samples, the Schuster–Kubelka–Munk absorption function was plotted against the photon energy (hv), according to the equation:25,26
ahv1/2 = A(hv − Eg) | (4) |
Ecb(CeO2) = χ(CeO2) − EC − 0.5Eg(CeO2) | (5) |
Evb(CeO2) = Eg(CeO2) − Ecb(CeO2) | (6) |
In this study, the amount of In dopant greatly influences the photocatalytic activity of the produced CeO2 nanocrystals. The photocatalytic activity of the sample increases with increasing In content before it reaches a maximum value of 10 at% In doping. Then, the activity obviously decreases with increasing In content above 10 at%. This is due to a small amount of In3+, which can act as an electron acceptor (from In3+ to In2+) and/or a hole donor (from In3+ to In4+) to facilitate charge carrier localization and thus, prolonged separation by trapping at energy levels close to the conduction or valence bands, respectively,25,28 as schematically illustrated in Fig. 8.
Fig. 8 Schematic diagram of the degradation mechanisms for the MO dye with indium doped CeO2 nanocrystals under ultraviolet-visible light irradiation. |
Based on crystal field theory, In2+ and In4+ are relatively unstable compared to In3+, and this likely leads to the transfer of the trapped charge carriers from In2+ or In4+ to the adsorbed O2 and surface hydroxyl (–OH) to regenerate In3+.28,29 The newly produced active species (such as ˙OH and O2−) will initiate the subsequent photocatalytic reactions.
It is important to note that the photocatalytic activity of In-doped CeO2 nanocrystals is strongly dependent on the concentration of the In dopant since the In3+ ions can serve not only as a mediator of interfacial charge transfer but also as a recombination centre.26 In this study, the optimal concentration of the In doping is 10 at%. Above this concentration, In3+ ions steadily become recombination centres and the photocatalytic activity steadily decreases through quantum tunnelling.28–30 Another possible reason for the enhancement in the photocatalytic activity of CS-10 is that In doping also induces oxygen vacancies (as confirmed by the XPS and Raman spectroscopy results), which are highly advantageous for the separation of photo-induced charges. The increase in the oxygen vacancy concentration can also generate an increment in the surface hydroxyl amount, which is beneficial for photocatalytic reactions31 to improve waste water treatment efficiencies.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5nr06588g |
This journal is © The Royal Society of Chemistry 2016 |