Bappaditya Mandal and
Aparna Mondal*
Department of Chemistry, National Institute of Technology, Rourkela – 769008, Odisha, India. E-mail: aparnamondal@gmail.com; Tel: +91 6612462658
First published on 17th April 2015
Novel, high surface area, mesoporous and crystalline samarium-doped ceria (CeO2:Sm3+) nanopowders were successfully synthesized by combining the excellent properties of both microwave heating and surfactants and were used as remarkably efficient new photocatalysts for the degradation of a representative azo dye, Acid Orange 7 (AO7), in an aqueous medium under natural sunlight without the addition of any external reagents like peroxide, acid or base. The synthesized nanopowders were characterized by X-ray diffraction (XRD), transmission electron microscopy (TEM), Brunauer–Emmett–Teller (BET) analysis and UV-vis diffuse reflectance spectroscopy. The effects of calcination temperature, pH of the medium, catalyst dosage and irradiation time on the decolorization of AO7 were investigated and are discussed in this paper. Sm3+ doping in CeO2 narrowed the band gap and significantly enhanced the photocatalytic degradation of the azo dye. The photocatalytic degradation of AO7 was also investigated by using certain radical scavengers and the results suggest that under solar light irradiation predominantly positive holes and superoxide radicals (O2˙−) act as the active species in the degradation process. Our results suggest that the materials developed here are a promising alternative solar light sensitive photocatalyst.
For degrading organics and toxic materials, many treatment methods require high pressure (0.5–20 Mpa) and high temperature (80–320 °C),4 which limit their practical application. To resolve this problem, heterogeneous photocatalysis using TiO2 and ZnO photocatalysts in particular has emerged as a promising advanced oxidation process for the purification of dye-containing wastewater and also provides solutions to many problems related to environmental pollution.5–7 Mostly these processes involve UV/H2O2, or UV/O3 for the oxidative degradation of dyes.6–8 Although H2O2 is an environmentally friendly oxidant and widely used for environmental remediation, disinfection, etc. the current industrial process for H2O2 production requires toxic solvents and a high energy input, and therefore is not environmentally benign.9,10 Photo-assisted oxidation by solar irradiation has been found to be a cost effective and ideal approach, as it is readily available and free and can be used directly to oxidize or degrade hazardous organic chemicals.5 The advantage of CeO2 (band gap Eg ∼ 2.9 eV) in comparison with TiO2 (Eg ∼ 3.0 to 3.2 eV for anatase and absorbs less than 5% UV light of the sunlight) is that it absorbs over a larger fraction of the UV spectrum and the onset of absorption of CeO2 is ∼440 nm.11 CeO2 nanopowder has been reported to be a more efficient photocatalyst than commercial TiO2 P25.11–13 So, recently, CeO2 has gained immense interest as a photocatalyst for the photocatalytic degradation of various dyes as it is chemically stable, inexpensive and can be easily synthesized through various reproducible solution-based synthesis routes.14–18 It has already been used in various catalytic applications such as fluid cracking, purification of harmful gases in three way automotive catalytic converters, water splitting for the generation of H2 gas, as well as in biomedicine, solar cells, and inorganic phosphors.19–23
The key to the wide range of catalytic applications of ceria-based materials is that CeO2, by shifting some Ce4+ to Ce3+ ions, can easily produce oxygen vacancies, which act as sources for the oxygen involved in reactions taking place on the catalyst surface.15,18,24 The effectiveness of a CeO2 photocatalyst can be enhanced by introducing other metal ions into it by forming a solid solution.16,18 It has also been reported that lanthanide ions including La3+, Nd3+, Pr3+, Sm3+, or Eu3+ doped in TiO2 photocatalysts improved the separation rate of photo-induced charge carriers and greatly enhanced the photocatalytic activity of TiO2 over pure TiO2.25 Cai et al. observed improved degradation of AO7 in the dark as well as under visible light irradiation in a CeO2–H2O2 system after Fe3+ doping.18
It is well known that the concentration of oxygen vacancies can be remarkably enhanced after doping with Sm, due to Sm3+ → Ce4+ substitutions.20 Sm3+ doping also induces the least distortion of the parent lattice when oxygen vacancies are created in the CeO2 lattice for charge compensation. Many studies have reported on the Sm0.2Ce0.8O1.9 system showing the highest electrical conductivity (specific to ionic conductivity) required for solid oxide fuel cell applications.22 Also, samarium-doped ceria nanoparticles are very important in developing new luminescence devices.20 But no investigation on Sm3+-doped CeO2 as a photocatalyst has been reported. Therefore, in this study, we aimed to obtain an efficient photocatalyst by doping CeO2 with Sm3+ ions.
Recently, Sun et al. reported a 97.6% degradation efficiency of AO7 by nitrogen-doped monocrystalline CeO2 (N:
Ce molar ratio of 0.3) nanoparticles synthesized solvothermally at 120 °C for 24 h in an autoclave.14 They investigated the degradation of a 20 mg L−1 AO7 solution (at pH = 3.0) containing 1 g L−1 of the as-synthesized sample, which had air bubbled through it at 60 °C at a flow rate of 300 mL min−1, using a domestic 10 W compact fluorescent lamp.14 Salker et al. reported the low solar light assisted photocatalytic activity of CeO2 for the degradation of the textile dye Naphthol Blue Black.26 But a significant increase in the photocatalytic activity was observed by substitution of 30 mol% Mn4+ in the CeO2 crystal lattice due to the enhancement of the redox couple Ce4+–Ce3+ and decrease in the band gap energy.26 A few studies have also demonstrated the potential use of CeO2,12,13,15 or doped CeO2 (ref. 16–18) for the photocatalytic degradation of dyes under visible light irradiation (mostly λ > 420 nm).13,16,18 However, in most cases they are successful in the presence of high concentrations of oxidizing agents,15,18 and/or under specific pH conditions.14 These conditions limit the industrial application of nanomaterials in environmental remediation. There is a need to develop general, simple and economical routes for designing novel nanomaterials which can efficiently remove various contaminants in the environment under solar illumination. Due to higher reaction rates and selectivities, microwave chemistry has recently been shown to be a fast-growing research area with immense potential.27–29 Microwave-assisted methods have been effectively employed to synthesize a great many novel nanostructures with various shapes.27–29
However, the application of microwave heating with the addition of surfactants to synthesize mesostructured nanomaterials has hardly been exploited until now. Here, we used a surfactant-assisted microwave heating method, which enables the convenient low temperature preparation of efficient CeO2 photocatalysts with a higher surface area and higher crystallinity. In the present study, we demonstrated a new photochemical remediation method for dye-polluted waters by using samarium-doped ceria nanoparticles and natural sunlight only. The effects of the calcination temperature, pH and the amount of nanopowder were also investigated. Though several studies on the mechanistic details of dye degradation on TiO2 have been reported under UV as well as visible light, some insight of the active species for the photodegradation process on CeO2 under solar light irradiation is given here for the first time.
![]() | ||
Fig. 1 XRD patterns of the as-prepared CeO2:Sm3+ powders (a) and the powders calcined at (b) 500 °C, and (c) 800 °C for 2 h. |
With increasing calcination temperature, the peaks become narrower and their intensity increases due to the increase in the crystallinity of the Sm3+-doped CeO2 crystallites. The average crystallite sizes of the Sm3+-doped CeO2 samples were calculated from the X-ray line broadening of the (111), (200), (220) and (311) reflections using Scherrer’s equation (peak width obtained after correction for the instrumental broadening).31 The instrumental peak width (βi) was subtracted from the measured peak width (βm) at half maximum intensity; i.e. where the subscript “s” indicates the contribution from the sample. The calculated average crystallite sizes were 7.9, 8.3, and 22.0 nm for the as-prepared sample and the samples heated at 500 °C and 800 °C, respectively. It is interesting to observe that a very slow increase of crystallite size occurs in the early stage of calcination from 100 °C to 500 °C. While an activated growth of crystallite size from 8 to 28 nm was observed in pure CeO2 on increasing the annealing temperature from 100 to 800 °C. This signifies that the Sm3+ doping in CeO2 stabilizes the solid solution against crystal growth during calcination.32
The lattice parameter (a) value (0.5429 nm) of the as-prepared mesoporous Sm3+-doped CeO2 sample is higher than that of the mesoporous as-prepared CeO2 sample (a = 0.53929 nm), prepared under the same experimental conditions, but very close to that of bulk cubic CeO2 (0.5411 nm). Due to the small size of the nanocrystallites, the formation energy of oxygen vacancy is reduced and the two electrons from the oxygen atom are captured by two lattice site Ce4+ ions and hence favors the formation of reducing Ce4+ ions to the Ce3+ state. Using Kroger–Vink notation, the vacancy formation reaction is represented as
The presence of Ce3+ ions allows the increase of the lattice parameter due to the fact that the Ce3+ ion has a larger ionic radius compared to Ce4+ and the concentration of the Ce3+ ion decreases with increase in calcination temperature.22 Again, the ionic radius of Sm3+ (0.1079 nm) is larger than Ce4+ (0.0971 nm), hence little lattice expansion is expected due to incorporation of Sm3+ into the CeO2 lattice, and the results are in good agreement with previous reports.21,22 In order to balance the charge compensation, Sm3+ additives create oxygen vacancies in the CeO2 lattice, for example
A decrease in the a value to 0.5420 nm was observed when the sample was calcined at 500 °C. This lattice contraction observed in the mesoporous structure (as evidenced from the HRTEM results) may be due to the strains imposed by the pore wall structure, and also the oxidation of Ce3+ to Ce4+ occurring during the calcination process. As expected, on the further increase of the calcination temperature to 800 °C the lattice parameter value decreased to 0.5381 nm due to the increase in the crystallite size. The consistency of the lattice parameter infers that the introduction of Sm3+ ions into CeO2 can stabilize the host lattice at high temperatures and no new phase is formed.
The textural properties of the Sm3+-doped CeO2 photocatalysts were investigated by nitrogen sorption analysis. The nitrogen sorption isotherms and the corresponding pore size distribution (in the inset) of the (a) as-prepared and (b) heated at 500 °C powders are shown in Fig. 2. Both the samples show a type IV isotherm characterized by the presence of a hysteresis loop covering a broad relative pressure range, indicating a mesoporous structure containing some micropores. The surface area and total pore volume of the as-prepared sample were as high as 112 m2 g−1 and 0.095 cm3 g−1 respectively, and increased to 135 m2 g−1 and 0.16 cm3 g−1 when the sample was calcined at 500 °C. Subsequent thermal treatment not only improves the crystallinity but also maintains a high specific surface area, which is an important criterion for good photocatalysts. The BJH pore size distributions (see the inset of Fig. 2) reveal that the as-prepared sample has a monomodal pore size distribution in the range of 2 to 30 nm, with an average pore size of 5.5 nm, while the sample calcined at 500 °C shows a similar pore size distribution but has a larger average pore size of 6.2 nm, in line with the TEM results. As shown in Table 1, the surface area of the Sm3+-doped CeO2 sample is higher than that of the pure CeO2 (82 m2 g−1 at 500 °C) heat treated at a particular temperature. This may be due to Sm3+ doping in the ceria stabilizing the solid solutions and decreasing the crystallite size. Similarly, increases in surface area on doping some trivalent cations such as Fe3+ and Sm3+ in CeO2 have also been reported.18,22
![]() | ||
Fig. 2 N2 adsorption–desorption isotherms (with the BJH pore size distribution plots in the inset) of the CeO2:Sm3+ powders (a) as-prepared and (b) calcined at 500 °C for 2 h. |
Sample | Crystallite size (nm) | Lattice volume (nm3) | Surface area (m2 g−1) | Pore volume (cm3 g−1) | Band gap (eV) | %photodegradation |
---|---|---|---|---|---|---|
CeO2 | 8.6 | 0.1582 | 82 | 0.15 | 2.86 | 60 |
CeO2:Sm3+ | 8.3 | 0.1592 | 135 | 0.16 | 2.79 | 82 |
On calcining the Sm3+-doped CeO2 material at the higher temperatures of 600 °C and 800 °C, the BET surface area decreased to 70 and 30 m2 g−1, respectively. This is because the mesoporous structure collapses due to the structural shrinkage of the skeleton and increased material density during the calcinations. Natile et al. reported a surface area value of 72 m2 g−1 for pure CeO2 calcined at 250 °C and prepared by the microwave-assisted heating of an aqueous solution of [(NH4)2Ce(NO3)6] containing 1 wt% poly(ethylene glycol) 2000 (PEG) and 1 wt% sodium acetate (NaOOCCH3).29 These mesoporous Sm3+:CeO2 materials having large pores and high surface areas are expected to show high activity in adsorption and photocatalytical degradation.
The TEM image shown in Fig. 3 further confirms the mesoporous structure of the Sm3+-doped CeO2 calcined at 500 °C for 2 h. It is evident from the HRTEM image in Fig. 3b that the particles were uniform in shape and also crystalline in nature as the lattice fringes could be observed. The estimated particle size was around 8 nm, which is consistent with the XRD results. The selected-area electron diffraction (SAED) patterns (inset of Fig. 3a) consist of single spots superimposed on rings, indicating that the crystal domains within the pore walls are quite small. As microwave heating leads to uniform heating of the whole synthesis mixture, it led to a homogeneous microstructure as is evident from the micrographs. From the EDS spectrum (inset of Fig. 3b), the expected peaks of Sm, Ce and O were observed from the CeO2:Sm3+ sample. The addition of 1.0 mol% Sm2O3 into the reaction solution led to the incorporation of 2.2 at% Sm in the CeO2:Sm3+, as confirmed by EDS analysis.
![]() | ||
Fig. 3 (a) TEM image, (b) HRTEM image and SAED pattern (inset of a) of the CeO2:Sm3+ powders calcined at 500 °C for 2 h. The EDS of the sample is shown as the inset of (b). |
The band gap energies (Eg) of the samples were calculated by fitting the absorption data to the equation Eg = 1240/λAbsorption Edge. The bandgap energies were calculated to be 2.85 eV for the pure as-prepared CeO2, and 2.78, 2.79, 2.81, and 2.91 eV for the Sm3+:CeO2 as-prepared, and calcined at 500, 600 and 800 °C, respectively. These values are smaller than the values obtained for the bulk CeO2 (3.08 eV),34 CeO2 microspheres (2.81 eV), and CeO2 microflowers (3.08 eV).35 This shows that Sm3+ doping can shift the absorption edge of CeO2 to the visible light range and narrow the band gap as can be seen from the inset of Fig. 4. Modification of the band gap is expected to improve the efficiency of the utilization of sunlight especially in the visible region. Thus, narrowing of the band gap is beneficial for improving the photoabsorption performance of Sm3+-doped CeO2 and enhances its photocatalytic performance. The redshift of the optical band gap of cerium oxide (CeOx) films was correlated with the increase of Ce3+ content at the grain boundary.36 Chen et al. also observed a blueshift of the band gap due to the valence transition of Ce3+–Ce4+ induced by thermal annealing.37 The quantum-size effect is expected to enhance the band gap of the materials with decreasing particle size due to the higher localization of the energy bands.36 The content of Ce3+ and oxygen vacancies from the surface progressively decreased with the growth of the CeO2 nanoparticles during annealing in an air atmosphere. The decrease of Ce3+ concentration and the corresponding decrease of vacancies (defects) content will eliminate some localized defect energy states within the band gap, and thus result in an increase in the band gap. So, the blueshift of Eg in our Sm3+:CeO2 samples with increasing calcination temperatures is due to the Ce3+ content decrease, not due to the results of the quantum-size effect. The results also suggest that the Sm3+ doping in CeO2 has induced a higher concentration of Ce3+. These defects provided many activity sites on the CeO2 in the photocatalytic process.17
Fig. 6 shows the direct influence of the initial solution pH values of 2.9, 4.2, 6.9, and 9.1 on the degradation efficiency of the dye on the catalyst. It is interesting to observe that the Sm3+:CeO2 photocatalyst was very efficient in degrading (99.5% to 95%) the dye over the pH range from 2.9 to 9.1. At a neutral pH of 6.9, the photodegradation of AO7 was 98%. The photodegradation efficiency of AO7 was slightly higher in acidic solutions than in alkaline solutions. The zero point charge of ceria is at pH 4.2, which is close to the neutral pH of 6.9 of the solution. The AO7 dye acts as a Lewis base due to the presence of a negatively charged acidic sulfonate group on it and can easily adsorb on the positively charged Ce–OH2+ catalyst surface in an acidic pH, and hence increase the degradation rate. In an alkaline medium, maybe due to the increased OH− ion concentration with the increased pH, there is a competitive adsorption between hydroxyl groups and the dye molecules. In addition, the negatively charged catalyst surface repels the dye, thereby the degradation is a bit slower.
Generally, the specific surface area, surface states and crystallinity play key roles in the degradation efficiency.38 The catalyst calcined at 600 °C possesses a lower surface area than the as-prepared catalyst though it shows a prominently 42% higher photodegradation efficiency. This indicates that surface area is not the only determining factor to cause the different efficiencies seen in this study. The as-prepared catalyst also contains a certain amount of SDS molecules, which are attached to the inorganic framework and are difficult to remove by just a simple washing process. The SO4− ions of the SDS molecules present in the as-prepared sample also decrease the adsorption of the negatively charged dye on the surface, and thereby decrease the photodegradation efficiency. Carbonaceous species formed during the sunlight irradiation and low crystallinity could be possible reasons for the unusually low activity of the as-prepared catalyst despite its high surface area. Wang et al. found that organic impurities (carbonaceous species) present in the well-known photocatalyst TiO2 diminished the photocatalytic activities by accelerating the electron and hole recombination by acting as intermediate levels under UV light.39
As the decomposition temperature of SDS is ∼210 °C, the catalyst calcined at 500 °C is surfactant free, as was also evident from IR analysis (not shown here). This calcination temperature is high enough to develop well the crystalline structure of the catalyst. The highest photocatalytic activity of the catalyst calcined at 500 °C may be attributed to the high surface area, pore volume and mesoporous morphology. Also the catalyst can absorb light in the near UV and also absorbs a larger fraction of the solar spectrum than the others. As the temperature reaches above 500 °C, the crystallites grow larger and the surface area and total pore volume go down, leading to the poor performance of the samples. The photocatalytic activity reduction for the catalyst calcined at 800 °C may be attributed to the extensive decrease of the specific surface area and its nonporous morphology.
![]() | ||
Fig. 8 Evolution of the UV-vis spectra with irradiation time for the photocatalytic degradation of AO7 in aqueous solution in the presence of the CeO2:Sm3+ catalyst calcined at 500 °C. pH neutral. |
![]() | ||
Fig. 9 A photograph of the AO7 solutions in the presence of the CeO2:Sm3+ calcined at 500 °C as the catalyst under solar irradiation at different time intervals (h) as marked therein. |
It was observed that after the first 1 h of solar irradiation the absorbance peaks including the shoulder in the visible region decreased by over 82% from their initial values and almost complete decolorization (98%) was achieved after 10 h of photocatalysis. The peaks at 228, 254 and 310 nm decreased significantly. This indicates that the benzene and naphthalene rings of AO7 were broken into smaller species.
A tentative photocatalytic reaction mechanism for the degradation of AO7 over the CeO2:Sm3+ can be proposed as the following. Under solar light irradiation, electron–hole pairs are generated from CeO2. Part of the photogenerated e− also gets trapped by oxygen vacancies (F centres) before moving to the conduction band.20 Adsorbed surface oxygen first interacts with the Ce3+ and oxygen vacancies to generate superoxide radicals and then forms H2O2, ˙OOH and ˙OH. Due to the short lifetime of the resulting ˙OH on the photocatalyst, the contribution of the oxidizing species in the bulk solution could be insignificant. Active h+ and O2˙− generated in the bulk solution is accepted to be produced by the following equations:
CeO2 + hν → h+ + e− |
Charge carrier recombination: hvb+ + ecb− → heat |
Ce4+ + e− → Ce3+ |
Ce3+ + O2(ads) → Ce4+ + O2˙− |
O2(ads) + F → O2˙− + F+ |
H2O + h+ → ˙OH + H+ |
O2˙− + H+ → ˙OOH |
˙OOH + H+ + e− → H2O2 |
H2O2 + e− → ˙OH + OH− |
h+ + dye → degraded product |
O2˙− + dye → degraded product. |
The adsorption experiments indicated that AO7 adsorbed strongly on the CeO2 surface. Two oxygen atoms from the sulfonate group of AO7 are linked with two Ce surfaces and form an inner-sphere complex. AO7 molecules, being strong electron donors due to the oxygen atoms of SO32−, are able to directly interact with valence band holes. The valence band holes are primarily captured by the adsorbed AO7 molecules rather than by the adsorbed water or hydroxyl groups as the oxidative potential of AO7 (0.76 V versus NHE),41 is much lower than that of the photogenerated holes. The pathway for photodegradation under solar light as presented in our study is different from the dye self-sensitization pathway implicated during visible light irradiation (λ > 420 nm) using pure CeO2 as the photocatalyst.16 Another study reported that active h+ play a major role in AO7 degradation using P-25 under UV irradiation.
Kondarides and coworkers reported that it takes about 50 h to totally decolorize AO7 solutions with a concentration of up to 100 mg L−1 by TiO2 under visible light irradiation.42 A Cu2O/CeO2 composite photocatalyst (1 g L−1) with a surface area of 53 m2 g−1 degraded about 96.2% of AO7 (35 mg L−1) after 4 h of visible light irradiation.16 When 0.05 g of Co-doped CeO2 nanocubes (catalyst) was added with 15 mL of AO7 (0.3 mmol) and exposed under UV illumination (λmax = 365 nm) for 8 h, a specific change in the color of the dye was observed due to its decomposition.43 In particular, it is interesting to evoke some reasons why the mesoporous CeO2:Sm3+ prepared through the microwave route exhibited enhanced solar light photocatalytic activities. The first explanation is that the shifts of the absorption edge of CeO2 to the visible-light range narrow the band gap and have strong absorption in the UV region as well as definite absorptions in the visible region. This decrease in band gap may help to facilitate the propagation of electrons to the conduction band initiating the photolysis reaction.44 Furthermore, the photocatalytic activity can be enhanced by the high crystallinity of these new mesoporous Sm3+:CeO2 catalysts. Mesoporous ceria structures with a high specific surface area, large pore size and thermal stability have several advantages for high adsorption, which is a key factor for the degradation rate and photocatalysis. The mesoporous structure space effectively enhances the spatial dispersion, which results in a higher surface area and also mass transportation of molecules both into and out of the pore structure.45 It has been suggested that the degradation of an organic dye in the ceria/H2O2 system relies significantly on its adsorption on the surface of the CeO2.15 We found that the adsorption effects of AO7 in the dark after 11 h by pure CeO2 and 1 mol% Sm3+-doped CeO2 are 52.7 and 65%, respectively. The very high adsorption of AO7 by CeO2 is attributed to the high surface area and pore volume of the mesoporous catalysts, high oxygen vacancies on the surface originating from the facile Ce3+/Ce4+ redox cycle,15 and to the ability of cerium ions to form complexes with the electron-rich sulfonate group (SO3−) of AO7.13 Xiao et al. also demonstrated that Sm3+-TiO2 had a higher methylene blue adsorption capacity than undoped TiO2.46 Sm3+ doping further improved the concentration of Ce3+ or, in other words, the concentration of oxygen vacancies in CeO2 samples as well as enhanced the adsorption capacity of AO7, which resulted in better catalytic performances. Tang et al. reported that the oxygen vacancies can not only act as impurity levels in the band structure of ZnO but can also work as electron traps to accept the photogenerated electrons.47 Oxygen vacancies on the CeO2 catalyst surface can easily trap the photogenerated electrons and efficiently benefit the separation of excited electron–hole pairs and thus can play a more functional and effective role in photocatalytic applications.
In our study, the optimum Sm3+ concentration in the CeO2 was 1.0 mol%, at which the recombination of photo-induced electrons and holes could be effectively inhibited, and thereby the highest photocatalytic activity was achieved. At natural pH conditions, the photodegradation of AO7 under 1 h of solar light irradiation by the CeO2 photocatalysts (calcined at 500 °C) with 0, 0.5, 1.0, 2.0 and 5.0 mol% Sm3+ doping were 60, 66, 82, 78, and 72%, respectively. It is well documented that a higher adsorption capacity with a higher Sm3+ content does not lead to a higher photocatalytic activity in TiO2.46 Cai et al. also showed that doping Fe3+ improved the defect concentration and thus favored the catalytic activity of CeO2 at low levels, but retarded the degradation of AO7 at high levels, with an optimal Fe/Ce ratio of 1/100.18 Probably at higher Sm3+ concentrations the Sm may be present at the surface of particles and form clusters, which are detrimental for photocatalytic activity carried out at the surface.48
This journal is © The Royal Society of Chemistry 2015 |