Guisheng
Li
*,
Ya
Zhang
,
Ling
Wu
,
Fang
Wu
,
Rong
Wang
,
Dieqing
Zhang
,
Jian
Zhu
and
Hexing
Li
*
The Education Ministry Key Lab of Resource Chemistry, Shanghai Key Laboratory of Rare Earth Functional Materials, Shanghai Normal University, Shanghai 200234, China. E-mail: lgscuhk@yahoo.com.cn; hexing-li@shnu.edu.cn; Fax: +(86)21-64322272; Tel: +(86)21-64322272
First published on 20th March 2012
A novel La2NiO4 photocatalyst in layered perovskite crystal was synthesized by co-alcoholysis of La(NO3)3 and Ni(NO3)2 under solvothermal conditions. It exhibited high activity for mineralizing 4-chlorophenol under visible light irradiation and even in the dark, leading to round-the-clock photocatalysis. The photocatalytic mechanism of the La2NiO4 in the absence of light irradiation was discussed based on both the structural characterizations and the kinetic studies. Both the La2NiO4 crystals and the organic molecules played key roles for starting the degradation reaction in dark. The 4-chlorophenol firstly ionized into anions, followed by donating electrons to La2NiO4 owing to the positively charged surface. Then, the electrons could react with dissolved O2 to produce ˙O2−, which subsequently reacted with H+ to form ˙OH radicals. These highly active ˙OH radicals could oxidize 4-chlorophenol into CO2, leading to complete degradation.
Photocatalysis has been widely used in degrading organic pollutants owing to the energy saving, easy operation and the absence of secondary pollution. Most studies are focused on the TiO2 photocatalyst owing to its strong oxidizing ability, excellent stability and low toxicity.19,20 However, TiO2 can only be excited by UV light due to its wide band gap (Eg = 3.2 eV), which accounts for merely 4–5% sun light spectrum on the earth surface. To date, great achievements have been obtained for fabricating visible-light photocatalysts by doping TiO2 or designing non-TiO2 photocatalysts.21–23 However, these catalysts can work well only in ther daytime and cannot work in the dark due to the absence of light irradiation. Recently, Wu et al. reported a novel approach to mineralize azo dyes in the absence of light irradiation.24 Rosseinsky and co-workers also found that HCa2Nb3O10 could oxidize methyl orange at room temperature in the dark.25 However, these photocatalysts exhibited low efficiencies in the dark and some co-catalysts or the additional oxidant reagents were required for enhancing the activity. Meanwhile, the degradation mechanism in the dark was not clearly elucidated. Herein, we report for the first time a layered perovskite La2NiO4 crystal, acting as a round-the-clock photocatalyst for breaking down 4-CP with excellent performance and strong durability at ambient conditions. The degradation mechanism in the absence of light irradiation was examined based on both the structural characterizations and kinetic studies, which revealed that the La2NiO4 could generate photoelectrons via activation with visible light irradiation and also trap electrons from some reactant molecules (e.g., 4-CP) in the dark, leading to round-the clock degradation reactions.
For comparison, the traditional visible-light photocatalyst (TiO2−xNx) was also synthesized according to the method reported elsewhere.30 Briefly, 17 mL tetrabutyl orthotitanate was added dropwise into 100 mL aqueous solution containing 1.0 wt% NH3. After being stirred for 1 h and aged at 25 °C for another 24 h, the solid product was filtrated, washed thoroughly with water, and dried at 60 °C, followed by calcining at 60 °C for 2 h.
![]() | ||
Fig. 1 XRD pattern (upper) and crystal cell of the La2NiO4. a: (100), b: (010), c: (001). |
Fig. 3a revealed that La2NiO4 displayed a type-II N2 adsorption–desorption isotherm indicative of a non-porous structure, corresponding to a low surface area (7.3 m2 g−1). Fig. 3b demonstrated that the La2NiO4 showed spectral responses in the whole UV-Vis-IR region, which could possibly be attributed to the presence of multiple consecutive energy levels.
![]() | ||
Fig. 3 (a) Nitrogen adsorption–desorption isotherm and (b) UV-Vis-IR adsorption spectrum of the La2NiO4 sample. |
4-CP was chosen as the model phenolic compound to test the photocatalytic performance of the as-prepared La2NiO4 crystals. Preliminary experiments demonstrate that no significant degradation of 4-CP was observed in the absence of La2NiO4, regardless of the light irradiation. Fig. 4 shows the influence of reaction temperature or reaction time on the 4-CP degradation yield under visible light irradiation and in the dark, where C0, C, k, and t refer to the initial concentration and the real-time concentrations of 4-CP, the reaction constant and the reaction time. Obviously, 4-CP could be effectively degraded by the La2NiO4 crystals both under visible light irradiation and in the dark. The linear relationship between Ln(C0/C) and t demonstrates that the reaction is pseudo-first-order with 4-CP concentration.27 The reaction constants at 10, 30 and 50 °C were calculated as following: kL50 (0.27 h−1) > kL30 (0.16 h−1) > kL10 (0.10 h−1) under visible light irradiation and kD50 (0.11 h−1) > kD30 (0.074 h−1) > kD10 (0.060 h−1) in the dark. These k values reveal that the reaction rate increases rapidly with increasing temperature. At the same reaction temperature, the k values obtained under visible light irradiation are much higher than those obtained in the dark, suggesting that 4-CP could be more easily degraded by photocatalysis. According to the influence of the reaction temperature on the reaction rate, the apparent activation energies were calculated as 73 kJ mol−1 and 61 kJ mol−1 for the 4-CP degradation reactions under visible light irradiation and in the dark, respectively. The difference in either the k value or apparent activation energy implied that the 4-CP degradation on La2NiO4 crystals under visible light irradiation and in the dark might follow different reaction mechanisms.
![]() | ||
Fig. 4 Kinetic curves of 4-CP degradation obtained under visible light irradiation (λ > 420 nm) (a) and in the dark (b). Reaction conditions: 75 mg La2NiO4, 50 mL 5.0 × 10−6 M 4-CP aqueous solution, reaction T = 10 °C (-★-), 30 °C (-▼-), and 50 °C (-■-). |
As shown in Fig. 5, the HPLC-MS spectra display a peak at retention time of around 7.0 min is indicative of a 4-CP molecule (m/z = 127). The other two peaks observed at retention times of around 3.0 min were ascribed to the injection fluctuation. With the increase of reaction time either under visible light irradiation or in the dark, the signal corresponding to the 4-CP molecule decreases gradually owing to the degradation reaction. No other peaks are detected by HPLC-MS during the reaction until the complete removal of 4-CP. It suggests that, different from those reported so far,28 the 4-CP molecule is directly mineralized without formation of relatively stable by-products including hydroquinone, benzoquinone, and hydroxyhydroquione etc., possibly owing to the super strong oxidizing ability of the active species generated from the La2NiO4 crystals. By measuring the total organic carbon (TOC), we found that a more than 95% mineralization degree of 4-CP was obtained after reaction for 12 h in the dark. This is in good accordance with the above product analysis by HPLC-MS.
![]() | ||
Fig. 5 HPLC-MS chromatograms of the resulting solution after photocatalytic degradation of 4-CP on La2NiO4 crystals for different times under visible light irradiation and in the dark, respectively. |
Table 1 summarized the degradation efficiencies of La2NiO4 and TiO2−xNx (standard visible-light photocatalyst) for different organic pollutants under different conditions. Both La2NiO4 and TiO2−xNx showed high activity under visible light irradiation since they could be easily activated owing to the narrow energy band gaps. La2NiO4 exhibited a slightly lower activity than TiO2−xNx in 4-CP degradation due to its extremely low surface area (7.3 m2 g−1). Under visible light irradiation, La2NiO4 exhibited a higher activity for MO degradation than that for 4-CP degradation since the MO molecule was more active to be oxidized than the 4-CP owing to the presence of NN double bond (see the structural formula in Scheme 1). Similarly, the phenol degradation yield was higher than the 4-CP degradation yield on La2NiO4 under visible light irradiation because the C–Cl bond was quite stable against oxidative cleavage. In comparison with the 4-CP, phenol, and MO degradation, La2NiO4 showed the lowest activity in RhB degradation since the RhB molecule was even more stable against oxidation due to the conjugated π system (see the structural formula in Scheme 1). Interestingly, it was found that the activity of La2NiO4 in the dark was quite different from that observed under light irradiation. The degradation yield decreased in the order of 4-CP > MO > phenol > RhB. Meanwhile, one could see that 4-CP degradation on La2NiO4 in the dark was almost completely inhibited by either bubbling of the reaction solution with N2 flow to remove dissolved O2, or by introducing benzoquinone for trapping ˙O2− radicals, or by decreasing the pH value from 7.1 to 5.6 to inhibit the ionization of 4-CP into 4-CP− anions. Moreover, although TiO2−xNx showed high activity in 4-CP degradation under visible light irradiation, it was almost inactive in the dark. These results further confirmed that the degradation mechanism in the dark was different from that under visible light irradiation, taking into account that no photoelectrons and holes could be produced directly in the dark due to the absence of light irradiation.
![]() | ||
Scheme 1 The structural formula of MO and RhB molecules. |
Photocatalyst | Irradtiation | Reactant | Gas | pH | Time | Yield (%) |
---|---|---|---|---|---|---|
a Reaction conditions: 75 mg photocatalyst, 50 mL 5.0 × 10−6 M 4-CP aqueous solution, 30 °C. b BQ = benzoquinone (10 ppm). | ||||||
La2NiO4 | > 420 nm | 4-CP | Air | 7.1 | 4 | 45 |
La2NiO4 | > 420 nm | Phenol | Air | 7.1 | 4 | 61 |
La2NiO4 | > 420 nm | MO | Air | 7.1 | 4 | 88 |
La2NiO4 | > 420 nm | RhB | Air | 7.1 | 4 | 10 |
TiO2-xNx | > 420 nm | 4-CP | Air | 7.1 | 4 | 60 |
La2NiO4 | In dark | 4-CP | Air | 7.1 | 12 | 93 |
La2NiO4 | In dark | 4-CP | Air | 5.6 | 12 | 4 |
La2NiO4 | In dark | 4-CP | N2 | 7.1 | 12 | 9 |
La2NiO4 | In dark | 4-CP + BQb | Air | 7.1 | 12 | 7 |
La2NiO4 | In dark | Phenol | Air | 7.1 | 12 | 20 |
La2NiO4 | In dark | MO | Air | 7.1 | 12 | 70 |
La2NiO4 | In dark | RhB | Air | 7.1 | 12 | 12 |
TiO2−xNx | In dark | 4-CP | Air | 7.1 | 12 | ∼0 |
As shown in Fig. 6, the La2NiO4 could be used repetitively for more than 6 times without any significant decrease in activity during 4-CP degradation in the dark, showing excellent durability. More importantly, it also demonstrated that the 4-CP degradation reaction in the dark was not induced by the light originally stored in the La2NiO4 crystals.
![]() | ||
Fig. 6 Recycling test of the La2NiO4 photocatalyst for 4-CP degradation in the dark. Reaction conditions are shown in Fig. 4. |
Scheme 2 illustrated a plausible mechanism for the 4-CP degradation on La2NiO4 in the dark. Firstly, the 4-CP molecules ionized into H+ cations and 4-CP− anions in aqueous solution. Then, the 4-CP− anions interacted with the La2NiO4 to donate electrons to La2NiO4, together with the formation of 4-CP radical (4-CP˙). Similar to the photo-induced electrons, the as-trapped electrons could react with dissolved O2 to produce ˙O2− radicals, which subsequently reacted with H+ to form ˙OH radicals. These highly active ˙OH radicals could oxidize 4-CP˙ radicals into CO2, leading to 4-CP degradation. Obviously, the degradation efficiency of the La2NiO4 in dark was mainly dependent on the ability to trap electrons from organic pollutants.
![]() | ||
Scheme 2 A plausible mechanism of 4-CP degradation reaction in dark. |
There are several pieces of evidence which support the as-proposed mechanism. On one hand, the zeta potential spectra in Fig. 7 revealed that the La2NiO4 crystal in pure water was positively charged and turns to neutral in the presence of 4-CP molecules, which confirmed the above conclusion that the La2NiO4 crystal could trap electrons from the 4-CP− anions, together with the formation of 4-CP˙ radicals. On the other hand, the XPS spectra in Fig. 8 revealed that both the Ni(II) and La(III) species displayed negative shifts of binding energy after the La2NiO4 was pre-treated with 4-CP aqueous solution, which also confirmed that the 4-CP donated partial electrons to La2NiO4.
![]() | ||
Fig. 7 Zeta potential spectra of the La2NiO4 in H2O (green line) and 4-CP aqueous solution (red line). |
![]() | ||
Fig. 8 XPS spectra of La2NiO4 crystals before (a) and after (b) being pre-treated with 4-CP aqueous solution. |
The electrochemical measurements could further support the above mechanism. Fig. 9a shows the Tafel plots of pure and La2NiO4 modified carbon/glassy carbon (C/GC) electrodes in Na2SO4 aqueous solution with and without adding 4-CP. The equilibrium electric potential on the La2NiO4 modified C/GC electrode decreased from −0.094 to −0.120 V by adding 4-CP, which further confirmed the electron donation from 4-CP− to La2NiO4. On one hand, it enhanced the electric current indicative of 4-CP oxidation since more electrons released. On the other hand, it decreased electric current indicative of O2 reduction since partial electrons directly transferred from La2NiO4 to O2. Similar results could also be observed on the pure C/GC electrode with and without adding 4-CP. However, the C/GC electrode without La2NiO4 modification displayed much lower electric current indicative either the 4-CP oxidation or the O2 reduction than the La2NiO4-modified C/GC electron, suggesting that the La2NiO4 favored to trap electrons from 4-CP−. Meanwhile, the CV curves in Fig. 9b revealed that the O2 reduction peak potential on the La2NiO4 modified C/GC electrode shifted positively and electric current greatly enhanced compared with that of the pure C/GC electrode, suggesting that the presence of La2NiO4 crystals promoted the O2 reduction.29 This could be attributed to the layered perovskite structure of the La2NiO4 crystal which facilitated the O2 transfer. Furthermore, the CV curves in Fig. 9c revealed that the La2NiO4-modification promoted the 4-CP reduction on the C/GC electrode, suggesting that the La2NiO4 could more easily trap electrons from 4-CP− owing to the positive surface (Fig. 7).
![]() | ||
Fig. 9 (a) Tafel plots of unmodified and La2NiO4 modified C/C/GC electrode with and without 4-CP (1.0 mg mL−1), (b) CV curves of the C/GC electrode and La2NiO4–C/GC electrode in O2 saturated solution, and (c) CV curves of the C/GC electrode and La2NiO4–C/GC electrode with 4-CP (1.0 mg mL−1). All those experiments were conducted in the dark. |
To further confirm the above mechanism and account for the changes in degradation activity of La2NiO4 in the dark, the hydroxyl radicals (˙OH) were detected in the dark by measuring photoluminescence excited at 312 nm, with terephthalic acid (TA) as the ˙OH trapping agent.29Fig. 10 shows the intensities of photoluminescence emission peaks around 426 nm, which are characteristic of 2-hydroxyterephthalic acid (TAOH), obtained at 120 min under different conditions. La2NiO4 displayed a strong TAOH peak intensity in the 4-CP solution in the dark. However, almost no TAOH peak was observed when the solution was bubbled with N2 flow, suggesting that the ˙OH radicals were mainly generated from dissolved O2 rather than from H2O. Meanwhile, it was also found that the TAOH peak intensity abruptly decreased upon introduction of benzoquinone for trapping the ˙O2− radicals, which confirmed that the ˙OH radicals were mainly resultant from the reaction between ˙O2− and H+. In addition, an abrupt decrease in the TAOH peak intensity was found when the pH value decreased from 7.1 to 5.6, indicating that the 4-CP− anions played an important role for donating electrons and the high acidity might inhibit the 4-CP ionization to 4-CP−. The La2NiO4 exhibited a stronger TAOH peak intensity in the presence of 4-CP than that in the presence of MO, suggesting that the 4-CP molecule could donate electrons to La2NiO4 for forming ˙OH radicals more easily than the MO. This could accountfor the higher degradation yield of 4-CP than that of MO on La2NiO4 in the dark, although the MO molecule was more active to be degraded under light irradiation. An abrupt decrease in the TAOH peak intensity was found on La2NiO4 when using phenol instead of 4-CP, obviously due to the lower ionization degree of phenol (pKa = 9.99) than that of 4-CP (pKa = 7.65), leading to the lower ability to donate electrons to La2NiO4, which could account for the lower degradation yield of phenol than that of 4-CP. No significant TAOH peak was observed in the presence of RhB since RhB could ionize into RhB+ cations rather than anions, which exhibited a poor ability to donate electrons to La2NiO4, leading to the extremely low degradation yield of RhB on La2NiO4 in the dark. Although the TiO2−xNx exhibited strong TAOH peak intensity under visible light irradiation (λ > 420 nm) even in the absence of 4-CP, it showed no significant TAOH peak in the dark. Thus, it displayed almost no activity towards the degradation in the dark. From these results, we could conclude that the interaction between La2NiO4 and the anions of the organic pollutants with strong electron-donating abilities was essential for producing ˙OH radicals in dark. The traditional visible photocatalyst (TiO2−xNx) displayed a poor ability to trap electrons from organic pollutants and thus, could not produce ˙OH radicals in the dark, meaning that it could only work well under light irradiation.
![]() | ||
Fig. 10 ˙OH-trapping photoluminescence spectra with terephthalic acid. (a) La2NiO4–4-CP, (b) La2NiO4–4-CP–N2, (c) La2NiO4–4-CP–benzoquinone (10 ppm), (d) La2NiO4–4-CP at pH 5.6, (e) La2NiO4–MO, (f) La2NiO4–phenol, (g) La2NiO4–RhB, (h) TiO2−xNx–4-CP, (i) TiO2−xNx–4-CP under visible light (λ > 420 nm) irradiation. All the photoluminescence spectra were obtained at 120 min. Besides (i), all the other photoluminescence spectra were obtained in the dark. |
This journal is © The Royal Society of Chemistry 2012 |