Shahzad Abu Bakar*ab and
Caue Ribeirob
aDepartment of Chemistry, Federal University of São Carlos, Rod. Washington Luiz, km 235, CEP: 13565-905 São Carlos, SP, Brazil. E-mail: shahzad.158@gmail.com
bEmbrapa CNPDIA, Rua XV de Novembro, 1452, 13560-970, CP 741, São Carlos, SP, Brazil
First published on 5th September 2016
In this report, we describe the synthesis of nitrogen and sulfur co-doped TiO2 photocatalysts (NST) with preferential {001} facets by surfactant- and template-free OPM routes and crystallized through hydrothermal treatment. The precursor solution formed a coordination complex containing Ti–peroxo complex and chelating ligands (thiourea and urea were used as ligands for sulfur and nitrogen-donating sources, respectively). The effects of dopant concentration on the structural and morphological properties of the as-prepared NST photocatalysts were studied using X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), field emission scanning electron microscopy (FE-SEM), high-resolution transmission electron microscopy (HR-TEM), N2-adsorption–desorption isotherms and defuse reflectance spectroscopy (DRS). The visible-light-driven photocatalytic activity of the as-prepared NST samples was tested for the photodegradation of organic compounds (i.e. rhodamine B and phenol). High-performance liquid chromatography (HPLC) and LC-MS spectroscopy were used to analyze the by-products from the photodegradation of RhB. Different scavengers were added to the photocatalysis system in order to identify the role of active species in the photodegradation of organic compounds on the surface of NST photocatalysts. It was found that hydroxyl radicals (˙OH) and photogenerated holes (h+) played an important role in the photodegradation of organic compounds under visible-light irradiation. The electronic and structural characterizations of the as-prepared NST samples proved the successful incorporation of dopant elements (i.e. N and S) into the crystal lattice of TiO2 – which shifted the absorption edge shoulder from UV to the visible-light region, due to the bandgap transition. Another reason for the red shift of the absorption edge in the visible-light region is expected to be the formation of new energy levels near to the conduction bands, because of the incorporation of dopant elements (N and S) into the bandgap of the TiO2 crystal lattice. More interestingly, the pristine NST-0 sample showed photocatalytic activity, expected to be due to the formation of a substrate–surface complexation, resulting in an absorption shift of TiO2 into the visible light region due to the transfer of charges from the ligand (i.e. the attached dye molecule) to the titanium atoms. Moreover, the photocatalytic efficiency of the as-prepared NST samples was higher than those of the other samples. The prominent synergetic factors responsible for enhanced photocatalytic activity of doped NST samples include: (1) the presence of preferential exposed {001} facets of the anatase TiO2 nanorods, and (2) the introduction of the shift in the absorption edge shoulder towards the visible-light region. The preferential {001} facets act as a reservoir for the photogenerated charge carriers (i.e. electron–hole pairs) and slow down their rate of recombination. The introduction of absorption shift facilitates the adsorption of organic compounds on the photocatalyst surface.
By addressing different concerns and limitations in modifying the crystal structure of TiO2 by other routes, it has been established that non-metal doping and, in particular, non-metal co-doping is the most effective, efficient, inexpensive and stable route to achieve visible-light-driven photocatalytic activity of TiO2.17–20 In particular, non-metal co-doping not only enhanced the absorbance edge of TiO2 in the visible-light region but also facilitated the transition and/or separation of photogenerated charge carriers (i.e. electron–hole pairs).17,18,21–23 Traditional routes applied to the synthesis of N,S-co-doped TiO2 materials are either the use of any S- and N-donating ligands (urea and thiourea, etc.) for the sol–gel process24,25 or thermal annealing for TiO2 crystallization in the presence of the above-mentioned modifiers.26–28 The as-synthesized N,S-co-doped TiO2 material showed prominent photocatalytic activity for the degradation of water pollutants under visible-light irradiation, but improvement is still needed for long-term utilization of their photocatalytic applications.29 A general list of reported approaches includes the preparation of photocatalysts with a double-region-structure,30 decorating the surface of TiO2 with specific molecules,31,32 and adjustment of pH to control the surface electric charge of TiO2.33 It was found that the photocatalytic efficiency of the as-prepared materials was not sufficient because the reported methods were not adequate to drive the visible-light activity of TiO2.29–32 To the best of our knowledge, no one has cited the preparation of N,S-co-doped TiO2 nanorods using an OPM-assisted hydrothermal method.
While keeping these advantages in mind, we have developed a simple and more effective OPM-assisted hydrothermal route for the synthesis of N,S-co-doped TiO2 photocatalysts, utilizing previous knowledge from our cited research work.2,22,27 The resulting N,S-co-doped TiO2 material showed enhanced visible-light absorption thresholds due to expected band-to-band excitation and lower bandgap energy values as compared to the pristine sample. The as-prepared N,S-co-doped TiO2 nanorods showed good adsorption capacity and enhanced visible-light-driven photocatalytic activity for the degradation of organic compounds (i.e. rhodamine B and phenol). The as-prepared samples showed good recycling selectivity for the photodegradation of target pollutants. The enhanced photocatalytic activity of the as-prepared NST samples compared to the pristine sample has been studied by structural and morphological characterizations. Thus, this report may offer a promising, facile and environmentally benign approach to the design of highly efficient visible-light-activated photocatalyst material.
The photocatalytic activity of the as-prepared NST samples was tested for the photodegradation of organic pollutants (i.e. rhodamine B (RhB) and phenol) under visible-light irradiation. For this purpose, six Osram radiation lamps (Model = L18/10 – daylight intensity 18 W and maximum intensity at 440 nm) were used as the visible-light source with 25% of the total spectrum below maximum intensity. In detail, 10 mg of the as-prepared samples was dispersed into 20 mL of the 10 mg L−1 of the organic compound solution in beakers and these were placed in a homemade photoreactor, maintained at 18 °C by a thermostatically controller. The resulting mixture solutions were placed in the dark overnight to reach an adsorption–desorption equilibrium prior to visible-light irradiation. The detailed description of the homemade photoreactor can be found in our recent publication.37 UV-Vis spectroscopy (Shimadzu UV-1601PC) was used to monitor the photocatalytic oxidation reaction of the organic compounds for fixed intervals of visible-light irradiation. Different aliquots of the reaction solution were taken for fixed intervals of visible-light irradiation and were separated by centrifugation before performing the UV-Vis analysis. The formation of different by-products during the photocatalytic oxidation of RhB dye was monitored with the help of Mass Spectrometer (ESI-MS, Varian 310-MS) and high-pressure liquid chromatography instruments. The aliquots of the reaction solution were analyzed with the help of an Electrospray Ionization Spectrometer. A glass micro-syringe was used to inject the aliquots into the ESI source (20 mL min−1 flow rate) and the spectra were obtained after stabilization of the equipment.
Additionally, different chemical scavengers were added to the NST-containing RhB solution systems to monitor the photodegradation pathway. For this purpose, silver nitrate (AgNO3), dimethyl sulfoxide (DMSO), and sodium oxalate (SO) were added separately into the photocatalytic system as scavengers for conduction band (CB) electrons, ˙OH radicals, and photogenerated valence band (VB) holes (h+), respectively.38 The photoluminescence (PL) technique was used to monitor the generation of hydroxyl radicals (˙OH) on the interface of photoilluminated photocatalyst/water in the presence of a probing molecule (i.e. terephthalic acid). This technique was initially described by Fujishima and Ishibashi.39,40 The basic theme for the use of terephthalic acid (5 × 10−4 M of terephthalic acid in 2 × 10−3 M NaOH aqueous solution) as a probing molecule is its ability to produce a highly fluorescent product (2-hydroxyterephthalic acid) as a result of a rapid reaction with the freshly generated hydroxyl radicals in the solution.41,42 The change in PL peak intensity signal gives an indication of the terephthalic acid hydroxylation reaction with the ˙OH radical at the interface of photocatalyst/water. The concentration of hydroxyl radicals generated is directly proportional to the PL intensity of the fluorescent product (2-hydroxyterephthalic acid) in water.41,42 The experimental setup used for measurement of hydroxyl radical concentration is similar to the photocatalytic experiment. A Hitachi F-7000 fluorescence spectrophotometer was used to measure the PL spectra of the fluorescent product (2-hydroxyterephthalic acid). The mixture solution was exposed to visible-light for a fixed interval of time and then the PL intensity was measured.
![]() | ||
Fig. 1 The EDS spectrum of the NST-2 sample for better illustration in the marked region as shown in the inset of FE-SEM image. |
Sample | Phasea | Crystal sizeb (nm) | SBET (m2 g−1) | Pore volume (cm3 g−1) | Average pore size (nm) | Porosity (%) | Relative crystallinityc |
---|---|---|---|---|---|---|---|
a A stands for anatase phase.b Stands for the average crystallite size of anatase titania (TiO2) calculated using Scherrer equation from the broadening of the {101} diffraction peak.c Relative crystallinity of the anatase and calculated from the ratio of the relative intensity of the anatase {101} diffraction peak to the same diffraction peak for undoped sample. SBET, pore volume and pore size is calculated from the nitrogen adsorption–desorption isotherm as discussed in the Results and discussion section. | |||||||
NST-0 | A | 23.7 | 55.3 | 0.41 | 34.7 | 53.7 | 1 |
NST-1 | A | 22.3 | 59.2 | 0.47 | 31.3 | 57.4 | 0.89 |
NST-2 | A | 21.8 | 63.7 | 0.52 | 27.4 | 59.2 | 0.85 |
NST-3 | A | 21.5 | 65.2 | 0.56 | 25.9 | 60.5 | 0.83 |
Fig. 4 shows the FE-SEM images of the as-prepared NST samples. From the FE-SEM images, it was revealed that the pristine sample has the surface morphology of a nanorod shape. The increase in dopant concentration results in the evolution of an anisotropic nanostructure to form an isotropic structure, as depicted in the FE-SEM images. The structure of the as-prepared samples consisted of nanorods with a regular shape with an average length and diameter in the range of ca. 60–80 nm and ca. 20–30 nm, respectively. The elemental composition and surface mapping of the as-prepared samples are shown in Fig. 2S and Table 1S.† More interestingly, an increase in dopant concentration did not change the surface morphology of the as-prepared samples. The nanorod-shaped morphology of the samples was also revealed with the help of TEM analysis. Fig. 5(a) and (b) show the typical TEM images of the NST-0 and NST-2 samples – to observe the surface morphologies the samples were prepared by the OPM route via hydrothermal treatment (see Experimental section). The as-prepared samples have a nanorod shape with anisotropic structures. The length and diameter of the nanorods are in accord with the values calculated from FE-SEM analysis. From the HR-TEM images, it was revealed that the length of the nanorods decreases to a marginal value with an increase in dopant concentration, whereas a slight difference in the diameter values is also observed for the higher concentration of dopant level. This suggests that nanorods are made by the restructuring and fusion of different original close-contact TiO2 crystalline lattices. These findings are in good agreement with the isotropic structure from the FE-SEM and XRD results. Fig. 5(c) and (d) show the HR-TEM and Fourier Transform (inset) images of the as-marked region, which reveal that the monocrystalline nature of the as-prepared samples is retained for the higher concentrations of dopant impurity. Close analysis of the HR-TEM and FT analyses revealed that no secondary phase was present and even the local agglomeration of dopant impurities did not happen (the formation of nitrogen oxide or sulfur oxide particles). Thus, these findings further support the earlier discussion that dopant atoms (N and S) are incorporated into the crystal lattice of TiO2 without forming any extra impurity phase and is true even for the higher concentration of dopants. Similar findings were observed for the XRD results of all the as-prepared samples.
![]() | ||
Fig. 4 FE-SEM images of the as-prepared samples: (a) & (b) NST-0 (pristine); (c) NST-1; (d) & (e) NST-2; (f) NST-3. |
![]() | ||
Fig. 5 TEM (a and c) and HRTEM (b and d) of the NST-0 and NST-2 samples. Inset shows the respective Fourier Transform (FT). |
HR-TEM and FT results show that NST-0 nanorods have grown in the direction of [001] as the lattice spacing parallel to the exposed facets is ca. 0.235 nm from top to bottom, as discussed in the XRD Results and discussion section. Moreover, the isotropic nanostructures have the small crystallite size of nanorods and preferential facets of (101) for the NST-2 sample. One factor for this abnormal change in preferential facets is the change in the surface energy for different planes of crystallite.46,47 The change in surface energy of crystalline planes is responsible for the predominant evolution of the nanorod morphology with the incorporation of dopant level. The average values of surface energy reported for the TiO2 anatase phase are 0.53 J m−2 for {100}, 0.44 J m−2 for {101}, and 0.90 J m−2 for {001}, respectively.46,48,49 Thus, the preferentially exposed facet for the anatase phase of TiO2 is high-energy {001} as expected from the cited values and is noted to be the most stable phase for thermal treatment. Conversely, HR-TEM analysis revealed that the incorporation of dopant atoms (N and S) into the TiO2 crystal lattice preferred exposed (001) planes either without thermal annealing of the material for high temperature or without the post-calcination step. One can predict from these findings that the existence and incorporation of dopant atoms have a significant impact on the structure and morphology of the materials by either affecting the mechanism of crystal growth or changing the surface energy.47,50 Therefore, our findings support the statement that dopant atoms (N and S) have played a significant role in the mechanism involved in crystal growth and controlling the morphology of the nanostructure materials.
Fig. 6 shows the nitrogen adsorption–desorption isotherms of the NST-0 and NST-2 samples, while the inset shows the pore diameter distribution of the pore volume of the as-prepared samples. Similar patterns were observed for the isotherms and pore size distributions of other samples (not shown here). The presence of large mesopores (i.e. a particle gap range between 2 and 50 nm) was indicated by the existence of type IV isotherms for the NST-2 sample with very narrow hysteresis loops close to unity at relative pressures. Moreover, type II isotherms were observed for nitrogen isotherms of the adsorption branch which steadily increases on approaching unity in relative pressure (P/P0) due to the existence of mesopores (>100 nm) with a high external surface area. As shown in the inset of Fig. 6, the pore volume distribution curve is broad in the range of 20 to over 70 nm which further supports the presence of mesopores. From the agglomeration of nanoparticles resulting in the formation of these mesopores it is evident that single-crystals of nanorods have nanoporous channels.22 For photocatalytic applications, this nanoporous channel structure is reported to be useful for efficient and fast transfer of reactant and product molecule fragments.22 The specific surface area, porosity, and pore volume of the as-prepared samples are shown in Table 1. An increase in dopant concentrations results in a decrease in average pore size due to the conquest of the crystal growth mechanism.
![]() | ||
Fig. 6 Nitrogen adsorption–desorption isotherm for NST-0 and NST-2, and inset shows the corresponding pore volume distribution curve for NST-2 sample. |
The chemical composition and chemical state of the dopant elements (N and S) in the lattice of TiO2 were identified with the help of XPS analysis. For this purpose, survey spectra of NST-0 and NST-2 samples are shown in Fig. 7(a). It was noted that the survey spectrum of pure TiO2 has sharp photoelectron peaks for Ti, O and C elements corresponding to the binding energies at 459.2 (Ti 2p), 530.7 (O 1s) and 284.8 eV (C 1s). The nominal atomic composition for anatase TiO2 was observed for the NST-0 sample from the XPS survey spectrum with an atomic ratio of Ti to O which is about 1:
2 and is in good agreement. More interestingly, the adventitious carbon peak is reported to arise from the XPS instrument grid. Whereas, for the NST-2 sample, minor XPS photoelectron peaks corresponding to dopant elements (N and S) were present at binding energies of at 401 (N 1s) and 168.7 eV (S 2p), beside the main Ti, O, and C photoelectron peaks.
The high-resolution XPS spectra of the C 1s photoelectron peaks for NST-0 and NST-2 samples are shown in Fig. 7(b). It was found that the C 1s photoelectron peaks are similar in shape and appeared at the same binding energies. More importantly, the symmetrical C 1s peak at a binding energy of 284.8 eV is ascribed to the presence of methyl groups (organic compound) and is expected to arise from the XPS instrument itself. In particular, no clue was found for the presence of either anionic or interstitial C 1s peaks corresponding to the binding energies at 282 and 288 eV for C-doped TiO2 material and thus eliminated the chance of carbon doping into the TiO2 crystal lattice.51
The high-resolution XPS spectrum was taken from the surface of the NST-2 sample and is shown in Fig. 7(c) for the deconvolution the N 1s photoelectron region. The absence of a sharp N 1s peak at 400.6 eV is attributed to the fact that the adopted synthesis method is superior to a number of previously reported methods to obtain N,S-co-doped samples without performing the post-annealing treatment. It also supports the fact that nitrogen atoms are not present in the oxidized form of Ti–N–O in the crystal lattice of TiO2.52,53 Whereas the presence of a sharp XPS signal at BE of 397.0 eV for the N 1s peak is attributed to the anionic N-doping in the TiO2 lattice forming O–Ti–N bonds because of the replacement of oxygen atoms by nitrogen atoms in the TiO2 crystal lattice. Thus, interstitial nitrogen atom-doping did not take place and only anionic nitrogen-doping occurred. It is reported in the literature that preparation routes control the localization of nitrogen atoms in the crystal lattice of TiO2 for N-doped TiO2 samples.54 Therefore, the form of nitrogen doping in the TiO2 crystal lattice largely depends on the synthesis route and the source of N-doping reagents.52 More interestingly, oxygen-rich conditions favored interstitial N-doping and the absence of oxygen (anaerobic condition) or the presence of a reducing environment favor the substitution of oxygen atoms by nitrogen atoms in the TiO2 crystal lattice. Thus, the absence of interstitial N-doped TiO2 is not surprising due to the absence of anaerobic atmosphere in the OMP route.
The high-resolution XPS spectrum of S 2p photoelectrons for the NST-2 sample is shown in Fig. 7(d). It was observed that the XPS spectrum consists of two types of peak for S species in the region of S 2p. The presence of a prominent XPS peak at about 168.7 eV corresponds to the S6+ state of the S species in the TiO2 crystal lattice. Whereas the trace of a minor peak at 163.8 eV corresponds to the presence of some sulfur atoms as anionic doping in the state of S2− in the crystal lattice of TiO2 due to the replacement of oxygen atoms. In detail, the cationic substitution of titanium atoms by sulfur atoms in the lattice is attributed to a rise in the XPS peak at 168.7 eV. On the contrary, the presence of a weak peak at 163.8 eV is attributed to the formation of some Ti–S bond.55,56 Similar findings were reported by Umebayashi et al.57 for anionic sulphur-doping, where titanium disulfide (TiS2) was used as a starting material. It was reported that during the preparation process most of the sulfur atoms were oxidized from TiS2 and remained as residual sulfur in the form of S2−. More importantly, the large ionic radius of S2− (1.7 Å) as compared to O2− (1.22 Å) make it difficult for the successful substitution of oxygen atoms by a sulfur atom in the lattice due to the requirement of large activation energy for the formation of a Ti–S bond. On the other hand, the lower value of the activation energy for the cationic substitution of Ti4+ by S6+ in the lattice makes it more feasible compared to the replacement of oxygen atoms by sulfur atoms.55 The XPS findings revealed that the exact amount of dopant elements (N and S) in the NST-2 sample are in accord. Furthermore, the XPS results confirmed that an increase in dopant concentration results in an increase in their atomic percentage in the lattice TiO2.
The optical properties of the selected samples were studied with the help of normalized absorbance data. Generally speaking, non-metal-doping into the TiO2 crystal lattice changes the light absorption pattern of TiO2.56,58 It was noted that incorporation of dopant atoms into the TiO2 crystal lattice not only changes the nanostructure morphology and composition, but also changes the bandgap value of the nanorods. Fig. 8 shows the onset of the shift in absorbance band edge for the as-prepared NST samples. The inset of the Fig. 8 shows the bandgap plot using the Kubelka–Munk function from the indirect transition for the as-prepared NST samples.59–61 In particular, the NST-2 sample shows two prominent features: (1) a new shift in the absorption shoulder to the visible wavelength region (400–500 nm) and (2) the absorbance enhanced in the range of 550–650 nm. The overall absorbance in the visible-light region increases with the increase in dopant concentration and is consistent with the appearance of a yellow color in the NST samples. It was reported that interstitial N-doping into the TiO2 crystal lattice resulted in a narrowing of the valence band edge by creating some localized states due to the introduction of a visible-light absorption shoulder in the 380 to 500 nm wavelength range.53 The synergistic effect of co-dopant elements (N and S) changed the TiO2 electronic structure and enhanced the absorbance shoulder in the visible region i.e. 500–750 nm. Whereas nitrogen doping is directly responsible for the absorbance shoulder in the wavelength range of 380–500 nm. The calculated bandgap value for the NST-2 sample was in the range of 2.70 eV (460 nm). The incorporation of dopant atoms results in a shift of the bandgap edge towards the longer wavelength of the visible-light region. The expected transition by impurity levels is related to the shift in bandgap edge as reported earlier for the incorporation of nitrogen and sulfur atoms into the TiO2 crystal lattice having a wider bandgap.62,63 These findings further support the successful incorporation of nitrogen and sulfur atoms into the TiO2 crystal lattice, which changes the crystallite size and electronic structure of the as-prepared NST samples.
The photocatalytic activity of the as-prepared NST samples was investigated for the degradation of organic compounds (i.e. rhodamine B (RhB) and phenol) in an aqueous solution beneath visible-light irradiation. In order to ensure the photocatalytic degradation of organic dyes over the as-prepared NST photocatalysts, the change in the organic compound concentration over the NST samples was measured under dark conditions, in the absence of visible-light irradiation. Fig. 3S† shows the change in RhB concentration over the as-prepared NST samples. It was found that the RhB concentration did not change to a significant amount under dark conditions. On the other hand, the RhB concentration decreased dramatically under visible-light irradiation. Therefore, the degradation of RhB molecules was performed over the surface of the NST photocatalysts under visible-light irradiation. The photocatalytic activity of the as-prepared NST samples was calculated by measuring the decrease in the RhB concentration over the different intervals of visible-light irradiation. The following eqn (1) was used to calculate the photodegradation rate constant:
![]() | (1) |
The change in the shape of UV-Vis absorbance spectra helps to analyze the dominance of the photodegradation routes – either by surface or bulk reaction. In the case of the bulk solution reaction, the RhB molecules degraded without changing the position/shift of the RhB absorbance peak. Whereas the involvement of the surface reaction changes the position of the UV-Vis peak during the degradation of the RhB aqueous solution. The shift in absorbance spectra for the as-prepared NST samples was more prominent after the visible-light irradiation. The rate of surface reaction for the as-prepared NST-2 sample was much higher than for other samples. In particular, the rate of surface reaction for the as-prepared NST-2 photocatalysts is higher due to the occurrence of rapid surface reactions as compared to other samples, where the surface reactions occurred slowly. The UV-Vis absorbance spectra for RhB aqueous solution over the NST-2 sample started to decrease under visible-light irradiation. The decrease in absorbance intensity is due to the degradation of RhB molecules over the photocatalysts under visible-light irradiation via both fast bulk solution and surface reactions.
To eliminate the chance of a photosensitization process, the photocatalytic activity of the as-prepared samples was investigated for the degradation of a colorless organic compound (i.e. phenol) under visible-light irradiation. It was noted that no change in phenol concentration occurred over NST samples in dark conditions (without visible-light irradiation). Whereas a blank photocatalytic test (in the absence of photocatalysts) showed that no photodegradation of phenol took place without the addition of photocatalysts. Therefore, the presence of both NST photocatalysts and visible-light illumination is required for the successful photodegradation of phenol aqueous solution. It was found that photocatalytic oxidation reactions for the degradation of phenol aqueous solution take place on the surface of the photocatalysts under visible-light irradiation. The change in absorption of phenol aqueous solution over the as-prepared NST-2 photocatalyst under visible-light irradiation is shown in Fig. 4S.† The photocatalytic degradation of phenol and rate constants over different photocatalysts under visible-light irradiation are shown in Fig. 9(c) and (d). It was observed that the molar ratio of titanium to the dopant concentration has a significant effect on the photocatalytic efficiency of NST samples. More interestingly, an obvious photocatalytic activity was noted for the pristine TiO2 sample under visible-light irradiation. Whereas it is theoretically established that anatase TiO2 has a large bandgap of 3.2 eV and cannot be excited under visible-light irradiation. Therefore, it is very interesting to figure out the main reason for the photocatalytic activity of pristine TiO2 towards organic compound degradation under visible-light irradiation. According to the literature, the surprising visible-light photocatalytic activity of TiO2 is expected to arise from substrate–surface complexation between the photocatalyst (TiO2) and an organic compound. The creation of substrate–surface complexation is involved in the transfer of charge from ligand-to-titanium atoms and extends the absorbance edge of titania in the visible-light region.65,66 This extension of the absorbance edge in the visible-region is explained on the basis that the adsorption of organic compounds on the surface of TiO2 is responsible for absorption in the visible region and TiO2 itself did not absorb visible-light.65,66 Hence, during the photocatalytic experiment, the adsorption of an organic compound on the surface of TiO2 is responsible for shifting the absorption edge in the visible-light region due to an intramolecular transfer of charge from ligand to titanium atoms owing to the formation of a substrate–surface complexation. The photodegradation of organic compounds (i.e. phenol) over a TiO2 photocatalyst under visible-light irradiation through substrate–surface complex mediation is presented in Fig. 10(a). Serpone et al.66 described the surface–substrate complex formation through phenolate linkage during the adsorption of ligand on the surface of TiO2 in the TiO2/phenol system as:
![]() ![]() | (2) |
The substrate–surface complex is excited upon visible-light irradiation through the transfer of charge from ligand to titanium. As a result, the photo-induced electrons started to shift to the TiO2 conduction band. Afterwards, the adsorbed O2 on the TiO2 surface reacts with injected electrons to form superoxide radical anions (˙O2−). The superoxide radicals react further by protonation to form ˙OOH radicals which trapped the injected electrons and subsequently form H2O2 and ˙OH radicals. It is well-known that ˙OH radicals are the most active species for the degradation of organic compounds.67,68 On the other hand, the photocatalytic activity of all the NST samples is much higher compared to the pristine sample under visible-light irradiation. The proposed mechanism involved for the higher photocatalytic activity of NST samples as compared to the pristine sample is shown in Fig. 10(b). Furthermore, the photocatalytic activity of NST samples increases with the increase in dopant concentrations and reaches a maximum value for the NST-2 sample. The photocatalytic efficiency of NST-2 is higher than for the other samples.
The formation of intermediate by-products during the degradation of organic pollutants was monitored with the help of HPLC and LC-MS analysis. The HPLC analysis was performed during the photocatalytic degradation experiment of RhB aqueous solution over NST-2 photocatalysts for the different intervals of visible-light irradiation. Fig. 11(a) shows the matrix peaks at retention times of tR1/4 = 8.5 min of RhB at 554 nm. The evidence for the photodegradation of RhB was obvious from the diminishing of the matrix peak with photocatalytic experiment time. The complete photodegradation of RhB into less-toxic fragments was achieved at the end of the photocatalytic experiment. In particular, no peak corresponding to the presence of any by-product was found – except for the RhB. The LC-MS analysis was performed to further confirm the preliminary degradation process of RhB over NST photocatalysts under visible-light irradiation. Fig. 5S† shows the LC-MS spectra for the pattern formation of different fragments/intermediates during the photocatalytic degradation of RhB under visible-light irradiation. The molecular ion peak for RhB started to decrease and diminished completely in the LC-MS analysis after the completion of the photocatalytic experiment. In detail, the intensity of the higher m/z value peaks started to decrease due to the degradation of RhB into non-toxic levels of inorganic products (i.e. CO2, H2O, etc.). Thus, this confirmed the complete degradation of RhB under visible-light irradiation.
Excellent reproducibility was observed for the NST-2 sample over 4 consecutive cycles of the photocatalytic experiments. Fig. 6S† shows the 4 cycles of photocatalytic degradation of RhB over the NST-2 sample under visible-light irradiation. The surface bulk solution reactions for the as-prepared samples were faster than for the reference samples. The high visible-light absorbance for the as-prepared photocatalysts was expected to bring about the rapid bulk solution reactions. The rate of surface reaction depends on the concentration of the doping elements in the as-prepared samples. It has been reported that, other than the doping element concentration in the TiO2 crystal structure, the variation in the crystal structure, degree of crystallinity, specific surface area and the chemical states of the dopants on the TiO2 surface have a considerable effect on the rate of surface reactions.69
Fig. 7S(a)† shows the rate constant of ˙OH radical formation as a function of visible-light irradiation time. It was found that the rate constant for ˙OH radical formation follows the same trend as that observed for the photodegradation of organic pollutants. This shows that an indirect mechanism played a key role in the photocatalytic oxidation reactions. The detection of ˙OH was performed to confirm the photocatalytic activity of as-prepared samples under visible-light irradiation. Fig. 7S(b)† shows the change in PL spectra for the terephthalic acid solution as a function of visible-light irradiation time. It was noted that the PL intensity of terephthalic acid solution increases gradually with time. Whereas no change in PL intensity was noted in the absence of either NST photocatalysts or visible-light irradiation. Therefore, chemical reactions between the generated ˙OH radicals and the terephthalic acid are responsible for the rise in fluorescence intensity. The change in PL intensities is linear over the as-prepared samples for a given interval of time. The concentration of ˙OH radical formation is directly proportional to the visible-light irradiation time. The photogenerated holes possessed enough oxidation power to produce ˙OH radicals by interacting with the surface-adsorbed hydroxyl ions/water molecules. The photogenerated electrons in the conduction bands participated in the redox reactions to produce ˙OH radicals on the surface of NST photocatalysts.70,71 It was observed that the PL intensity for the NST-2 sample was higher than for other samples i.e. higher amounts of ˙OH radical were produced over the NST-2 sample under visible-light irradiation. Therefore, higher photocatalytic activity was observed for the NST-2 sample. For pristine TiO2, the intensity of the PL peak is very low, as expected due to the fact that TiO2 is not activated under visible-light irradiation. As evidence, the concentration of generated ˙OH on the surface of photocatalysts played an active role in the photocatalytic efficiency of the as-prepared samples. Therefore, the more hydroxyl radicals produced, the higher the photocatalytic activity.
A detailed study was performed to characterize the mechanism involved in the photocatalytic process. The involvement of active species was studied in the photocatalytic degradation of the organic pollutant (RhB) over the NST-2 sample under visible-light irradiation – by adding different chemical scavengers into the reaction systems. For this purpose, AgNO3 (a CB electron acceptor plus a strong oxidant), DMSO (an ˙OH scavenger) and sodium oxalate (a photogenerated holes (h+) near-VB scavenger) were added into the photocatalysis system.38,72,73 Furthermore, the dissolved oxygen is responsible for inhibiting the recombination of charge carriers due to the formation of superoxide radicals (˙O2−) – which played an important role in some of the photocatalytic processes.74 As a rational approach, if the photocatalytic process is performed by superoxide radicals – as an active species, the photocatalytic reaction rate would be reduced substantially with the addition of Ag+ into the reaction system.72,73 On the contrary, the reduction of reaction rate with the addition of either DMSO or SO showed that ˙OH radicals and/or holes played active roles in the photodegradation process.72,73,75 Fig. 11(b) shows that the addition of sodium oxalate (0.5 mmol L−1) reduced the photodegradation efficiency of RhB from 78 to 48% as compared to the experiment in the absence of scavengers. This shows that h+ played an active role in the photodegradation of RhB aqueous solution under visible-light irradiation. Similarly, an excess of DMSO (5 mmol L−1) was added into the reaction system to monitor the effect of ˙OH formation on the photodegradation processes. The excess DMSO captured all the ˙OH during the photodegradation reaction of RhB aqueous solution.75 It was found that the photodegradation efficiency of RhB reduced greatly from 78 to 42% after 2 hours of visible-light irradiation. Thus, ˙OH played a more important part in the photodegradation of RhB as compared to the photogenerated holes. Overall, photogenerated h+ and ˙OH played an important role in the photodegradation of RhB aqueous solution over the NST photocatalysts under visible-light irradiation. These findings further support the results obtained for the formation of ˙OH from the terephthalic acid experiments. On the contrary, the addition of excess Ag+ (1 mmol L−1) did not change the photodegradation efficiency of RhB to an obvious degree. This shows that O2−˙ played a minor part in the photocatalytic degradation of RhB aqueous solution under visible-light irradiation.
The synergetic effects of enhanced photocatalytic activity for NST samples are ascribed to the presence of red-shift of intense absorption edges in the visible-light region and the presence of highly reactive preferential {001} facets. As discussed earlier, the introduction of dopant elements into the TiO2 lattice shift the absorption edge in the visible-light region. Thus, the NST samples were activated under visible-light irradiation and generated photo-induced charge carriers (i.e. electrons and holes pairs) to participate in the photocatalytic oxidation reactions. Furthermore, the presence of preferential highly reactive exposed {001} facets as compared to thermodynamically stable {101} facets dominates the active site on the surface of the photocatalysts for photocatalytic oxidation reactions.76,77 As a result, the 1-D nanorods with preferential {001} facets showed enhanced visible-light photocatalytic activity due to more feasible adsorption of organic pollutants on the surface of the photocatalysts for degradation.78 Moreover, most of the time anatase TiO2 nanocrystals are prepared with preferential {101} facets and showed lower photocatalytic activity. Therefore, the higher photocatalytic activity of as-prepared NST nanorods was obtained with preferential exposed {001} facets under visible-light irradiation. More interestingly, the NST-2 nanoparticles with less dominant {001} facets prepared in another experiment showed lower photocatalytic activity under visible-light irradiation as compared to NST-2 nanorods having preferential exposed {001} facets. Thus, we can conclude from these findings that the morphology of TiO2 played an important role in the photocatalytic performance of TiO2 nanorods having preferential {001} facets under visible-light irradiation as compared to corresponding nanoparticles. In the end, a further increase in dopant element concentration results in a gradual decrease in photocatalytic activity. This gradual decrease in photocatalytic activity is expected due to the deterioration in the degree of crystallinity and creation of more defects sites which act as recombination centers in the NST samples (see Table 1).
Another explanation for the enhanced photocatalytic activity of the as-prepared samples is based on the presence of dopant elements on the surface of NST photocatalysts – which efficiently enhanced the rate of surface reaction due to the enhanced interaction/adsorption of the organic dyes on the surface of the NST photocatalysts. The strong interaction between the dopant elements and the organic compounds promotes the adsorption and rate of the surface reaction of NST photocatalysts under visible-light irradiation. The photoinduced holes from the conduction band – i.e. those excited to the surface of the photocatalysts by visible-light irradiation – are adsorbed by the surface adsorbed oxygen molecules to produce hydroxyl radicals. The nascent hydroxyl radicals facilitate the surface reactions for the degradation of organic dyes on the surface of NST photocatalysts. The presence of delocalized states in the bandgap of the NST samples promotes the rate of bulk solution reaction by shifting the absorbance edge of the NST sample in the visible-light region. Thus, the as-prepared NST photocatalysts showed enhanced photocatalytic activity for the degradation of organic compounds under visible-light irradiation. The enhancement of dopant concentration after an ambient concentration reduced the photocatalytic efficiency due to the lower rate of surface and bulk solution reactions – as observed for the NST-3 sample.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra16881g |
This journal is © The Royal Society of Chemistry 2016 |