Yijun Yanga,
Dawei Nib,
Ye Yaob,
Yeteng Zhonga,
Ying Ma*b and
Jiannian Yao*a
aBeijing National Laboratory for Molecular Science, Key Laboratory of Photochemistry, Institute of Chemistry, Chinese Academy of Sciences, Beijing, 100190, People's Republic of China. E-mail: jnyao@iccas.ac.cn
bState Key Laboratory of Material Processing and Die & Mould Technology, School of Material Sciences and Engineering, Huazhong University of Science and Technology, Wuhan, 430074, People's Republic of China. E-mail: yingma@hust.edu.cn
First published on 26th October 2015
Carbon doping has been widely applied to modify TiO2 to improve photocatalytic activity and initiate visible light activity. Using oleylamine wrapped TiO2 nanomaterials as precursor, carbon doped TiO2 photocatalysts have been synthesized by control of heating ramp rates and temperatures in air. Optical absorption of all these photocatalysts are extended to visible light, and photogenerated electron–hole separation is enhanced by carbon doping. Especially, those prepared by fast combustion of oleylamine ligands exhibit excellent photocatalytic activity and visible light activity for hydrogen production. EPR analysis demonstrates that more oxygen vacancies present in carbon doped TiO2 with high activity. This indicates that oxygen vacancies may play key roles in photocatalytic water splitting. Fast heating process may have offered an oxygen-poor atmosphere in which oxygen vacancies are favoured.
In addition to nitrogen,10,11 carbon has also been well accepted as an efficient doping element to make TiO2 sensitive to visible light. Khan12 et al. firstly reported that TiO2−xCx (x ∼ 0.15) absorbed light at wavelengths below 535 nm and performed water splitting at an applied potential of 0.3 V. Sakthivel and Kisch found good photocatalytic property of carbon doped TiO2 in degradation of 4-chlorophenol under diffuse indoor daylight.13 Various methods have been developed to synthesize carbon doped TiO2 and enhanced visible light activity of the doped photocatalyst has been verified by many groups subsequently.14–18 Among these methods, heating precursor containing carbon species is the most commonly used one for doping TiO2 with carbon.19–21 However, the existing states of carbon in these photocatalysts have triggered a lot of arguments. Carbon was proposed to substitute a lattice atom in some reports,12,15,22 while interstitial carbon atoms were believed to dominate in other works.16,23 Recently, Kisch and co-workers have attributed the visible light activity of some “carbon-doped” TiO2 to the sensitization of TiO2 by aromatic carbon compound.24 Moreover, no visible activity or detrimental effect of C doping on photocatalytic activity of TiO2 under UV light was observed in some cases.25–27
Herein we report an easy way to fabricate carbon doped TiO2 with high photocatalytic activity for hydrogen production via fast combustion of organic capping reagents. During surface modification of TiO2 with carbon via heating oleylamine wrapped ultrathin TiO2 nanosheets,28 we noticed that elevating calcination temperature or heating rate in air would dope carbon into TiO2 lattice. By varying the temperature ramp-up rates, we prepared several carbon doped TiO2 photocatalysts. A high and stable hydrogen generation rate was observed on those prepared under a high heating rate, which benefits fast combustion of oleylamine ligands. In contrast, those prepared under a low heating rate exhibited a relative low activity under UV light and no activity under visible light. We think fast and slow combustion of oleylamine ligands may lead to different carbon doping states in the final products.
To better understand the key factors leading to such a difference in activity among these carbon doped TiO2 photocatalysts, detailed structural analyses were carried out. The crystalline structures of the photocatalysts were confirmed by powder XRD patterns (shown in Fig. 2a). The peaks at 25.52°, 48.01°, 53.96°, 55.04° and 62.68° can be assigned to the (101), (004), (200), (105), (211) and (204) planes of anatase (space group: I41/amd; tetragonal symmetry, a = 3.7852 Å, c = 9.5139 Å, JCPDS card no. 21-1272), respectively.29 No diffraction peaks belonging to potential impurities, rutile or brookite can be discerned. According to the half-width at half maximum (FWHM) of (101) peaks of F-400-1, F-500-0.5, N-300-450 and N-300-500, the crystallite size of anatase estimated using the Scherrer equation are about 11.0, 12.4, 13.2, 14.2 nm, respectively. The crystallite size slightly increases with the rise of calcination temperature and the prolongation of calcination time.
![]() | ||
Fig. 2 (a) XRD patterns, (b) IR spectra, (c) Raman spectra and (d) Raman shifts of Eg modes of the photocatalysts. |
Fig. 2b shows the FTIR spectra of the photocatalysts. The characteristic absorption bands of adsorbed water molecules including stretching vibration ranging from 3200 to 3400 cm−1 and bending vibration at around 1633 cm−1 are clearly visible in all samples, indicating that water is adsorbed on their surface.30 The strong absorption band at ca. 462 cm−1 originates from the vibration of Ti–O bonds, and the absorption bands at 1384 cm−1 can be assigned to O–H in-plane deformation. The latter becomes weaker when high calcination temperature or long calcination time was adopted. The emergence of absorption bands at 2923, 2853 and 1463 cm−1, assigned to νas (–CH2), vs. (–CH2) and δas (–CH2), suggests the existence of trace organic residue in sample F-500-0.5 due to short time combustion.31 Raman spectroscopy is a powerful non-destructive technique for the investigation of the crystalline quality of photocatalysts. Six typical Raman-active vibrational modes of anatase have been reported including modes A1g (513 cm−1), B1g (399 and 519 cm−1), and Eg (144, 197 and 639 cm−1).32 All the Raman peaks shown in Fig. 2c correspond to these vibrational modes, indicating that anatase is the predominant phase structure in the photocatalysts. The blue shift in Eg modes of F-500-0.5 and F-400-1 (Fig. 2d) may be attributed to more defects generated in fast combustion and crystallization process.
To investigate the chemical states of carbon atoms incorporated into the TiO2 photocatalysts, X-ray photoelectron spectroscopy (XPS) was also utilized to record the Ti 2p, O 1s and C 1s elements of the photocatalysts. Similar XPS spectra were obtained for all the photocatalysts including P25 and typical spectra are shown in Fig. 3 for clarity. As shown in Fig. 3a, the survey spectra of all photocatalysts exhibit all peaks of elements of Ti, O and C. In Fig. 3b, two intense symmetric peaks at 458.4 and 464.2 eV in Ti 2p spectra of F-400-1 and N-300-450 can be ascribed to Ti 2p3/2 and Ti 2p1/2, respectively. The ΔE value between Ti 2p3/2 and Ti 2p1/2 was about 5.8 eV, indicating that the element of Ti in photocatalysts is predominantly Ti4+.33 The main peak at 529.6 eV in O 1s spectra of photocatalysts (Fig. 3c) can be assigned to O in the form of the O–Ti bond (lattice O), while the peak at 531.4 eV can be assigned to CO bond (and COO).34 The peak at around 533.1 eV of N-300-450 and P25 (not shown here) may be caused by hydroxyl groups and chemisorbed water.35,36 For the C 1s spectra shown in Fig. 3d, peaks at 284.8, 285.7 and 288.8 eV can be attributed to the C–C neutral bond, C–OR(H) group and COO or COOR(H) group, respectively,34 since nearly the same C 1s spectrum was obtained for all samples including P25. These oxidised carbon species originate from the incomplete combustion of organic molecules in the TiO2–oleylamine precursor or adventitious hydrocarbonaceous and carbonate species.18 No peak at 281.8 eV can be discerned, as usually observed for carbon-doped TiO2 prepared by heating a carbon-containing precursor.19–21 This cannot exclude oxygen substitution of carbon to form a Ti–C bond because this peak often appears after surface cleaning by Ar+ sputtering.18,37 Although the carbon doped photocatalysts exhibit different O 1s XPS spectra, the different surface oxygen species seem to have little effect on their photocatalytic activities for hydrogen evolution. Both N-300-450 and N-300-500 show low activity, but surface oxygen species related to 533.1 eV was only observed in N-300-450. In addition, weak signals of N 1s were detected (Fig. S1†) and the nitrogen contents were estimated by XPS analyses to be 0.66, 0.79, 0.54 and 0.49 atom% for F-400-1, F-500-0.5, N-300-450 and N-300-500, respectively. The carbon doping contents of F-400-1, F-500-0.5, N-300-450 and N-300-500 are about 1.65, 1.78, 1.60 and 1.48 atom%, respectively, determined by EPMA. No nitrogen contents were detected by EPMA, indicating trace nitrogen exits only on the surface of these samples.
According to the above XPS spectra and other structural analyses, no obvious structural difference seems to contribute to different photocatalytic behaviour observed on the samples prepared by slow and fast combustion processes. The morphologies of carbon-doped photocatalysts were further confirmed by TEM and HRTEM techniques. TEM images of these photocatalysts (Fig. 4) show that the samples are composed of nanoparticles with sizes of 5–15 nm. As clearly observed in HRTEM images, the (101) planes of these carbon-doped anatase particles are well-defined, indicating the well crystallinity of the photocatalysts. The average size of the photocatalysts increases slightly with the temperature elevating and the calcination time prolonging, which is consistent with the result obtained from the XRD patterns. In particular, the polydispersity of the nanoparticles increases significantly during slow heating process. Furthermore, these nanoparticles seem to aggregate more severely, as shown in Fig. 4e and g. Better dispersion and less aggregation of nanoparticles in both F-samples may benefit their high photocatalytic activity, whereas the BET surface areas determined by N2 adsorption experiments are ca. 96, 69, 72 and 59 m2 g−1 for F-400-1, F-500-0.5, N-300-450 and N-300-500, respectively. In fact, F-500-0.5 and N-300-450 exhibit very different activities although they possess similar surface areas, while both F-samples show similar activities instead.
![]() | ||
Fig. 4 TEM and HRTEM images of F-400-1 (a and b), F-500-0.5 (c and d), N-300-450 (e and f) and N-300-500 (g and h). |
UV-vis diffuse reflectance absorption spectroscopy was utilized to characterize the optical property of the photocatalysts, and the absorption spectrum of commercial P25 was also shown for comparison. As illustrated in Fig. 5a, all the samples exhibit typical intense absorption in the UV region, mediated by the intrinsic bandgap absorption of TiO2 resulting from the electron transitions from the valence band to the conduction band (O2p → Ti3d). Compared to P25, redshift of absorption onset is obviously observed in all of the carbon-doped photocatalysts with only anatase phase, which corresponds to the narrow of bandgap energy. As a well-known indirect semiconductor, the bandgap energy (Eg) of TiO2 can be calculated from the intersection of the extrapolated linear portion in the plot of (αhν)1/2 versus the photon energy (hν) following the equation:16 (αhν)1/2 ∝ hν − Eg, where (α) is the optical absorption coefficient and linearly proportional to the absorbance (A). As shown in Fig. 5b, the bandgap of all carbon-doped photocatalysts is narrower (F-400-1, 2.62 eV; F-500-0.5, 2.55 eV; N-300-450, 2.0 eV; N-300-500, 2.4 eV; respectively) than that of P25 (3.0 eV). It has been well accepted that doping TiO2 with carbon can effectively decrease the bandgap of TiO2 although the electronic structure of carbon doped TiO2 is still a matter of disputation.24,38,39 In addition, the visible absorbance of TiO2 is significantly strengthened after carbon doping as evidenced by a long tail absorption in the visible range in Fig. 5a.
![]() | ||
Fig. 5 (a) UV-vis diffuse reflectance spectra of the carbon doped TiO2 and commercial P25; (b) plot of (Ahν)1/2 versus photon energy (eV). |
PL emission spectra have been widely used to investigate the efficiency of charge carrier trapping, migration, and transfer and to understand the recombination of free charge carriers in semiconductor particles. In this study, the PL emission spectra of all samples were examined with an Edinburgh Instruments FLS920 spectrometer equipped with a 375 nm continuous laser as the excitation light source. As shown in Fig. 6, an intense peak at about 530 nm (2.34 eV) with a broad spectral width can be observed for all samples, which has been detected in many TiO2 photocatalysts and attributed to the charge transfer from Ti3+ to the oxygen anion in a TiO68− complex,40,41 while the weak luminescence at about 640 nm (1.94 eV) might be a consequence of the Frank–Condon principle and the polarizability of the lattice ions surrounding the vacancy42 or originate from the radiative recombination of excitons trapped to surface and subsurface defects.43 Close inspection of these spectra reveals a small shift of the emission peaks toward the longer wavelength region and a decrease in emission intensity in the carbon doped TiO2 samples except for N-300-500. The decrease in emission intensity suggests that the doping of carbon into TiO2 leads to the efficient quenching of the photoluminescence with different efficiencies. Similar quenching in the luminescence intensity has also been observed for In,44 F,40 and N8 doped TiO2. Consequently, separation efficiency of the photoinduced electron and hole and thus the photocatalytic activity of the photocatalysts for water splitting may be enhanced. It should be noted that N-300-450 exhibits low activity although weak PL was also observed, which may be due to nonradiative recombination of charge carriers. This indicates different carbon states possibly exist in N-300-450 from those in F-400-1 and F-500-0.5.
It is reported that photocurrent response mainly depends on the photoelectron generation, electron–hole pair separation and the electron-transfer efficiency on the surface of a semiconductor catalyst. In a sense, responsive photocurrent intensity could reflect the overall photoelectron conversion efficiency. To further understand the enhanced photocatalytic activity of both F-samples, the transient photocurrent on–off cycles of intermittent UV-light irradiation (365 nm) with 10 mW cm−2 light density. As can be obviously seen from Fig. 7, an apparent rise in the photocurrent responses can be discerned for all the electrodes, when the light is on, and the on–off cycles of the photocurrents are reproducible. We assume that the photocurrent should be attributed to the generation and separation of photo-generated electron–hole pairs at the TiO2/electrolyte interface. Holes move to the surface of TiO2, where they are trapped or captured by reduced species in the electrolyte, while the electrons are transported to the back contact substrate via TiO2. Under dark conditions, the weak dark currents at an anodic potential of 0.5 V may result from pure electrochemical reaction on carbon-doped TiO2 electrodes.
![]() | ||
Fig. 7 Transient photocurrent responses of the photocatalysts for three 50 s light-on–off cycles in 1 M Na2SO4 aqueous solution under UV light irradiation with a bias potential of 0.5 V. |
Further observation indicates that the photocurrents of F-400-1 and F-500-0.5 are higher than those of N-300-450 and N-300-500. A higher photocurrent response means a lower electron–hole recombination and a higher photoelectron transfer efficiency, which may contribute to higher photocatalytic activity of the photocatalysts. Except for P25, the order of photocurrent intensities is nearly consistent with that of photocatalytic activity for carbon doped TiO2 (Fig. 1). Although N-300-450 exhibits very weak photoluminescence (Fig. 6), its photocurrent at 0.5 V is much lower than that of F-samples, suggesting its low efficiency in electron–hole pair separation. Lowest photocurrent was observed on P25 electrode, manifesting poorer photogenerated charge separation in P25 than that in carbon doped samples. In other words, carbon doping seems to be favourable for charge separation.
Striking difference of these photocatalysts can be easily detected by electron spin resonance (ESR) investigations at room temperature. ESR/EPR is used to characterize the unpaired electrons or paramagnetic centers such as Ti3+ and oxygen vacancy (one electron trapped) in TiO2 crystals.19,45,46 As shown in Fig. 8, F-400-1 gave rise to a very strong EPR signal at about g = 2.005, which has been attributed to a single-electron trapped oxygen vacancy.27 A weak signal was also observed for F-500-0.5, while an even weaker signal or no signal was seen for N-samples and P25. The absence of EPR signal at g < 2.0 indicates the absence of the Ti3+ spins. EPR results demonstrate that fast heating process may favour the generation of oxygen vacancies as TiO2 is doped by carbon. For the same heating rate, the lower calcination temperature used, the more oxygen vacancies generated, similar to EPR signals of Co doped TiO2 B nanotubes reported previously.47
![]() | ||
Fig. 8 EPR spectra of TiO2 photocatalysts at room temperature operating at 9.7 GHz. Inset shows the enlarged weak signals. |
As complexity of carbon doping states in TiO2 is concerned, many theoretical discussions have been reported based on density functional theory (DFT) calculations.38,39,48 Our experimental results have been supported by previous theoretical calculations. Valentin et al.49 predicted that carbon may be substituted for oxygen and oxygen vacancies would be formed under oxygen-poor conditions, while carbon may be substituted for Ti or situated in an interstice under oxygen-rich conditions when carbon concentration is relatively low. In this work, the atmosphere should be deficient in oxygen during fast combustion of oleylamine, while slow combustion of oleylamine may give rise to an atmosphere with richer oxygen. As a result, substitutional carbon atoms and oxygen vacancies are favoured when carbon is doped at a high heating rate, whereas, few oxygen vacancies will be formed at a normal heating rate. More importantly, such different doping states seem to have a great effect on photocatalytic activity toward water splitting of the carbon doped TiO2 nanomaterials. In comparison to pure TiO2 nanoparticles prepared by the same precursor reported in our previous work (hydrogen production rate of ∼50 μmol h−1 in first 10 h),28 the photocatalytic activity for all of the carbon doped TiO2 nanomaterials here is enhanced more or less. As compared with P25, a commercial TiO2 photocatalyst with mixed phases of anatase and rutile, improved photogenerated electron–hole pair separation in N-samples has been demonstrated via PL and photocurrent responses, whereas, N-samples presented lower activity than P25, which may be due to aggregation of nanoparticles during calcination (Fig. 4) and loss of surface active sites. In particular, the significant improvement in photocatalytic activity can be only realized when oxygen vacancies are formed in TiO2 during carbon doping. Such carbon doped TiO2 photocatalysts (F-samples herein) exhibited superior photocatalytic performance to P25. Moreover, high photocatalytic activity of F-samples remains after 20 h irradiation, in contrast to gradual decrease in activity for P25 and N-samples. According to previous investigations, intermediates produced during photocatalytic reaction will block active sites of the photocatalyst and result in its deactivation. When methanol is used as a sacrificial reagent in this work, enhanced electron–hole separation and more O–H species on surface evinced by FTIR spectra (Fig. 2b) in fast combustion samples may benefit complete oxidation of methanol to CO2. This possibly avoids the blockage of the active sites and high stability of these fast combustion samples are observed as a result.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra19058d |
This journal is © The Royal Society of Chemistry 2015 |