DOI:
10.1039/C5RA04100G
(Paper)
RSC Adv., 2015,
5, 34281-34291
Facile synthesis of sheet-like N–TiO2/g-C3N4 heterojunctions with highly enhanced and stable visible-light photocatalytic activities†
Received
8th March 2015
, Accepted 7th April 2015
First published on 7th April 2015
Abstract
Sheet-like N–TiO2/g-C3N4 heterojunctions with well-controlled structures as high-efficiency visible-light photocatalysts were synthesized by direct co-calcination of preformed N–TiO2 nanoparticles and g-C3N4 nanosheets. The composition, structure, morphology, and optical absorption properties of the as-prepared g-C3N4 nanosheets and N–TiO2/g-C3N4 heterojunctions were thoroughly investigated with XRD, FT-IR, SEM, TEM, XPS, and UV-vis DRS, respectively. The visible-light photocatalytic activity of the heterojunctions with different N–TiO2/g-C3N4 mass ratios was evaluated by photodegradation of rhodamine B dye and the highest enhancement was attained at a mass ratio of 1
:
3, which was 19 and 5.3 times as high as that of the individual N–TiO2 nanoparticles and g-C3N4 nanosheets, respectively. The photocatalytic activity of the composites originating from nanosheets is 11 times as high as that of the bulk g-C3N4/N–TiO2 heterojunctions. Moreover, the sheet-like N–TiO2/g-C3N4 photocatalyst exhibits excellent stability against repeated use and could still retain over 98.8% of the initial activity after seven cycles. Such a good visible-light photocatalytic performance is mainly due to the effective structure and chemical composition provided by the present method, which endow the as-prepared heterojunctions with exceptionally high specific surface area (>80 m2 g−1) as well as efficient charge transfer at the heterojunction interfaces. The charge separation mechanism during the photodegradation process was also studied and a possible photocatalytic mechanism was proposed.
1. Introduction
Since the discovery of photoinduced water-splitting on TiO2 electrodes by Fujishima and Honda,1 semiconductor-based photocatalysis has received considerable interest for enabling pollutant decomposition,2,3 and renewable fuel synthesis.4,5 TiO2-based materials have been considered as the most promising candidate in terms of catalytic activity and chemical stability.6 Unfortunately, the wide band gap of TiO2 (∼3.2 eV for anatase phase) restricts its efficiency in photoelectric conversion under solar light.5 Therefore, a lot of efforts have been devoted in recent years to the development of photocatalysts with desirable high reactivity in visible light.7–10 However, the major challenge is to develop a highly efficient, low-cost, and robust photocatalyst that can satisfy industrial requirements.
Graphitic carbon nitride (g-C3N4) has drawn much attention due to its absorption in visible light, corrosion resistance to acid and alkali and easy adjustment of the structure and performance.11 The admirable properties of g-C3N4 make it widely used in the field of decomposition of contaminants,12 water-splitting,13 and photosynthesis.14 Still, the photocatalytic activity of g-C3N4 is relatively low owing to the high recombination rate of photo-generated electron–hole pairs. Improvements of g-C3N4 photocatalytic activity have been made, for example, by chemical doping with nonmetal or metal elements,15,16 designing mesoporous structures,17 building heterojunctions,18 and so on.19,20 To date, the formation of heterojunctions by coupling g-C3N4 to another semiconductor has proven to be an effective way to the enhancement of photocatalytic activity.21 It can lead to enhanced quantum yields by decreasing the recombination rate of the photogenerated electron–hole pairs and inducing a synergistic effect including efficient charge separation and upgrading of photostability, thereby compensating the disadvantages of the individual components.22 The intense interest in two-dimensional g-C3N4 is driven by large specific area and abundant active sites.23 The excellent value of g-C3N4 nanosheets is significantly favorable for application as a photocatalyst.
It is remarkable that many heterojunctions are formed by combining ultraviolet-light photocatalysts with g-C3N4, such as TiO2,24 ZnO,25 and BiPO4.26 Only the g-C3N4 could be excited in visible light. To get most utilization of the light source, some visible-light photocatalysts were coupled with g-C3N4 to form composites.27–29 N–TiO2, which exhibits visible-light photocatalytic activity owing to the narrow band-gap induced by the hybrid of the N2p state with O2p states,8 is one of the most interesting semiconductor for coupling g-C3N4 with advantages such as low cost, high stability and activity and most importantly it possesses a band structure that matches well with the energy levels of g-C3N4. Therefore, the formation of N–TiO2/g-C3N4 heterojunctions are supposed to be quite attractive as visible-light photocatalytic materials. However, until now, methods for the preparation of N–TiO2/g-C3N4 heterojunctions are still very limited in the literature. Generally, N–TiO2/g-C3N4 heterojunctions were reported by heating the mixture of the hydrolysis product of TiCl4 and C3N4,30 or an in situ microwave-assisted approach.31 Yet, the aforementioned N–TiO2/g-C3N4 composite photocatalysts only exhibited ordinary improvement of photocatalytic activity and poor stability. The main reason is due to the lack of structural control over the individual components as well as the overall composites. Therefore, the development of powerful and less complicated preparation methods, in-depth illumination of morphology/structure related electrons/holes transfer mechanism and further improvement of the visible-light driven photocatalytic performance on both activity and stability are substantially demanded.
Herein, we report on the facile synthesis of sheet-like N–TiO2/g-C3N4 heterojuctions with highly enhanced and stable visible-light photocatalytic activities. The heterojunctions with a well-defined sheet-like morphology were prepared by direct co-calcination of pre-synthesized N–TiO2 nanoparticles and g-C3N4 nanosheets, which ensures an effective structure/morphology control over individual components as well as an exceptionally high specific surface area. The visible-light photocatalytic activity and the stability/reusability of the N–TiO2/g-C3N4 heterojunctions were evaluated by photodegradation of RhB dye. The possible photocatalytic degradation mechanism of the N–TiO2/g-C3N4 heterojunctions was also investigated.
2. Experimental section
2.1 Chemicals
Titanium tetrachloride (TiCl4, 99.9%), ammonium thiocyanate (NH4SCN, 98.5%), ammonium hydroxide (25–28% NH3·H2O in water) and rhodamine B (RhB) were all purchased from Sinopharm Chemical Co. Ltd (China) and used without further purification.
2.2 Synthesis
Synthesis of N–TiO2 nanoparticles. Excessive NH3·H2O was added to an aqueous solution of TiCl4 (3.5 M, diluted with water in an ice-water bath) and a white precipitate immediately formed, which was collected by filtration, dried under 70 °C and subsequently calcined at 400 °C for 2 h in a muffle furnace, with a heating rate of 5 °C min−1.
g-C3N4 nanosheets and pristine g-C3N4. g-C3N4 nanosheets were synthesized by thermal polycondensation of NH4SCN. Typically, 6 g of NH4SCN was put into a crucible with a half-open cover and calcined at 550 °C for 4 h in a muffle furnace, with a heating rate of 10 °C min−1. The resulting light yellow product was collected and ground into powders for further use. The prinstine g-C3N4 was synthesized in a well-closed crucible and heated at the same calcining condition. The g-C3N4 nanosheets and pristine g-C3N4 were labeled as g-C3N4 and p-g-C3N4.
N–TiO2/g-C3N4 and N–TiO2/p-g-C3N4 heterojunctions. The as-prepared g-C3N4 nanosheets were mixed with various mass ratios and ground to form a uniform mixture before co-calcination in a muffle furnace at 400 °C for 2 h, with a heating rate of 1 °C min−1. Then, the resultant composites were collected and grounded for further characterization. The N–TiO2/p-g-C3N4 heterojunctions were prepared by the same method.
2.3 Characterizations
Transmission electron microscopy (TEM), high resolution TEM (HR-TEM) and scanning electron microscopy (SEM) images were taken from a JEOL JEM-1011, a JEOL JEM-2010 and a Hitachi S-4800 electron microscope, respectively. Powder X-ray diffraction (XRD) patterns were collected on a Rigaku D/Max 2200PC diffractometer with a graphite-monochromatized Cu Kα radiation source (λ = 0.15418 nm). Fourier-transform infrared (FT-IR) spectra were recorded on a Nicolet 5DX FT-IR spectrophotometer using KBr pellets. X-ray photoelectron spectra (XPS) were recorded on a Perkin-Elmer PHI-5300 ESCA spectrometer with a 35.75 eV pass energy and an AlKα line excitation source. The core levels were calibrated according to the C1s binding energy (BE) of 284.6 eV. Thermal gravimetric analysis (TGA) was carried out with a Mettler Toledo TGA/SDTA 851E analyzer at a heating rate of 10 °C min−1 from room temperature to 800 °C under air atmosphere. Nitrogen adsorption–desorption isotherms at 77 K, Brunauer–Emmett–Teller (BET) surface areas and Barrett–Joyner–Halenda (BJH) pore size distributions were measured and calculated with a Quantachrome QuadraSorb SI analyzer and the samples were vacuum-degassed at 100 °C for 12 h before measurements. Ultraviolet-visible (UV-vis) diffuse reflectance spectra (DRS) were recorded on an Agilent Cary 100 UV-vis spectrophotometer coupled to an integrating sphere with BaSO4 as reference. Photoluminescence (PL) spectra were recorded on an Edinburgh FI/FSTCSPC 920 spectrophotometer under an excitation wavelength of 400 nm. The element analysis was performed on the Elemental Analyzer (Vario EL III, Elementar Analysensysteme) under the high temperature combustion (ca. 1100 °C furnace temperature) in an oxygen-rich environment for determining the CHNS contents of the samples. The O contents of graphitic carbon nitride were obtained by deducting the CHNS contents from the total weight.
2.4 Photocatalytic activity testing
The photocatalytic activities of the as-prepared N–TiO2/g-C3N4 composites were evaluated by photodegradation of RhB dye. An 800 W xenon lamp with a UV cut-off filter (emission range >420 nm, maximum emission at ∼470 nm) is loaded in the empty chamber of an annular quartz tube, with a circulating water jacket surrounding it to keep the temperature at 25 °C. 0.050 g of the catalyst was dispersed in a 50 mL 20 ppm RhB aqueous solution by ultrasonication for 20 min. In order to attain adsorption equilibrium, the suspension was first stirred continuously in dark for 30 min before irradiated under the xenon lamp. 6.0 mL of the suspension was taken out from the reaction flask every 5 min and centrifuged to remove the catalyst. The concentration of the remaining RhB was determined by its maximum absorbance at 554 nm (λmax) measured with a Perkin-Elmer Lambda-35 UV-vis spectrometer. Blank (without catalysts) and contrast experiments were performed following identical procedures.
3. Results and discussion
3.1 Structure and composition of g-C3N4 sheets and bulk p-g-C3N4
The morphology and microstructures of the g-C3N4 nanosheets and the p-g-C3N4 were thoroughly studied using SEM and TEM. A typically aggregated morphology with a large size and wrinkles is found in Fig. 1a and c. The surface of the sample is very rough. The essential stacked lamellar structure of g-C3N4 can be obtained from the SEM image. As shown in Fig. 1b and d, the g-C3N4 nanosheets appear as loose nanosheets with a thickness of 8–15 nm rather than condensed solid agglomerates. The sheets tend to bend and form ragged edges as a result of minimizing their surface energy.32 Fig. 1b shows the typical TEM image of an isolated piece of g-C3N4 nanosheets, which displays a two-dimensional sheet-like structure with a lateral scale of several micrometers. The FT-IR spectra (Fig. S1a†) of g-C3N4 and p-g-C3N4 are almost the same. These bands between 1700 cm−1 and 1250 cm−1 are ascribed to the C
N and C–N stretching vibration.33 The peak at 810 cm−1 is associated with breathing mode of s-triazine.34 The IR spectra suggest the conjugated triazine ring forms in the compounds after being calcined.
 |
| Fig. 1 TEM images of p-g-C3N4 (a) and g-C3N4 (b) and typical FE-SEM images of p-g-C3N4 (c) and g-C3N4 (d). | |
In order to confirm the surface composition and chemical state of the p-g-C3N4 and g-C3N4 nanosheets, the samples were analyzed by XPS technique. The XPS spectra of C1s, N1s and O1s are shown in Fig. 2a–c. For the C1s spectrum (Fig. 2a), the peak at 284.6 eV is typically attributed to graphitic carbon or external carbon contamination35 and the other peak at 287.7 eV can be assigned to the C sp2 atoms (N–C
N) in triazine rings.36 The N1s features show a broad peak extending from 395 to 403 eV as shown in Fig. 2b. The main N1s peak at 398.3 eV in g-C3N4 can be assigned to sp2 N (C
N–C) in triazine rings and the peak at 400.2 eV can be attributed to the bridging tertiary nitrogen N–(C)3.35,36 The C1s and N1s spectra of XPS are coincided with the two samples, indicating the formation of indistinctive triazine ring structure. The p-g-C3N4 shows O1s core level at 532.9 eV (Fig. 2c), ascribed to adsorbed water.31 No additional signal related to the C–O or N–O bond is observed. For the g-C3N4 nanosheets, the other two core levels at 531.6 eV and 534.0 eV are obviously detected on the doped sample. These signals could be attributed to the formation of N–C–O species and the adsorbed O2, respectively.37 It implies that the oxygen could be directly bonding with sp2-hybridized carbon atom in the g-C3N4 matrix. Fig. S1b† shows the XRD patterns of the two samples. The main peak of the p-g-C3N4 was detected at around 27.3°, while the peaks of g-C3N4 shifts to 27.5° corresponding to a decrease in the interplanar stacking distance from 0.326 to 0.324 nm. The change is probably due to the introduction of oxygen into the graphitic structure.37
|
3NH4SCN → 2NH3 + C3N4 + 3H2S
| (1) |
 |
| Fig. 2 XPS spectra (a–c) of g-C3N4 and p-g-C3N4. | |
As previous reports, the g-C3N4 can be obtained by the self-polymerization of NH4SCN according to the eqn (1). The formation processes are shown in Scheme S1a.†38 The present oxygen in the experiment can probably participate in the reaction and be doped into the g-C3N4 matrix (Scheme S1b†). According to the former report, the nanosheets of g-C3N4 could be obtained in the process of the thermal oxidation etching.32 In the experimental condition, sufficient oxygen contents can also thermally oxidate the matrix to form sheet-like morphology. As a result, the BET surface area of the g-C3N4 nanosheets increases up to 42.4 m2 g−1, which is approximately 2 times as high as that of the bulk p-g-C3N4 (22.0 m2 g−1) (Fig. S2†). The hierarchical structures with mesopores are expected to provide more active sites.
 |
| Fig. 3 UV-visible diffuse reflectance spectra (a) and fluorescence emission spectra (b) of g-C3N4 and p-g-C3N4. The inset is the plots of (αhν)2 versus photon energy (hν). | |
As above discussion, the crystalline structure of g-C3N4 nanosheets basically corresponds with bulk p-g-C3N4. The electronic band structures of the nanosheets were studied by diffuse reflectance spectra (DRS) and fluorescence emission spectra. The UV-visible DRS in Fig. 3a show a little blue shift of the intrinsic absorption edge compared with bluk-p-g-C3N4. The derived band-gaps from the plots of the (αhν)n versus the photo energy of the light adsorbed (according to the Kubelka–Munk equation, n = 2 for the direct band-gap19) become larger due to the well-known quantum confinement.32 The result is further confirmed by the blue shift of the fluorescence emission spectra in Fig. 3b. Also, there is a significant decrease in the PL intensity of the g-C3N4 nanosheets, compared to p-g-C3N4. It suggests that the recombination of electron–hole pairs might be effectively inhibited on the g-C3N4 nanosheets.
 |
| Fig. 4 Powder XRD patterns (a) and FT-IR spectra (b) of g-C3N4, N–TiO2 and N–TiO2/g-C3N4 composites with different mass ratios. | |
To evaluate the thermostability of the g-C3N4 nanosheets, the TG-DSC experiment was performed at a heating rate of 10 °C min−1 (Fig. S3†). The g-C3N4 nanosheets become unstable when the heating temperature is above 500 °C and there is an endothermal peak centered at 688 °C, which can be attributed to the decomposition of g-C3N4.39 From room temperature to 500 °C, there is no clear change for the g-C3N4 nanosheets. Thus, the g-C3N4 is relatively stable when the calcined temperature is below 500 °C. As prepared, the N content of N–TiO2, the O content of g-C3N4 nanosheets and the O content of pristine g-C3N4 agglomeration are 0.469%, 6.96% and 2.08%, respectively. After calcinations at 400 °C, the corresponding contents are changed to 0.435%, 7.45% and 2.22%. The changes of the N and O contents in the samples are very small. The high thermostability of nanosheets makes the formation of heterojunctions with N–TiO2 possible by direct co-calcination.
3.2 Morphology and microstructures of the heterojunctions of N–TiO2/g-C3N4
The structure and composition of the as-prepared samples was characterized by XRD and FT-IR. Powder XRD patterns of N–TiO2/g-C3N4 composites with different mass ratios together with the diffraction patterns of pristine N–TiO2 and g-C3N4 were shown in Fig. 4a. For the pure N–TiO2 sample, the diffraction peaks are indexed to the (101), (004), (200), (105), (211), (204), (116), (220) and (215) planes of the anatase phase (JCPDS no. 21-1272). The patterns of the N–TiO2/g-C3N4 composites show no other diffraction peaks except that of N–TiO2 and g-C3N4, indicating the coexistence of N–TiO2 and g-C3N4, and no impurities are formed during the fabrication of the composites. Furthermore, with a decreasing of the N–TiO2/g-C3N4 mass ratio from 1
:
1 to 1
:
5 the intensity of the typical diffraction peaks of the g-C3N4 gradually increases, indicating an increase of the g-C3N4 content in the as-formed composite. Fig. 4b shows the FT-IR spectra of pure N–TiO2, pure g-C3N4 and N–TiO2/g-C3N4 composites with mass ratios of 1
:
1, 1
:
2, 1
:
3 and 1
:
5. For the pure N–TiO2, the most obvious absorption at 454 cm−1 corresponds to the Ti–O–Ti vibration.40 Besides, the observed broad peak around 3500 cm−1 and the peak at 1661 cm−1 belong to the O–H stretching vibrations. For the heterojunctions, the characteristic bands between 1200 and 1640 cm−1 and the peak at 810 cm−1 are attributed to the typical stretching vibration of C–N heterocycles, and the breathing mode of triazine unites. The FT-IR spectra of the N–TiO2/g-C3N4 composites represent the overlap of the spectra of both N–TiO2 and g-C3N4, indicating composite formation does not change the structure and chemical skeleton of both components.
To investigate the surface composition and chemical state as well as the interaction between N–TiO2 and g-C3N4, the as-prepared composite was also analyzed using XPS technique. The XPS survey spectra (Fig. S4†) of the as-prepared N–TiO2/g-C3N4 (1
:
3) composite displays no other elements than Ti, C, N, and O. The high resolution XPS spectra of Ti2p, C1s, N1s and O1s for the N–TiO2/g-C3N4 (1
:
3) composite are displayed in Fig. 5. The Ti2p3/2 and Ti2p1/2 core level appear at 458.7 and 464.0 eV, respectively, which is very close to the binding energy value of N–TiO2. The inset shows the N1s spectrum for N–TiO2. The weak peak at 399.0 eV can be attributed to the N atoms located at the interstitial sites of the TiO2 lattice, such as Ti–N–O or Ti–O–N,41 and its low intensity indicates the trace amount of nitrogen in N–TiO2. The C1s and N1s spectra of the composite are similar with the g-C3N4 nanosheet. It implies the structure of g-C3N4 was retained during the formation of heterojunctions. In the O1s spectrum, the peak centered at 529.4 eV is associated with O2− in N–TiO2,42 and those at 531.7 and 533.1 eV is ascribed to N–C–O species or the absorbed H2O on the surface of photocatalysts.37,38 The higher binding energy is attributed to the heterojunctions of N–TiO2 and g-C3N4 nanosheets.
 |
| Fig. 5 High resolution XPS spectra of Ti2p (a), C1s (b), N1s (c) and O1s (d) for the N–TiO2/g-C3N4 (1 : 3) composite. The inset shows the N1s spectra of N–TiO2. | |
The morphology of the pure N–TiO2 and the N–TiO2/g-C3N4 composite with the mass ratio of 1
:
3 were studied using SEM, TEM, HR-TEM and SAED. As shown in Fig. 6a and b, the as-prepared N–TiO2 is well-defined particles in good dispersity with sizes ranging from 10 to 30 nm. The HR-TEM image (Fig. 6c) reveals the lattice fringes of an individual nanoparticle and the lattice spacing are measured to be 0.239 and 0.356 nm, corresponding to the (004) and (101) planes of anatase TiO2, respectively. The selected area electron diffraction (SAED) pattern in the inset of Fig. 6c are made up of distinct diffraction rings, indicating the high crystallinity of the nanoparticles. By mixing and co-calcination of the as-obtained N–TiO2 nanoparticles and g-C3N4 nanosheets the N–TiO2/g-C3N4 composites are formed. As can be seen from Fig. 6d, the g-C3N4 still maintains the sheet-like structure although the surface became a little rougher. The TEM image of an isolating composite sheet in Fig. 6e shows clearly darker N–TiO2 nanoparticles scattered on the g-C3N4 nanosheet, as denoted by the arrows. The interface between the two components was further revealed by HR-TEM image in Fig. 6f, which shows the N–TiO2 nanoparticle with a lattice spacing of 0.232 nm that corresponds to the (112) plane of the anatase phase were tightly contacted to the surface of the g-C3N4 nanosheet. Besides, the size of N–TiO2 nanoparticles in the composite stays almost the same as that of the pristine ones. All of the above results imply that intimate heterojunctions between N–TiO2 and g-C3N4 were indeed formed via co-calcination of the pre-synthesized nanoparticles and nanosheets. Moreover, the sheet-like structure of the N–TiO2/g-C3N4 heterojunctions is also beneficial for charge transfer compared with condensed agglomerates.
 |
| Fig. 6 SEM (a and d), TEM (b and e) and HR-TEM (c and f) images of N–TiO2 (a–c), N–TiO2/g-C3N4 (1 : 3) composite (d–f). Insets show corresponding SAED patterns. | |
3.3 Optical properties and specific surface area
The optical properties and the band gap energies of the as-prepared samples were investigated by UV-vis DRS measurement. Fig. S5a† shows the UV-vis DRS of N–TiO2 nanoparticles, g-C3N4 nanosheets and the N–TiO2/g-C3N4 (1
:
3) heterojunctions, respectively. The absorption edge of the N–TiO2 nanoparticles is above the wavelength of 420 nm. Compared with anatase TiO2, such a red shift toward the visible light range is attributed to the nitrogen doping effect. The g-C3N4 nanosheets display absorption around 460 nm. For the N–TiO2/g-C3N4 (1
:
3) heterojunctions, the absorption edge is very close to the g-C3N4 nanosheets. Accordingly, the plots of (αhν)2 versus photon energy (hν) are shown in Fig. S5b.† The corresponding band gap energies are estimated to be 3.10, 2.75 and 2.73 eV for N–TiO2, g-C3N4 and the N–TiO2/g-C3N4 (1
:
3) heterojunctions, respectively. The resulting narrow band gap of the as-prepared N–TiO2/g-C3N4 heterojunctions can lead to enhanced visible-light absorption, and consequently higher photocatalytic activities under visible-light irradiation.
The porosity and the specific surface area of the samples, which are important for photocatalytic performances, were carefully examined by nitrogen adsorption–desorption measurements (Fig. S6†). As a comparison, g-C3N4 nanosheets after being calcined at 400 °C and a mechanically mixed N–TiO2/g-C3N4 (1
:
3) composite (without calcination) were also characterized. All the samples display typical type IV nitrogen isotherms, indicating the existence of mesoporous structures. The N–TiO2 nanoparticles show the H1 type hysteresis loops with a large pore volume, a wide pore size distribution and a most probable pore size of 11 nm, which are evaluated with the Barrett–Joyner–Halenda (BJH) method. These pores are derived from the irregular aggregation of the nanoparticles. All other samples present H4 type hysteresis loops. The g-C3N4 nanosheets, both before and after being calcined at 400 °C, display negligible total pore volumes. The as-prepared N–TiO2/g-C3N4 (1
:
3) heterojunctions present a wide pore size distribution and a very small total pore volume, which is a little bit larger than the mechanically mixed N–TiO2/g-C3N4 (1
:
3) composite. The corresponding Barrett–Emmett–Teller specific surface area (SBET) for all the samples are listed in Table 1. As expected, the N–TiO2 nanoparticles show a relatively high SBET of 69.68 m2 g−1 as a result of the small size of the particles. The SBET of the g-C3N4 nanosheets increases significantly from 42.40 to 98.72 m2 g−1 after being calcined at 400 °C, which can be attributed to the thermal oxide etching of g-C3N4.27 The mechanically mixed N–TiO2/g-C3N4 (1
:
3) composite presents a SBET of 52.45 m2 g−1, whereas the N–TiO2/g-C3N4 heterojunctions display much higher SBET values, which are 74.29, 84.69, 81.83 and 82.22 m2 g−1 for the 1
:
1, 1
:
2, 1
:
3 and 1
:
5 mass ratio samples, respectively. These results indicate that the present co-calcination synthesis endows the N–TiO2/g-C3N4 heterojunctions with a large porosity as well as an exceptionally high specific surface area, which is superior to the reported N–TiO2/g-C3N4 heterojunctions, and would provide more reaction sites that are favorable for photocatalytic applications.
Table 1 Specific surface area SBET and reaction rate constant k of various samples for photodegradation of RhB
Sample |
Weight ratio |
SBET (m2 g−1) |
k (min−1) |
N–TiO2 |
— |
69.68 |
0.00881 |
g-C3N4 |
— |
42.40 |
0.0328 |
g-C3N4 (400 °C calcined) |
— |
98.72 |
0.0366 |
N–TiO2/g-C3N4 (mechanically mixed) |
1 : 3 |
52.45 |
0.0283 |
N–TiO2/g-C3N4 heterojunctions |
1 : 1 |
74.29 |
0.0513 |
1 : 2 |
84.69 |
0.128 |
1 : 3 |
81.83 |
0.173 |
1 : 5 |
82.22 |
0.102 |
3.4 Photocatalytic activity
The photocatalytic activities of the as-prepared N–TiO2/g-C3N4 heterojunctions with different mass ratios were evaluated by visible-light induced photocatalytic degradation of RhB. For comparison, the photocatalytic activities of the pure N–TiO2, the pure g-C3N4, the pure g-C3N4 after being calcined at 400 °C and the mechanically mixed N–TiO2/g-C3N4 (1
:
3) composite were also tested under the same conditions. The characteristic absorption peak of RhB at 554 nm was employed to determine the degradation degree. Fig. 7 demonstrates the evolution of C/C0 (wherein C0 is the initial concentration of RhB and C represents the concentration at reaction time t) with the visible-light irradiation time in the presence of different photocatalysts. The blank experiment indicates that the direct photolysis of RhB is almost negligible in the absence of photocatalyst, which excludes the possibility of photolysis in the present system and therefore the degradation of RhB is indeed resulted from photocatalytic reactions. The pure N–TiO2 nanoparticles show a very poor and the lowest photocatalytic activity for degradation of RhB among all these samples. The pure g-C3N4, both before and after being calcined at 400 °C, and the mechanically mixed N–TiO2/g-C3N4 composite display a very close degradation efficiency toward RhB, among which the mechanically mixed N–TiO2/g-C3N4 composite shows the lowest activity. Compared to the above control samples, the as-synthesized heterojunctions with different mass radio of 1
:
1, 1
:
2, 1
:
3 and 1
:
5 all exhibit much higher photocatalytic activities. Specifically, the photocatalytic degradation efficiency of the heterojunctions increases as the g-C3N4 portion increase from 1
:
1 to 1
:
3, yet decreases at the weight radio of 1
:
5. These results strongly suggest that the intimate junctions between N–TiO2 and g-C3N4 indeed bring a synergistic effect that is crucial to the enhancement of photocatalytic activities, and that a suitable mass ratio and effective formation of heterojunctions in the composites play an important role for the improvement of photocatalytic performance. In this case, the N–TiO2/g-C3N4 heterojunctions show the optimum photocatalytic property at the mass radio of 1
:
3, which is probably due to the formation of the most proper amount of heterojunctions in the composite. The photocatalytic degradation process of RhB in the presence of the g-C3N4/N–TiO2 (1
:
3) heterojunctions is presented in Fig. S7.† As can be seen, the adsorption equilibrium of RhB onto the photocatalyst could be reached within the first 30 min and about 40% of the total amount of RhB was adsorbed. Once the visible light was turned on, the intensity of the absorption peak of RhB at 554 nm decreased rapidly and almost disappeared after 30 min. The gradual blue shift of the maximum absorption indicates the formation of intermediate products, which were also completely degraded at the end of the degradation process.
 |
| Fig. 7 Visible-light photocatalytic degradation efficiency of RhB in the presence of N–TiO2, g-C3N4, g-C3N4 after being calcined at 400 °C, a mechanically mixed N–TiO2/g-C3N4 (1 : 3) composite and N–TiO2/g-C3N4 heterojunctions with different mass ratios. | |
To compare the photocatalytic efficiency of these samples quantitatively, the reaction kinetics of the RhB degradation were fitted by a first-order kinetic model as expressed by eqn (2):
where
k is the rate constant (min
−1). Fitting plots of the first-order kinetic model for the photocatalysis kinetic data were provided in Fig. S8.
† The
k values of all the above samples for RhB degradation were calculated and drawn in
Fig. 8. As has been discussed, the N–TiO
2/g-C
3N
4 heterojunctions with a mass ratio of 1
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
:
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
3 show the highest rate constant (0.173 min
−1), which is 19 and 5.3 times as high as that of the individual N–TiO
2 nanoparticles (0.00881 min
−1) and g-C
3N
4 nanosheets (0.0328 min
−1), respectively. Such a significant enhancement of the photocatalytic efficiency of g-C
3N
4 nanosheets is 2.0 times higher than pristine bulk g-C
3N
4 (0.0111 min
−1). At the same time, the heterojunctions of
p-g-C
3N
4 with N–TiO
2 are also prepared using the same method. The optimal photocatalytic activity of heterojunctions (1
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
:
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
3) is improved by 0.44 times compared with
p-g-C
3N
4. The rate constant of sheet-like N–TiO
2/g-C
3N
4 heterojunctions is about 11 times as high as that of the composites of
p-g-C
3N
4 and N–TiO
2 (0.0159 min
−1). Therefore, the as-prepared sheet-like N–TiO
2/g-C
3N
4 heterojunctions can be regarded as a very competitive visible-light photocatalyst.
 |
| Fig. 8 Rate constants (k) of RhB degradation over N–TiO2, g-C3N4 nanosheets and their heterojunctions (a) and N–TiO2, p-g-C3N4 and their heterojunctions (b). | |
3.5 Stability of the N–TiO2/g-C3N4 photocatalysts
The stability and recycle/reusability of the photocatalysts were evaluated over multiple cycles of RhB degradation and photocatalyst regeneration with the N–TiO2/g-C3N4 (1
:
3) heterojunctions as the representative sample. In every cycle experiment, the used photocatalyst was regenerated by centrifuging and drying at 60 °C for 12 h. As shown in Fig. 9, the as-prepared N–TiO2/g-C3N4 heterojunctions have demonstrated excellent photocatalytic stability, maintaining nearly the same high level of reactivity after seven cycles. The degraded proportion of RhB over the photocatalyst was still up to 98.8% at the end of the 7th cycle. In addition, the composition and microstructure of the recycled composite were characterized by XRD, FT-IR and SEM. The XRD pattern and FT-IR spectra of the photocatalyst after seven cycles are almost the same as those of the freshly prepared sample (Fig. S9†). Moreover, the regenerated photocatalyst still appears the light yellow color without any visible change. The SEM image (Fig. S10†) shows that the sheet-like morphology of the photocatalyst remained. These results verify that the stability and recycle/reusability of the as-prepared N–TiO2/g-C3N4 heterojunctions are good enough for applications in practical photocatalytic degradations.
 |
| Fig. 9 Cycling runs (a) and degraded amount (b) for the photocatalytic degradation of RhB over N–TiO2/g-C3N4 (1 : 3) heterojunctions under visible-light irradiation. | |
3.6 Possible photocatalytic mechanisms of N–TiO2/g-C3N4 heterojunctions
To investigate the fate of electron–hole pairs within the as-prepared heterojunctions, photoluminescence (PL) emission has been carried out. For comparison, the PL of the pure g-C3N4 nanosheets was also measured. As shown in Fig. 10, the main emission peak for pure g-C3N4 is centered at 450 nm, which is approximately equal to the band gap energy of pure g-C3N4, whereas the as-prepared N–TiO2/g-C3N4 heterojunctions with different mass ratios all show significant quenching of the PL peaks. As the intensity of the PL peaks is closely related to the recombination of the electron–hole pairs, this result further indicates an efficient charge transfer within the N–TiO2/g-C3N4 heterojunctions and thus an improvement in the separation efficiency of the light-stimulated carriers. That is to say, the photogenerated electron–hole pairs can efficiently transfer at the interface of N–TiO2/g-C3N4 heterojunctions, resulting in higher photocatalytic activity under visible light irradiation.
In order to elucidate the photocatalytic oxidation process of RhB in the presence of N–TiO2/g-C3N4, radical and hole trapping experiments are designed by introducing different scavengers to the degradation process. Fig. 11 displays the effect of different scavengers on the RhB degradation in the presence of the N–TiO2/g-C3N4 (1
:
3) heterojunctions under visible light irradiation for 20 min. The less percentage of RhB is converted, the more important role the active species play in the reaction. Superoxide radical (˙O2−), hole (h+), electron (e−) and hydroxyl radical (˙OH) are explored in the reaction. The scavengers used in the experiments are benzoquinone (BQ, 2 mM, the ˙O2− scavenger),43 disodium ethylenediaminetetraacetate (EDTA, 2 mM, the h+ scavenger),44 AgNO3 (2 mM, the e− scavenger)45 and isopropanol (IPA, 2 mM, the ˙OH scavenger).25 As shown in Fig. 11, without any scavenger addition the conversion percentage is 99.1%. The addition of IPA and AgNO3 both lead to non-significant change, suggesting that ˙OH and e− are not the main reactive species involved in the RhB photocatalytic oxidation process. However, upon the addition of BQ and EDTA, the conversion percentage of RhB is greatly suppressed to 59% and 45%, respectively. This result indicates that ˙O2− and h+ are the major species involved in the photocatalytic oxidation process.
 |
| Fig. 10 PL spectra of g-C3N4 and N–TiO2/g-C3N4 heterojunctions with different mass ratios. | |
 |
| Fig. 11 Conversion percentage of RhB in the presence of N–TiO2/g-C3N4 photocatalyst with different scavengers. | |
Based on the above results, a possible photocatalytic mechanism of the as-prepared N–TiO2/g-C3N4 heterojunctions under visible light irradiation was proposed and illustrated in Fig. 12. In this study, the band gap energy of g-C3N4 and N–TiO2 are determined to be 2.75 and 3.10 eV, respectively, suggesting the two semiconductors have well-matched energy band structure for the formation of heterojunctions. Both the N–TiO2 and g-C3N4 have the ability to absorb visible light, resulting in the excitation of electrons from valence band (VB) to the conductor band (CB). As the CB potential of g-C3N4 is more negative than that of N–TiO2,31 the photogenerated electrons can directly inject into the CB of N–TiO2 via the interfaces, while the holes on the N–TiO2 surface can migrate to the VB of g-C3N4 in the same manner. Such transference can efficiently suppress the recombination of electrons and holes. The untransferred electrons on the CB of g-C3N4 and electrons on the CB of N–TiO2 can capture the O2 to generated ˙O2− radicals easily, which is one of the main active species for oxidizing RhB. On the other hand, the g-C3N4 with congregated holes can be charged with the segments of RhB and fall back to ground state according to the above results, thus also playing an important part in the degradation of RhB. These two paths present the most likely photocatalytic mechanism of the degradation of RhB over the as-prepared N–TiO2/g-C3N4 photocatalysts under visible-light irradiation. Compared with individual components, the formation of N–TiO2/g-C3N4 heterojunctions induces more electrons to congregate on the surface of N–TiO2 and more holes on the surface of g-C3N4, thus leading to the overall enhancement of photocatalytic activity.
 |
| Fig. 12 Schematic diagrams for the possible photocatalytic mechanism of the N–TiO2/g-C3N4 heterojunction under visible light irradiation. Band energy diagrams for N–TiO2 and g-C3N4. | |
4. Conclusions
Sheet-like N–TiO2/g-C3N4 heterojunctions with different mass ratios were synthesized by direct co-calcination of separately formed N–TiO2 nanoparticles and g-C3N4 nanosheets. Such a processing method not only ensures the successful formation of N–TiO2/g-C3N4 heterojunctions but also allows the structure/morphology control over individual components as well as the overall composite. Compared to individual components, the as-prepared N–TiO2/g-C3N4 heterojunctions have exhibited highly enhanced photocatalytic activity toward the degradation of RhB under visible-light irradiation, which is mainly attributed to the synergetic effects between N–TiO2 nanoparticles and g-C3N4 nanosheets. In the degradation of RhB, ˙O2− and h+ have played a major role whereas the contribution from ˙OH and e− can be negligible. The highly enhanced visible-light photocatalytic activity and its excellent stability surpass existing N–TiO2/g-C3N4 composite systems obtained via other synthetic routes mainly due to the effective structural control over the individual components as well as the overall composites provided by the present method, which endow the as-prepared heterojunctions with exceptionally high specific surface area (>80 m2 g−1) and efficient charge transfer at the heterojunction interfaces. This work shows that structural control over heterojunctions via designing synthetic strategy can bring about improved photocatalytic performance and provides a facile and powerful method for fabricating highly efficient and stable heterostructured photocatalysts for environmental remediation.
Acknowledgements
This work is supported by the Taishan Scholars Climbing Program of Shandong Province, Natural Science Foundation of China (NSFC 21271118, 21303095), Independent Innovation Foundation of Shandong University (2013TB001). The author W. Li acknowledges Dr Y. Lai for helpful discussions.
References
- A. Fujishima and K. Honda, Nature, 1972, 238, 37 CrossRef CAS.
- J. M. Herrmann, Catal. Today, 1999, 53, 115 CrossRef CAS.
- H. Tada, T. Kiyonaga and S. Naya, Chem. Soc. Rev., 2009, 38, 1849 RSC.
- A. Kudo and Y. Miseki, Chem. Soc. Rev., 2009, 38, 253 RSC.
- X. Chen, S. Shen, L. Guo and S. Mao, Chem. Rev., 2010, 110, 65030 Search PubMed.
- G. Mor, O. K. Varghese, M. Paulose, K. Shankar and C. A. Grimes, Sol. Energy Mater. Sol. Cells, 2006, 90, 2011 CrossRef CAS PubMed.
- M. Anpo and M. Takeuchi, J. Catal., 2003, 216, 505 CrossRef CAS.
- R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki and Y. Taga, Science, 2001, 293, 269 CrossRef CAS PubMed.
- J. Kim, C. W. Lee and W. Choi, Environ. Sci. Technol., 2010, 44, 6849 CrossRef CAS PubMed.
- Z. Yi, J. Ye, N. Kikugawa, T. Kako, S. Ouyang, H. Stuart-Williams, H. Yang, J. Cao, W. Luo, Z. Li, Y. Liu and R. L. Withers, Nat. Mater., 2010, 9, 559 CrossRef CAS PubMed.
- X. Wang, K. Maeda, A. Thomas, K. Takanabe, G. Xin, J. M. Carlsson, K. Domen and M. Antonietti, Nat. Mater., 2009, 8, 76 CrossRef CAS PubMed.
- Y. Cui, J. Huang, X. Fu and X. Wang, Catal. Sci. Technol., 2012, 2, 1396 CAS.
- D. J. Martin, K. Qiu, S. A. Shevlin, A. D. Handoko, X. Chen, Z. Guo and J. Tang, Angew. Chem., Int. Ed., 2014, 53, 9240 CrossRef CAS PubMed.
- Q. Su, J. Sun, J. Wang, Z. Yang, W. Cheng and S. Zhang, Catal. Sci. Technol., 2014, 4, 1556 CAS.
- S. Yan, Z. Li and Z. Zou, Langmuir, 2010, 26, 3894 CrossRef CAS PubMed.
- S. Samanta, S. Martha and K. Parida, ChemCatChem, 2014, 6, 1453 CAS.
- X. Wang, K. Maeda, X. Chen, K. Takanabe, K. Domen, Y. Hou, X. Fu and M. Antonietti, J. Am. Chem. Soc., 2009, 131, 1680 CrossRef CAS PubMed.
- J. Fu, Y. Tian, B. Chang, F. Xi and X. Dong, J. Mater. Chem., 2012, 22, 21159 RSC.
- J. Liu, Y. Liu, N. Liu, Y. Han, X. Zhang, H. Huang, Y. Lifshitz, S. Lee, J. Zhong and Z. Kang, Science, 2015, 347, 970 CrossRef CAS PubMed.
- M. Tahir, C. Cao, N. Mahmood, F. K. Butt, A. Mahmood, F. Idrees, S. Hussain, M. Tanveer, Z. Ali and I. Aslam, ACS Appl. Mater. Interfaces, 2014, 6, 1258 CAS.
- Z. Zhao, Y. Sun and F. Dong, Nanoscale, 2015, 7, 15 RSC.
- S. Cao, J. Low, J. Yu and M. Jaroniec, Adv. Mater., 2015, 27, 2150 CrossRef CAS PubMed.
- H. Zhao, H. Yu, X. Quan, S. Chen, H. Zhao and H. Wang, RSC Adv., 2014, 4, 624 RSC.
- K. Sridharan, E. Jang and T. Park, Appl. Catal., B, 2013, 142, 718 CrossRef PubMed.
- W. Liu, M. Wang, C. Xu and S. Chen, Chem. Eng. J., 2012, 209, 386 CrossRef CAS PubMed.
- Z. Li, S. Yang, J. Zhou, D. Li, X. Zhou, C. Ge and Y. Fang, Chem. Eng. J., 2014, 241, 344 CrossRef CAS PubMed.
- K. Katsumata, R. Motoyoshi, N. Matsushita and K. Okada, J. Hazard. Mater., 2013, 260, 475 CrossRef CAS PubMed.
- S. Yan, S. Lv, Z. Li and Z. Zou, Dalton Trans., 2010, 39, 1488 RSC.
- S. Kumar, T. Surendar, A. Baruah and V. Shanker, J. Mater. Chem. A, 2013, 1, 5333 CAS.
- N. Yang, G. Li, W. Wang, X. Yang and W. Zhang, J. Phys. Chem. Solids, 2011, 72, 1319 CrossRef CAS PubMed.
- X. Wang, W. Yang, F. Li, Y. Xue, R. Liu and Y. Hao, Ind. Eng. Chem. Res., 2013, 52, 17140 CrossRef CAS.
- P. Niu, L. Zhang, G. Liu and H. Cheng, Adv. Funct. Mater., 2012, 22, 4763 CrossRef CAS PubMed.
- F. Dong, L. Wu, Y. Sun, M. Fu, Z. Wu and S. Lee, J. Mater. Chem., 2011, 21, 15171 RSC.
- M. Shalom, S. Inal, C. Fettkenhauer, D. Neher and M. Antonietti, J. Am. Chem. Soc., 2013, 135, 7118 CrossRef CAS PubMed.
- H. Li, J. Liu, W. Hou, N. Du, R. Zhang and X. Tao, Appl. Catal., B, 2014, 160–161, 89 CrossRef CAS PubMed.
- S. Yang, Y. Gong, J. Zhang, L. Zhan, L. Ma, Z. Fang, R. Vajtai, X. Wang and P. M. Ajayan, Adv. Mater., 2013, 25, 2452 CrossRef CAS PubMed.
- J. Li, B. Shen, Z. Hong, B. Lin, B. Gao and Y. Chen, Chem. Commun., 2012, 48, 12017 RSC.
- Y. Cui, J. Zhang, G. Zhang, J. Huang, P. Liu, M. Antonietti and X. Wang, J. Mater. Chem., 2011, 21, 13032 RSC.
- S. Yan, Z. Li and Z. Zou, Langmuir, 2009, 25, 10397 CrossRef CAS PubMed.
- T. C. Jagadale, S. P. Takale, R. S. Sonawane, H. M. Joshi, S. I. Patil, B. B. Kale and S. B. Ogale, J. Phys. Chem. C, 2008, 112, 14595 CAS.
- D. Mitoraj and H. Kisch, Angew. Chem., Int. Ed., 2008, 47, 9975 CrossRef CAS PubMed.
- J. Wang, D. N. Tafen, J. P. Lewis, Z. Hong, A. Manivannan, M. Zhi, M. Li and N. Wu, J. Am. Chem. Soc., 2009, 131, 12290 CrossRef CAS PubMed.
- L. Mohapatra, K. Parida and M. Satpathy, J. Phys. Chem. C, 2012, 116, 13063 CAS.
- Y. Tian, B. Chang, J. Lu, J. Fu, F. Xi and X. Dong, ACS Appl. Mater. Interfaces, 2013, 5, 7079 CAS.
- Y. Huo, J. Zhang, M. Miao and Y. Jin, Appl. Catal., B, 2012, 111–112, 334 CrossRef CAS PubMed.
Footnote |
† Electronic supplementary information (ESI) available: X-ray photoelectron spectra, UV-vis diffuse reflectance spectra, FT-IR spectra, TG-DSC thermograms, nitrogen adsorption–desorption isotherms of different photocatalysts, evolution of UV-vis spectra of RhB in the presence of photocatalysts, XRD, FT-IR and SEM of the recycled photocatalyst. See DOI: 10.1039/c5ra04100g |
|
This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.