Xin Liu‡
,
Ailing Jin‡,
Yushuai Jia
*,
Junzhe Jiang,
Na Hu and
Xiangshu Chen
Jiangxi Inorganic Membrane Materials Engineering Research Centre, College of Chemistry and Chemical Engineering, Jiangxi Normal University, Nanchang 330022, P. R. China. E-mail: ysjia@jxnu.edu.cn
First published on 21st October 2015
Hybrid nanocomposites based on graphitic carbon nitride (g-C3N4) nanosheet supported ultrafine Fe2O3 nanoparticles have been successfully prepared by a facile thermal polymerization and deposition-precipitation method. Characterization results demonstrate that Fe2O3/g-C3N4 nanocomposites exhibit a well-defined morphology, in which Fe2O3 nanocrystals of 3 nm size with a narrow particle distribution are uniformly dispersed on the layers of the g-C3N4 nanosheets. Photocatalytic reaction results indicate that the photocatalytic activity of g-C3N4 is significantly enhanced after introduction of a small amount of Fe2O3, and the optimum activity of the Fe2O3/g-C3N4 nanocomposites with a weight ratio of Fe2O3 at 0.1% is up to about 3 times and 62 times higher than those of pure g-C3N4 and pure Fe2O3, respectively, for the degradation of organic dye Rhodamine B (RhB) under visible light irradiation. The high performance of the Fe2O3/g-C3N4 photocatalysts is mainly attributed to the synergistic effect at the interface of the heterojunction between Fe2O3 nanoparticles and g-C3N4 nanosheets, including improved separation efficiency of the charge carriers, a suppressed recombination process and suitable band position of the composites. These results underline the potential for the development of effective, low-cost and earth-abundant photocatalysts for the promotion of water splitting and environmental remediation under natural sunlight by construction of sustainable g-C3N4 polymeric materials.
To enhance the visible-light photocatalytic activity of g-C3N4, a variety of methods have been developed, including metal and nonmetal doping,6,14–17 nanostructure designing,18,19 sensitizing with organic dyes20,21 and coupling with other semiconductors to form composite photocatalysts.22–27 Among these strategies, semiconductor combination is one of the most effective and convenient methods for improving the photocatalytic performance with a formation of a heterojunction structure. Ge et al.28 reported a CdS quantum dots coupled g-C3N4 photocatalyst with a hydrogen evolution rate 9 times that of pure g-C3N4 owing to the synergistic effect of the heterostructure. He et al.23 prepared SnO2−x/g-C3N4 composite photocatalyst via a calcination process and suggested that the enhanced RhB photodegradation activity benefited from a direct Z-scheme mechanism, which promoted light absorption and the separation of charge carriers. Huang et al.29 developed WO3/g-C3N4 heterojunction photocatalyst by a calcination method that showed much higher efficiency for the photodegradation of organic pollutants than that of pure samples.
As a promising candidate photocatalyst, Fe2O3 is a typical non-toxic semiconductor with a narrow band gap of about 2.2 eV (much smaller than that of g-C3N4), which can absorb light up to 600 nm, corresponding to approximately 40% of the solar spectrum energy. Besides, the superior properties of good chemical stability, high electrical conductivity, earth abundance and low cost make Fe2O3 advantageous over other semiconductor materials.30 However, the photocatalytic activity of Fe2O3 is rather low due to high electron–hole recombination rate. It has been reported that fabrication of nanosized Fe2O3 particles could largely enhance the photocatalytic performance of this material.31,32 Therefore, it is then expected that if g-C3N4 was combined with Fe2O3 nanoparticles of high dispersion to form heterojunction nanocomposites, some intrinsic disadvantages of g-C3N4 such as the low separation efficiency of charge carriers and the limited visible-light absorption ability can be well overcome and it can be highly possible to achieve excellent photocatalytic activity under visible light irradiation.
Up to now, for synthesizing composites of g-C3N4 coupled with another semiconductor, most of the preparation processes employed conventional calcination or hydrothermal methods by simply combining the mixture of two pure samples or precursors together, while the direct and controllable growth of highly dispersed semiconductor nanoparticles on the layers of g-C3N4 has been rarely exploited, and the relationship between well-defined structure and photocatalytic performance needs to be further clarified. Ye et al.25 synthesized Fe2O3/g-C3N4 photocatalyst by calcination two precursors (FeCl3·6H2O and melamine) at 520 °C. Theerthagiri et al.26 mixed the two pure composites (Fe2O3 and g-C3N4) and sintered them at 300 °C to synthesize Fe2O3/g-C3N4 hybrid. In this work, we report a well-defined Fe2O3/g-C3N4 nanocomposite for the photocatalytic degradation of organic pollutant Rhodamine B (RhB) under visible light illumination. A facile strategy via a deposition-precipitation method has been developed and Fe2O3 nanoparticles with a diameter of 3 nm are homogeneously deposited on g-C3N4 nanosheets. Our results reveal that the photocatalytic activity of g-C3N4 is significantly improved after introduction of Fe2O3 nanoparticles and the optimum photocatalytic activity of Fe2O3/g-C3N4 composites is achieved at a weight ratio of 0.1% Fe2O3. XRD, SEM, high-resolution TEM, XPS, DRS and PL spectra are employed to analyze the origin of the superior photocatalytic performance and a possible mechanism is proposed.
Metal-free g-C3N4 nanosheets were prepared by a thermal polymerization method. In a typical synthesis, urea was loaded into an alumina crucible with a lid and heated in a muffle furnace at a ramping rate of 5 °C min−1 to 550 °C. This temperature was held for 4 h prior to cooling to room temperature naturally. The yellow product was collected and ground into a fine powder.
Fe2O3/g-C3N4 nanocomposites were synthesized using a facile deposition-precipitation method. As illustrated in Fig. 1, 200 mg g-C3N4 powder was first dispersed in 50 ml water by ultrasonication for 30 min. Then the mixture was transferred into a water-bath and continuously stirred at 70 °C, followed by adding of 10 ml Fe(NO3)3–ethanol solution (calculated amount of Fe(NO3)3·9H2O dissolved in ethanol) dropwise. After that, the pH value of the mixture was adjusted to 8 by careful adding of 10% ammonium hydroxide. Subsequently, the obtained solid after evaporation of the solvent was gradually heated to 400 °C in Ar atmosphere and kept for 3 h to reach complete decomposition of Fe precursors. The final composite Fe2O3/g-C3N4 photocatalysts with various Fe2O3 content (0.05 wt%, 0.1 wt%, 0.3 wt%, 0.6 wt% and 1 wt%) were obtained. For comparison, pure Fe2O3 nanopowder was also prepared following the above-mentioned procedure in the absence of g-C3N4.
![]() | ||
| Fig. 2 (a) XRD patterns of Fe2O3/g-C3N4 nanocomposites with different Fe2O3 weight percentage; (b) EDX analysis of 1 wt% Fe2O3/g-C3N4. | ||
The SEM image of pure g-C3N4 is shown in Fig. 3a. It can be seen that pure g-C3N4 prepared from urea by high temperature calcination presents flowerlike structure, which is composed of thin graphitic nanosheets with imbedded open nanopores of tens to several hundred nanometers diameter. The pure g-C3N4 of high porosity contributes to a large surface area of 38.9 m2 g−1, as determined by BET nitrogen adsorption/desorption isotherm (ESI Fig. S2†). A high surface area means exposure of more active sites for g-C3N4 in the photocatalytic reaction. Fig. 3b–d exhibit typical TEM images of 1 wt% Fe2O3/g-C3N4 nanocomposite. As can be seen, ultrafine Fe2O3 nanoparticles are uniformly dispersed on the nanosheets of g-C3N4, with close contact between each other via forming some chemical bonds. This tight interaction is favorable for the formation of a heterojunction structure in the composites and promotes effective charge transfer and separation.23 The d-spacing determined from HRTEM shown in Fig. 3e is 0.221 nm, corresponding to the (113) plane of hematite α-Fe2O3. To further confirm the microstructure of the composite, selected-area electron diffraction (SAED) was carried out on the area shown in Fig. 3d. The SAED exhibits concentric diffraction rings, indicating that the dispersed Fe2O3 nanoparticles are polycrystalline and well crystallized (Fig. 3f). The particle size distribution of Fe2O3 nanocrystals is presented in Fig. 3g. By measuring more than 300 particles from HRTEM images, it is found that Fe2O3 particles have a narrow size distribution, with most of them falling in the range of 2.5–4.0 nm with a mean size of approximately 3 nm. The above results demonstrate that Fe2O3/g-C3N4 hybrid composites with well-defined structure have been successfully prepared by the facile deposition-precipitation method used in our work, which may have promising application in the photocatalytic reaction.
![]() | ||
| Fig. 3 FESEM image of pure g-C3N4 (a); TEM images (b–d), HRTEM image (e), SAED pattern (f) and particle size distribution (g) of 1 wt% Fe2O3/g-C3N4 nanocomposite. | ||
XPS measurements were carried out to investigate the chemical states of the surface composition in pure g-C3N4 and 0.1 wt% Fe2O3/g-C3N4 nanocomposite. Fig. 4a shows the survey scan XPS spectra of both samples. The strong peaks of C 1s, N 1s and O 1s are observed in Fig. 4a, indicating the presence of C, N and O elements in both samples. While for 0.1 wt% Fe2O3/g-C3N4 sample, the Fe 2p signal is negligible, owing to the low content of Fe2O3 in the composite. High resolution spectra of C 1s, N 1s and Fe 2p are shown in Fig. 4b–d. As illustrated in Fig. 4b, three binding energy values of C 1s are observed at 284.8, 286.2 and 288.2 eV for both samples, which are assigned to sp2 C–C bonds from carbon-containing contaminations, sp3 hybridized carbon from the defects of g-C3N4 surface and sp2-bonded carbon of N–C
N in the graphitic structure, respectively, with the latter constituting the major carbon composition in the g-C3N4.33–35 In the N 1s spectrum of g-C3N4 (Fig. 4c), three peaks at 398.7, 399.8 and 401.2 eV can be clearly resolved, corresponding to sp2 hybridized nitrogen (C–N
C), sp3 hybridized nitrogen (H–N–(C)3) and amino functional groups (C–NHx), respectively.17,34,36 However, two N 1s core levels of 0.1 wt% Fe2O3/g-C3N4 are found to be slightly shifted to lower binding energies (399.5 and 401.1 eV) compared with that of g-C3N4. This is probably due to the strong interaction between Fe2O3 and g-C3N4 in the heterostructure, which gives rise to charge transfer between each other.37 The atomic ratio of C/N determined by the quantitative analysis results of XPS is 0.80 (the contribution of adventitious C 1s signal at 284.8 eV was deducted from the total C 1s peak area), which is close to the expected value of C/N in ideal g-C3N4 (0.75), reflecting the high graphitization degree of the g-C3N4 used in our experiments. The Fe 2p spectrum of 0.1 wt% Fe2O3/g-C3N4 (Fig. 4d) exhibits two characteristic peaks at 710.9 and 723.4 eV, attributed to typical Fe 2p3/2 and Fe 2p1/2 signals of Fe2O3 phase according to the literatures.38,39 Therefore, the XPS results are in good agreement with the results of XRD and TEM that ultrafine Fe2O3 nanoparticles are evenly distributed on g-C3N4 nanosheets and thus heterostructures with close interconnection are formed.
![]() | ||
| Fig. 4 XPS of synthesized g-C3N4 and 0.1 wt% Fe2O3/g-C3N4 nanocomposite: (a) survey spectra, (b) C 1s spectra, (c) N 1s spectra and (d) Fe 2p spectrum. | ||
To investigate the optical absorption properties of the as-prepared catalysts, we examined UV-vis diffuse reflectance spectra (DRS) of pure g-C3N4 and Fe2O3/g-C3N4 composites. As can be seen in Fig. 5, the g-C3N4 nanosheets show a strong absorption edge at about 450 nm, corresponding to the band gap of approximately 2.7 eV.9,29,35 Compared to g-C3N4, all of the Fe2O3/g-C3N4 composites display similar visible light absorbance property but exhibit some red-shifts in absorption wavelength, and the absorption edge gradually moves towards longer wavelength as the Fe2O3 loading increases (the color of the composites evolves from yellow to light brick-red), implying the strong interaction between Fe2O3 and g-C3N4 in the heterojunction structure, which results in modifications of optical properties of Fe2O3/g-C3N4 composites.23
![]() | ||
| Fig. 6 Changes in UV-vis absorption spectra of RhB by (a) pure g-C3N4 and (b) 0.1 wt% Fe2O3/g-C3N4 nanocomposite under visible light irradiation. | ||
The detailed photocatalytic degradation activities for Fe2O3/g-C3N4 nanocomposites with various Fe2O3 content were further investigated and shown in Fig. 7a together with that of pure Fe2O3 nanopowder and g-C3N4 for comparison. A control experiment in the absence of a photocatalyst was also carried out, and it is observed that RhB self-degradation is almost negligible. No obvious degradation is observed with Fe2O3 nanopowder alone. For pure g-C3N4, only 50% of RhB has been degraded under visible light after irradiation for 90 min. With regard to the degradation of RhB over Fe2O3/g-C3N4 composites, the case is quite different. It is found that the Fe2O3 content has a great influence on the photocatalytic activities of g-C3N4, which is significantly enhanced after loading a very low amount of Fe2O3 nanoparticles (lower than 1 wt%). The photocatalytic activity first increases as the Fe2O3 content increases from 0.05 wt% to 0.1 wt%, and then decreases at a higher Fe2O3 content. When further increasing the loading of Fe2O3 higher than 1 wt% leads to activity lower than g-C3N4. These results suggest the existence of the synergistic effect of Fe2O3/g-C3N4 heterojunction, which can effectively improve the photocatalytic activity of Fe2O3/g-C3N4 composite via facilitated separation of charge carriers on the heterojunction interface. However, it should be noted that excess Fe2O3 species are detrimental to RhB degradation, which may be related to the recombination property of Fe2O3, more Fe2O3 greatly reducing the separation efficiency of photogenerated carriers and consequently lowering the photocatalytic activity of Fe2O3/g-C3N4. As a result, the optimum photocatalytic activity is achieved for the catalyst of 0.1 wt% Fe2O3/g-C3N4 composite (about 90% of RhB has been photodegraded within 90 min).
The experimental data in Fig. 7a can be well fit to a pseudo first-order kinetics equation
| −ln(C/C0) = kt | (1) |
![]() | ||
| Fig. 8 Influences of various hole and radical scavengers in the system of photocatalytic degradation of RhB by pure g-C3N4 and 0.1 wt% Fe2O3/g-C3N4 under visible light irradiation for 90 min. | ||
The lifetime of charge carriers is of great importance to the performance of a photocatalytic reaction. The formation of heterojunction structures in the composite photocatalysts may lead to the fast separation of electrons and holes and thus result in the high photodegradation activity. As is well known that the photoluminescence (PL) emission spectra mainly stem from the recombination of free charge carriers in semiconductors,42 we carried out the room-temperature photoluminescence studies on the as-prepared samples to investigate the migration and separation efficiency of photogenerated electron–hole pairs. Fig. 9 represents the comparison of PL spectra for pure g-C3N4 and 0.1 wt% Fe2O3/g-C3N4 with an excitation wavelength of 325 nm. As for pure g-C3N4, a strong emission peak at about 445 nm and a shoulder peak around 475 nm are observed. After hybridization with Fe2O3 nanoparticles, the overall intensity of PL signals shows an obvious decrease, indicating that the recombination rate of photoinduced electron–hole pairs is greatly suppressed, which lengthens the lifetime of photogenerated carriers and subsequently favors the enhancement of photocatalytic performance.23,43 Therefore, it can be deduced from Fig. 9 that the higher photocatalytic activity for Fe2O3/g-C3N4 compared with pure g-C3N4 may arise from the higher separation efficiency of electron–hole pairs.
![]() | ||
| Fig. 9 Photoluminescence emission spectra of as-prepared pure g-C3N4 and 0.1 wt% Fe2O3/g-C3N4 samples. | ||
Based on above experimental results, a possible mechanism has been proposed for the enhanced visible-light activity of Fe2O3/g-C3N4 composite catalysts and a schematic diagram is illustrated in Fig. 10. The significant improvement of photocatalytic performance can be ascribed to the synergistic effect between Fe2O3 nanocrystals and g-C3N4 nanosheets. The band gaps of Fe2O3 and g-C3N4 are 2.2 eV (ref. 44) and 2.7 eV,9 respectively, so both pure samples can be excited by visible light and produce photoinduced electrons and holes. After Fe2O3 nanoparticles are dispersed on g-C3N4 nanosheets, the two types of semiconductors closely combine together and form heterojunction interface. Since both the bottom of the conduction band (CB) and the top of the valence band (VB) of g-C3N4 (−1.3 V and 1.4 V vs. NHE, pH = 7)45,46 are more negative than that of Fe2O3 (0.28 V and 2.48 V vs. NHE, pH = 7),44 the photogenerated electrons could transfer easily from g-C3N4 into the CB of Fe2O3, while the holes could migrate from Fe2O3 into the VB of g-C3N4 via the interface of the heterojunction. As a result, the redistribution of electrons and holes on each side of the heterojunction is achieved (as shown in Fig. 10), which greatly hinders the recombination process of electron–hole pairs and thus improves the efficiency of interfacial charge separation. In addition, the photogenerated electrons in the CB of Fe2O3 can reduce the dissolved O2 to yield abundant active ˙O2−, together with the holes in the VB of g-C3N4 employed for further effective oxidation of RhB. Therefore, owing to the high transfer and separation efficiency of charge carriers in this heterojunction system, Fe2O3/g-C3N4 composites demonstrate superior photocatalytic performance compared with that of pure g-C3N4 and Fe2O3.
![]() | ||
| Fig. 10 Proposed mechanism for the photocatalytic degradation of RhB over Fe2O3/g-C3N4 nanocomposites under visible light irradiation. | ||
Footnotes |
| † Electronic supplementary information (ESI) available: XRD pattern of 20 wt% Fe2O3/g-C3N4 nanocomposite and nitrogen adsorption/desorption isotherm of pure g-C3N4. See DOI: 10.1039/c5ra18466e |
| ‡ The authors contribute equally to this work. |
| This journal is © The Royal Society of Chemistry 2015 |