Shang-Yi Choua,
Wen-Hsin Chungb,
Li-Wen Chena,
Yong-Ming Daia,
Wan-Yu Linb,
Jia-Hao Lina and
Chiing-Chang Chen*a
aDepartment of Science Education and Application, National Taichung University of Education, Taichung 403, Taiwan. E-mail: ccchen@mail.ntcu.edu.tw; Fax: +886-4-2218-3560; Tel: +886-4-2218-3406
bDepartment of Plant Pathology, National Chung Hsing University, Taichung 402, Taiwan
First published on 17th August 2016
A series of bismuth oxyiodide (BiOxIy)-grafted graphene oxide (GO) sheets with different GO contents were synthesized through a simple hydrothermal method. This is the first report where four composites of BiOI/GO, Bi4O5I2/GO, Bi7O9I3/GO, and Bi5O7I/GO have been characterized using X-ray diffraction, transmission electron microscopy, scanning electron microscopy energy-dispersive spectroscopy, Fourier transform infrared spectroscopy, X-ray photoelectron spectroscopy, and diffuse reflectance spectroscopy. The assembled BiOxIy/GO composites exhibited excellent photocatalytic activities in the degradation of crystal violet (CV) under visible light irradiation. The order of rate constants was as follows: Bi7O9I3/GO > Bi4O5I2/GO > Bi4O5I2 > Bi7O9I3 > Bi5O7I/GO > BiOI/GO > BiOI > Bi5O7I > GO. The photocatalytic activity of the Bi7O9I3/GO (or Bi4O5I2/GO) composite reached a maximum rate constant of 0.351 (or 0.322) h−1, which was 1.8 (or 1.7) times higher than that of Bi7O9I3 (or Bi4O5I2), 6–7 times higher than that of BiOI/GO, and 119–130 times higher than that of BiOI. The quenching effects of different scavengers and electron paramagnetic resonance demonstrated that the superoxide radical (O2˙−) played a major role and holes (h+) and hydroxyl radicals (˙OH) played a minor role as active species in the degradation of crystal violet (CV) and salicylic acid (SA). Possible photodegradation mechanisms are proposed and discussed in this research.
In recent years, semiconductor photocatalysis driven by visible light has sparked significant research interest because it provides a promising mechanism for solving energy supply and environmental pollution problems. An environmentally cheap and powerful photocatalyst is an important constituent for practical applications of photocatalysis.5 The photocatalytic degradation of CV dyes has been researched using several systems to determine active species, including bismuth oxyhalide,6–8 TiO2,9 PbBiO2Br/BiOBr,10 Bi2WO6,11 BaTiO3,12 and ZnO.13
Recently, the study of visible light-driven photocatalysts has attracted considerable attention as an alternative for the elimination of toxic materials from wastewater. An effective and simple tactic for improving the photocatalytic activity of a photocatalyst is the incorporation of a heterostructure because hetero-composites have great potential for tuning the desired electronic properties of photocatalysts and efficiently separating photo-formed electron–hole pairs.14,15
In recent years, as a new family of advantageous photocatalysts, bismuth oxyhalides16–18 have demonstrated unusual photocatalytic activities because their unique layered structure features an internal static electric field vertical in each layer that may occasion more effective separation of photo-formed charge carriers. Bismuth oxyiodides have received increased interest because of their suitable energy gaps, stability, and relatively superior photocatalytic activities.19–22
Because the valence band for bismuth oxyiodides contains mostly O2p and I5p orbitals, whereas the conduction band is based on the Bi6p orbital,23 iodine-poor bismuth oxyiodides could be illustrated to have band-gap energies higher than those of BiOI but lower than those of Bi2O3;24,25 hence, these materials might be used as visible light-responsive photocatalysts. In particular, the structure and composition of bismuth oxyiodides strongly influence their electronic, optical, oxidizing abilities, and other physicochemical properties, giving an opportunity to acquire novel photocatalysts for effective degradation of environmental and toxic pollutants. However, the synthesis methods, characterization, and evaluated properties of a series of bismuth oxyiodides have remained rare until recently.
In the past several years, graphene (GR) has drawn attention because of its remarkable properties.26,27 According to a previous report, the theoretical specific surface area of GR is 2630 m2 g−1.26,27 Therefore, it can be used as an ideal support material with improved interfacial contact and enhanced adsorption activity. GR has a powerful but flexible structure with a high carrier mobility. Thus, a GR-based hybrid photocatalyst will show excellent photocatalytic efficiency. Furthermore, it can be easily produced from graphite, which is cheap and naturally abundant. Currently, many studies have reported that the integration of GR and a semiconductor photocatalyst, such as TiO2,28 Ag/Ag2CO3,29 Bi3.64Mo0.36O6.55,30 and ZnO,31 could form hybrid materials with superior photocatalytic activity. For example, Trapalis et al.32 revealed that TiO2–graphene composites prepared through a solvothermal process displayed improved photocatalytic performance in removing NOx compared with pure commercial TiO2.
GR or reduced graphene oxide (rGO) are suitable candidates because of the high electron mobility (>15000 cm2 V−1 s−1) and the flexible sheet nature that is beneficial for supporting photocatalysts. A previous study showed that the incorporation of GO with a metal oxide could enhance the photocatalytic activity.33 Apparently, photocatalysis enhancement through GO is because a conjugate structure provides a pathway for the transport of charge carriers.
One specific branch of GR research dealt with GO. This could be considered as a precursor of semiconductor/GO synthesis by either chemical or thermal processes. Song et al.29 synthesized an Ag/Ag2CO3/rGO composite that exhibited enhanced photocatalytic performance for the photocatalytic oxidation of organic pollutants. Zhang et al.28 prepared TiO2–graphene composites for photochemical processes to ultimately degrade organic compounds into CO2 and H2O at ambient conditions. Yang et al.34 reported that Ag3PO4/GR was applied to the suspended photocatalytic degradation of organic dyes. Bhirud et al.31 obtained a ZnO–graphene composite through a hydrothermal treatment using Zn powder as the reducing agent and precursor. Recently, BiOI/graphene,35 BiOBr/graphene,36 and BiOBrxI1−x/graphene37 composites have been synthesized to improve the photocatalytic activity of materials. Functionalizing GO nanosheets with a series of BiOxIy should not only combine the advantages of both BiOxIy and GO nanosheets but also result in new properties. However, no work related to a series of GO-based BiOxIy photocatalysts has been reported.
As shown in Table 1, BiOxIy/GO (or rGO, GR) composites have obtained remarkable interest in recent years because of their suitable band gaps, stability, and relatively superior photocatalytic activities. It is found that the BiOxIy/GO (or rGO, GR) composite shows higher photocatalytic activities than BiOxIy and GO (or rGO, GR) for the photocatalytic degradation of rhodamine B (or methyl orange, phenol).38–42
Composite photocatalyst | Mass fraction of GO | Parameters of photocatalytic experiments | Photocatalytic activity | Reference photocatalyst/photocatalytic activity | Enhancement factor | Reference |
---|---|---|---|---|---|---|
BiOI/GO | 1.2% | Phenol | 55% decomposition in 2.5 h | GO: 0% | — | 66 |
BiOI: 40% | 1.4 | |||||
Bi7O9I3/rGO | 10% | Rhodamine B | 95.8% in 100 min | rGO: 0% | — | 67 |
78.5% in 150 min | Bi7O9I3: 74.7% | 2.13 | ||||
Phenol | rGO: 0% | — | ||||
Bi7O9I3: 49.2% | 2.29 | |||||
BiOI/rGO | 0.5% | Methyl orange | 99% in 3 h. | rGO | — | 68 |
BiOI: 46% | 2.15 | |||||
BiOI/GR | 5% | Rhodamine B | 80% in 4 h. | GR: 0% | — | 69 |
BiOI: 11% | 7.3 | |||||
BiOI/GR | 2% | Methyl orange | 88.0% in 4 h. | GR: 0% | — | 70 |
BiOI: 27.6% | 6.0 |
According to our literature search, a series of BiOxIy/GO-assisted photocatalytic degradations of CV dyes under visible light irradiation has never been reported. This study synthesized BiOI/GO, Bi4O5I2/GO, Bi7O9I3/GO, and Bi5O7I/GO heterojunctions and compared their photocatalytic activities in degrading CV in aqueous solutions under visible light irradiation.
X-ray diffraction (XRD) patterns were recorded on a MAC Science MXP18 instrument equipped with Cu-Kα radiation, operating at 40 kV and 80 mA. Field-emission transmission electron microscopy (FE-TEM) images, selected area electron diffraction patterns, high resolution transmission electron microscopy (HRTEM) images, and energy-dispersive X-ray spectroscopy (EDS) spectra were obtained using a JEOL-2010 instrument with an accelerating voltage of 200 kV. Al-Kα radiation was generated at 15 kV. Field emission scanning electron microscopy electron-dispersive X-ray spectroscopy (FE-SEM-EDS) measurements were conducted using a JEOL JSM-7401F instrument at an acceleration voltage of 15 kV. High resolution X-ray photoelectron spectroscopy measurements were conducted using an ULVAC-PHI instrument. Photoluminescence (PL) measurements were conducted on a Hitachi F-7000 instrument. Ultraviolet photoelectron spectroscopic measurements were performed using an ULVAC-PHI XPS, PHI Quantera SXM. Brunauer–Emmett–Teller (BET) specific surface areas of the samples (SBET) were measured with an automated system (Micrometrics Gemini) by using nitrogen gas as the adsorbate at liquid nitrogen temperature. The mineralization of BCEXM was monitored by measuring the total organic carbon (TOC) content with a Dohrmann Phoenix 8000 Carbon Analyzer using a UV/persulfate oxidation method by directly injecting the aqueous solution into the instrument.
Five mmol Bi (NO3)3·5H2O was first mixed in a 50 mL flask, followed by the addition of 5 mL 4 M ethylene glycerol and GO powder. With continuous stirring, 2 M NaOH was added dropwise to adjust the pH value; when a precipitate was formed, 2 mL KI was also added dropwise. The solution was then stirred vigorously for 30 min and transferred into a 30 mL Teflon-lined autoclave, which was heated to 100–250 °C for 12 h and then naturally cooled to room temperature. The resulting solid precipitate was collected by filtration, washed with deionized water and methanol to remove any possible ionic species in the solid precipitate, and then dried at 60 °C overnight. Depending on the molar ratio of Bi(NO3)3·5H2O to KI, pH value, temperature, and time, different BiOxIy/GO samples could be synthesized.
A series of quenchers were introduced to scavenge the relevant active species to evaluate the effect of the active species during the photocatalytic reaction. Superoxide radicals, hydroxyl radicals, holes, and singlet oxygen (1O2) were studied by adding 1.0 mM benzoquinone (BQ, a quencher of superoxide radicals),44 1.0 mM isopropanol (IPA, a quencher of hydroxyl radicals),45 1.0 mM ammonium oxalate (AO, a quencher of holes),46 and 1.0 mM sodium azide (SA, a quencher of singlet oxygen).47 The method was similar to the previously reported photocatalytic activity test.44–47
Fig. 1 shows the XRD patterns of the as-prepared samples; the patterns clearly show the existence of different BiOxIy phase composites with GO. All the as-prepared samples contained the BiOI phase (JCPDS 73-2062), Bi4O5I2 phase,48 Bi7O9I3 phase,49 Bi5O7I phase (JCPDS 40-0548), and GO.50 At pH = 1, the XRD patterns (Fig. 1(a)) were identical to those reported for the BiOI/GO binary phases; at pH = 4, the XRD patterns (Fig. 1(b)) were identical to those reported for the Bi4O5I2/GO binary phases; at pH = 7, the XRD patterns (Fig. 1(c)) were identical to those reported for the Bi4O5I2/GO binary phases (0.15 g) and Bi7O9I3/GO binary phases (0.005–0.10 g); at pH = 10, the XRD patterns (Fig. 1(d)) were identical to those reported for the Bi7O9I3/GO binary phases; and at pH = 13, the XRD patterns (Fig. 1(e)) were identical to those reported for the Bi5O7I/GO binary phases. Table 2 summarizes the results of the XRD measurements.
Fig. 2 and S1–S5 of the ESI† illustrate that the as-prepared samples were composed of differently sized layers, consistent with the TEM observations. The graphene oxide nanosheets and bismuth oxyiodide are clearly observed. The graphene oxide sheets are not very flat but display intrinsic microscopic wrinkles, and bismuth oxyiodide is dispersed on the graphene oxide nanosheets. In addition, the EDS spectrum shows that the sample contained the elements of Bi, I, O, and C. In Fig. 2(a) and (b), the HRTEM image shows that two sets of different lattice images were found with a d-spacing of 0.282 nm, corresponding to the (110) plane of BiOI, which is in satisfactory agreement with the XRD results (Fig. 1(a)). In Fig. 2(c) and (d), the HRTEM image shows that two sets of different lattice images were found with a d-spacing of 0.286 nm, corresponding to the (402) plane of Bi4O5I2, which is in satisfactory agreement with the XRD results (Fig. 1(b)). In Fig. 2(e) and (f), the HRTEM image shows that two sets of different lattice images were found with a d-spacing of 0.317 nm, corresponding to the (110) plane of Bi7O9I3, which is in satisfactory agreement with the XRD results (Fig. 1(c)). In Fig. 2(g) and (h), the HRTEM image shows that three sets of different lattice images were found with a d-spacing of 0.313 nm, corresponding to the (110) plane of Bi7O9I3, which is in satisfactory agreement with the XRD results (Fig. 1(d)). In Fig. 2(i) and (j), the HRTEM image shows that two sets of different lattice images were found with a d-spacing of 0.318 nm, corresponding to the (312) plane of Bi5O7I, which is in satisfactory agreement with the XRD results (Fig. 1(e)). The results suggest that the series of BiOxIy/GO phases were produced in the composites, which are favorable for the separation of photoinduced carriers, yielding high photocatalytic activities.
![]() | ||
Fig. 2 FE-TEM of the as-prepared (a) and (b) BiOI/GO, (c) and (d) Bi4O5I2/GO, (e) and (f) Bi7O9I3/GO, (g) and (h) Bi7O9I3/GO, and (i) and (j) Bi5O7I/GO. |
The results illustrate that at different pH values, a series of changes occurred in the products. The proposed processes for the formation of the BiOxIy/GO composites are described in eqn (1)–(7). The results demonstrate that a series of changes in the compounds occurred at different hydrothermal conditions, expressed as BiOI → Bi4O5I2 → Bi7O9I3 → Bi5O7I → α-Bi2O3. By controlling the pH of the hydrothermal reaction, different compositions of bismuth oxyiodides were obtained.
Bi3+ + 3OH− → Bi(OH)3(s) | (1) |
2Bi(OH)3 + I− + g-C3N4 → BiOI/g-C3N4 + H2O + OH− | (2) |
4BiOI + 2OH− + g-C3N4 → Bi4O5I2/g-C3N4 + 2I− + H2O | (3) |
7Bi4O5I2 + 2OH− + g-C3N4 → 4Bi7O9I3/g-C3N4 + 2I− + H2O | (4) |
3Bi7O9I3 + 2OH− + g-C3N4 → 7Bi3O4I/g-C3N4 + 2I− + H2O | (5) |
5Bi3O4I + 2OH− + g-C3N4 → 3Bi5O7I/g-C3N4 + 2I− + H2O | (6) |
2Bi5O7I + 2OH− + g-C3N4 → 5Bi2O3/g-C3N4 + 2I− + H2O | (7) |
Fig. 3 shows the Fourier transform infrared (FT-IR) spectra of the Bi7O9I3/GO composite produced at different weight percentages, where a strong absorption was located mainly in the range of 400–700 cm−1, as a result of the stretching vibrations of Bi–O, Bi–O–I, and Bi–O–Bi in bismuth oxyiodides.7 The FT-IR spectrum of GO shows a strong absorption band at 3429 cm−1 because of the O–H stretching vibration. The spectrum also exhibits bands at approximately 1726 cm−1 because of the CO stretching of COOH groups situated at the edges of the GO sheets, and the O–H bending vibration, epoxide groups, and skeletal ring vibrations are observed at approximately 1631 cm−1.51 The absorption at 1396 and 1000 cm−1 may be attributed to tertiary C–OH and the stretching of C–O–C groups. After the composition process, the intensities of the absorption bands because of the O–H stretching vibration (3429 cm−1) and C
O stretching vibration (1726 cm−1) decreased substantially, and the band at 1631 cm−1 was absent.52 Instead, a new absorption band appeared at 1570 cm−1, which was attributed to the skeletal vibration of the graphene sheets.53 This result agrees with that of the XRD and TEM experiments.
![]() | ||
Fig. 4 XPS spectra of Bi7O9I3/GO (pH = 7) for (a) Bi 4f, (b) I 3d, (c) O 1s, (d) GO O 1s, (e) C 1s, (f) GO C 1s, and (g) 0.05 g GO C 1s. |
From Fig. 4(b), the binding energies of 630.7 eV and 619.1 eV were attributed to I 3d3/2 and 3d5/2, respectively, which could point to I in the monovalent oxidation state. Fig. 4(c) shows the high-resolution XPS spectra for the O 1s region of the Bi7O9I3/GO composites, which could be resolved into two peaks; the main peak at 529.7 eV was attributed to the Bi–O bonds in the (Bi2O2)2+ slabs of the BiOX layered structure, whereas the peak at 531.3 eV was assigned to the hydroxyl groups on the surface.55 The asymmetric O 1s peak shown in Fig. 4(d) can be split by using the XPS peak-fitting program for pure GO. The peak at 531.2 eV was assigned to the external –OH group or the water molecule adsorbed on the surface, and the other O 1s peak appearing at 532.4 eV corresponded to the C–O bonds in the GO.54 Fig. 4(e)–(g) show the high-resolution C 1s spectra of the Bi7O9I3/GO composites and pure GO. Three carbon species were displayed mainly in the C 1s spectra of pure GO and the Bi7O9I3/GO composites including unoxidized carbons (sp2 carbon), C–O, and CO. As seen in Fig. 4(f) and (g), three different chemically shifted components are visible, which could be deconvoluted into sp2 carbons in aromatic rings (284.4 eV) and C atoms bonded to oxygen (C–O 286.8 eV) and carbonyls (C
O, 288.7 eV).56,57
Tables 3 and 4 and Fig. 8 illustrate the degradation efficiency as a function of reaction time; the removal efficiency was significantly enhanced in the presence of 0.05–0.15 g Bi7O9I3/GO and 0.10 g Bi4O5I2/GO. To further understand the reaction kinetics of CV degradation, the apparent pseudo-first-order model,59 ln(Co/C) = kt, was applied to the experiments. Through the first-order linear fit of the data shown in Fig. 8 and Table 3, the k values of 0.05 g Bi7O9I3/GO and 0.10 g Bi4O5I2/GO were obtained as the maximum degradation rates of 3.51 × 10−1 and 3.22 × 10−1 h−1 by using the first-order linear fit of the data, which are much higher than those of the other composites. The 0.05 g Bi4O5I2/GO composite had a larger SBET and pore volume (Table S2†). However, the results in Table S2† show that the 0.05 g Bi7O9I3/GO sample—which does not show the highest SBET—did represent the highest photocatalytic activity (k = 3.22 × 10−1 h−1) among the samples, suggesting that the changes in the photocatalytic activity resulted from both the SBET and the BiOxIy/GO composites. Table 5 shows a comparison of the rate constants of the different photocatalysts. The order of rate constants was as follows: Bi4O5I2/GO > Bi7O9I3/GO > Bi4O5I2 > Bi7O9I3 > Bi7O9I3/GO > BiOI/GO > BiOI > Bi5O7I > GO. The photocatalytic activity of the Bi4O5I2/GO (or Bi7O9I3/GO) heterojunctions reached a maximum rate constant of 0.351 (or 0.322) h−1, which is 1.8 (or 1.7) times higher than that of Bi4O5I2 (or Bi7O9I3), 6–7 times higher than that of BiOI/GO, and 119–130 times higher than that of BiOI. Thus, the BiOxIy/GO composites may also play a role in enhancing the photocatalytic activity.
BiOxIy/GO | ||||||||||
---|---|---|---|---|---|---|---|---|---|---|
Graphene oxide weight (g) | pH | |||||||||
1 | 4 | 7 | 10 | 13 | ||||||
k (h−1) | R2 | k (h−1) | R2 | k (h−1) | R2 | k (h−1) | R2 | k (h−1) | R2 | |
0 | 0.027 | 0.940 | 0.198 | 0.949 | 0.190 | 0.954 | 0.191 | 0.983 | 0.004 | 0.973 |
0.005 | 0.020 | 0.976 | 0.252 | 0.967 | 0.255 | 0.973 | 0.183 | 0.973 | 0.012 | 0.985 |
0.01 | 0.051 | 0.942 | 0.256 | 0.931 | 0.277 | 0.982 | 0.192 | 0.970 | 0.005 | 0.910 |
0.05 | 0.050 | 0.940 | 0.218 | 0.953 | 0.351 | 0.953 | 0.191 | 0.955 | 0.065 | 0.914 |
0.10 | — | — | 0.322 | 0.924 | 0.286 | 0.976 | 0.272 | 0.904 | 0.014 | 0.944 |
0.15 | — | — | 0.080 | 0.913 | 0.311 | 0.913 | 0.319 | 0.911 | 0.034 | 0.865 |
BiOxIy/GO | ||||||||||
---|---|---|---|---|---|---|---|---|---|---|
Graphene oxide weight (g) | pH | |||||||||
1 | 4 | 7 | 10 | 13 | ||||||
k (h−1) | R2 | k (h−1) | R2 | k (h−1) | R2 | k (h−1) | R2 | k (h−1) | R2 | |
0 | 0.005 | 0.615 | 0.250 | 0.947 | 0.239 | 0.982 | 0.314 | 0.939 | 0.010 | 0.816 |
0.005 | 0.007 | 0.550 | 0.145 | 0.910 | 0.203 | 0.818 | 0.119 | 0.986 | 0.003 | 0.952 |
0.01 | 0.006 | 0.912 | 0.181 | 0.931 | 0.167 | 0.989 | 0.170 | 0.777 | 0.001 | 0.631 |
0.05 | 0.008 | 0.618 | 0.135 | 0.807 | 0.209 | 0.912 | 0.134 | 0.984 | 0.005 | 0.574 |
0.10 | — | — | 0.130 | 0.868 | 0.201 | 0.916 | 0.136 | 0.927 | 0.002 | 0.486 |
0.15 | — | — | 0.019 | 0.765 | 0.128 | 0.869 | 0.110 | 0.839 | 0.003 | 0.617 |
![]() | ||
Fig. 8 Photodegradation of CV as a function of irradiation time for the different Bi7O9I3/GO samples (pH = 7). |
Photocatalyst | Rate constant, k (h−1) |
---|---|
GO | 0 |
BiOI | 0.027 |
Bi4O5I2 | 0.198 |
Bi7O9I3 | 0.191 |
Bi5O7I | 0.004 |
BiOI/GO | 0.051 |
Bi4O5I2/GO | 0.322 |
Bi7O9I3/GO | 0.351 |
Bi5O7I/GO | 0.065 |
However, the photocatalytic activity over the BiOxIy/GO nanocomposites decreased when the GO content exceeded approximately 0.15 g. This decrease may be attributed to the joint effect between the excellent charge transfer capability of GO and its detrimental effect on visible light absorption. The durability of the 0.05 g Bi7O9I3/GO composite was evaluated by recycling the used catalyst. After each cycle, the catalyst was collected by centrifugation. No apparent decline was observed in the photocatalytic activity when CV was removed in the 3rd cycle; even during the fifth run, the decline in the photocatalytic activity was 5% (Fig. 9(a)). The used 0.05 g Bi7O9I3/GO composite was also examined by XRD and no detectable difference was observed between the as-prepared and the used samples (Fig. 9(b)); hence, the 0.05 g Bi7O9I3/GO composite had excellent photostability.
![]() | ||
Fig. 9 (a) Cycling runs in the photocatalytic degradation of CV in the presence of Bi7O9I3/GO (pH = 7, 0.15 g GO); (b) XRD of the powder sample before and after the degradation reaction. |
Photocatalysts are excited to generate electron–hole pairs directly after the illumination in the photocatalytic process. Photocatalytic efficiency depends mainly on the recombination rate or the lifetime of the photogenerated electron–hole pairs. The faster the recombination occurs, the shorter the chemical reaction time is. Therefore, PL spectroscopy was utilized for investigating the recombination rate of the photogenerated electron–hole pairs.60 To investigate the separation capacity of the photogenerated carriers in the heterostructures, the PL spectra of 0.005–0.15 g Bi7O9I3/GO and Bi7O9I3 were measured; the results are shown in Fig. 10. A strong emission peak at approximately 530 nm appeared for the as-prepared samples, which could have originated from the direct electron–hole recombination of band transitions. However, the characteristic emission peak within 530 nm for the 0.005–0.15 g Bi7O9I3/GO photocatalysts is low intensity indicated that the recombination of photogenerated charge carriers was greatly inhibited. The efficient separation of charge could increase the lifetime of the charge carriers and enhance the efficiency of interfacial charge transfer to the adsorbed substrates, thus improving the photocatalytic activity.55 The lowest relative PL intensity of the 0.15 g Bi7O9I3/GO composite, as shown in Fig. 10, suggests that it possessed a lower recombination rate of electron–hole pairs, resulting in the lower photocatalytic activity, as shown in Fig. 8. The PL results confirm the importance of the composites in hindering the recombination of electrons and holes and explain the reason for the increasing photocatalytic performance of the BiOxIy/GO composites.
Presumably, the enhanced photocatalytic activities of BiOxIy/GO composites could be ascribed to a synergistic effect including a high BET surface area, the formation of the composites (or heterojunction), a layered structure, and the low energy band structure. In the absence of photocatalysts, CV could not be degraded under visible light irradiation; the superior photocatalytic ability of BiOxIy/GO may be ascribed to its efficient utilization of visible light and the high separation efficiency of the electron–hole pairs within its composites.
Various primary active species, such as hydroxyl radicals, holes, superoxide radicals, hydrogen radicals (H˙), and singlet oxygen, can be generated during photocatalytic decomposition processes in UV-vis/semiconductor systems.63 Dimitrijevic et al.64 proposed that water, which dissociated both on the surface of TiO2 and in subsequent molecular layers, had a three-fold role as (i) a stabilizer of charges, preventing electron–hole recombination; (ii) an electron acceptor, forming H atoms in a reaction of photogenerated electrons with protons on the surface, –OH2+; and (iii) an electron donor, resulting in the reaction of water with photogenerated holes to give ˙OH radicals.
Theoretically, GR nanosheets with 100% sp2-hybridized carbon atoms have a high electrical conductivity (250000 cm2 V−1 s−1) for storing and shuttling electrons, and a high surface area (2630 m2 g−1). When GR is combined with other materials, electrons would flow from one material to the other (from a higher to a lower Fermi level) to align the Fermi energy levels at the interface of two materials.65
Wang et al.66 revealed that O2−˙ and ˙OH were the main reactive species for the degradation of rhodamine B with BiVO4/RGO. Bai et al.67 reported that active species trapping measurements, superoxide radicals (O2−˙) and hydroxy radicals played a crucial role during the catalytic process in the methylene blue degradation process using ZnWO4/graphene hybrids.67 Shenawi-Khalil et al.68 reported that ˙OH radicals were generated through the multistep reduction of O2−˙. The generation of O2−˙ could not only inhibit the recombination of photoinduced charge carriers but also benefit the dechlorination of chlorinated phenol derivatives. The hydroxyl radical HO˙ might only be formed through an e− → O2−˙ → H2O2 → ˙OH route. However, the ˙OH radical was formed through the multistep reduction of O2−˙ in the system.61 Zhu et al.69 reported that the g-C3N4/BiOBr-mediated photodegradation of methylene blue molecules was attributed mainly to the oxidation action of the generated O2−˙ radicals and partly to the action of h+ through the direct hole oxidation process. According to a previous study,67 a photocatalytic process was governed mainly by O2−˙ rather than by ˙OH, e−, or h+. In a previous study,7 the CV photodegradation by BiOmXn/BiOpYq (X, Y = Cl, Br, I) under visible light was dominated by oxidation, with O2−˙ being the main active species and ˙OH and h+ being the minor active species. On the basis of the aforementioned references, the probability of forming ˙OH should be much lower than that for O2−˙; however, ˙OH is an extremely strong and nonselective oxidant, which leads to a partial or complete mineralization of several organic chemicals.
Fig. 11(a) and (b) show not only the six characteristic peaks (weak) of the DMPO-O2−˙ adducts but also the four characteristic peaks (weak) of the DMPO-˙OH adducts (1:
2
:
2
:
1 quartet pattern) under visible light irradiating the 0.05 g Bi7O9I3/GO composite dispersion. Fig. 11(a) and (b) indicate that no electron paramagnetic resonance (EPR) signal was observed when the reaction was performed in the dark, whereas the signals with intensities corresponding to the characteristic peaks of DMPO-˙OH and DMPO-O2−˙ adducts22 were observed during the reaction process under visible light irradiation, suggesting that O2˙− and hydroxyl radicals (˙OH) as active species were formed in the presence of 0.05 g Bi7O9I3/GO composites and oxygen under visible light irradiation.
To re-evaluate the effect of the active species during the photocatalytic reaction, a series of quenchers were introduced to scavenge the relevant active species. As shown in Fig. 11(c), the photocatalytic degradation of CV was not affected by the addition of IPA, whereas the degradation efficiency of BQ, IPA, and AO quenching evidently decreased when compared with that of no quenching. O2˙− was a major active species and h+ and hydroxyl radicals were minor active species in the process of photocatalytic degradation of CV. Therefore, the quenching effects of different scavengers and EPR indicated that the reactive superoxide radical played a major role, and holes and hydroxyl radicals played a minor role in CV photocatalytic degradation.
On the basis of the aforementioned experimental results, a detailed pathway of decomposition is illustrated in Fig. 12. Once the electron reached the conduction band of BiOxIy, it induced the formation of active oxygen species, which caused the decomposition of the CV dye. Except for the photodegradation of CV through the route of BiOxIy/GO-mediated and photosensitized processes, another type of photocatalytic route accounted for the enhanced photocatalytic activity. Both photosensitized and photocatalytic processes proceeded concurrently (Fig. 12). However, under the photosensitized and photocatalytic reaction conditions, O2˙− radicals were formed by the reaction of photogenerated and photosensitized electrons with oxygen gas on the photocatalyst surface; hydroxyl radicals were also produced by the reaction of O2˙− radicals with H+ ions and h+ holes with OH− ions (or H2O). The hydroxyl radicals were produced subsequently.70 This cycle continuously occurred when the system was exposed to visible light irradiation;8 after several cycles of photooxidation, the decomposition of CV by the generated oxidant species can be expressed by eqn (8) and (9).
CV + O2−˙ → decomposed compounds | (8) |
CV + OH˙ → decomposed compounds | (9) |
Hydroxylated compounds were identified for the photocatalytic degradation of CV under visible light-induced semiconductor systems.7 Under UV light irradiation, N-dealkylation processes were preceded by the formation of a nitrogen-centered radical, and the destruction of the dye chromophore structure was preceded by the generation of a carbon-centered radical in the photocatalytic degradation of the CV dye.8,70,71 All the intermediates identified in these two researched topics had the same results under UV or visible light irradiation. Undoubtedly, the major oxidant was ˙OH radicals, not O2−˙ radicals. The reaction pathways of BiOxIy/GO-mediated photocatalytic processes proposed in this study should offer some guidance for applications in the decomposition of dyes.
![]() | ||
Fig. 13 HPLC chromatogram of the intermediates with 24 h of photocatalytic reaction, recorded at (a) 580 nm, (b) 350 nm, and (c) 300 nm. |
The absorption spectra of each intermediate in the absorption spectral region are measured corresponding to the peaks. The data show that the absorption spectral bands shift hypsochromically from 588.2 nm to 543.4 nm (A–J), 376.4 to 337.5 nm (a–f), and 309.7 to 278.1 nm (α–γ) in Table S3 of the ESI.† As can be seen, the 19 compounds can be distributed into three different aromatic groups: (i) intermediates of the N-demethylation of N,N,N′,N′,N′′,N′′-hexamethylpararosaniline (CV; A), (ii) intermediates of the N-demethylation of 4-(N,N-dimethylamino)-4′-(N′,N’-dimethylamino)benzophenone (a), and (iii) intermediates of the N-demethylation of 4-(N,N-dimethylamino)phenol (α), respectively. This hypsochromic shift of the absorption band was possible due to the formation of a series of N-demethylated intermediates in a stepwise manner. For example, the λmax of A, B, C, D, E, F, G, H, I, and J are 588.2, 581.1, 573.4, 579.5, 566.7, 570.1, 561.8, 566.2, 554.1 and 543.4 nm, respectively; the λmax of a, b, c, d, e, and f are 376.4, 366.4, 364.4, 358.8, 357.3, and 337.5 nm, respectively; and the λmax of α, β, and γ are 309.7, 283.6, and 278.1 nm, respectively. Similar phenomena were observed during the photocatalytic degradation of CV.6,12 The wavelength shift depicted in Table S3 of the ESI† is caused by the N-demethylation of CV because of the attack by one of the ˙OH radicals on the N,N-dimethyl group. Correspondingly, the N-demethylated intermediates of the photocatalytic reactions were further identified using the HPLC-ESI mass spectrometric method due to a similar situation in the intermediates category. Therefore, we used the higher removal rate of the Bi7O9I3/GO composite to identify the intermediates by HPLC-ESI mass spectrometry.
The N-demethylated intermediates were further identified using the HPLC-ESI mass spectrometric method. The molecular ion peaks appeared in the acid forms of the intermediates. As shown in Table S3,† the mass spectral analysis confirmed the components A (m/z = 372.18), B (m/z = 358.14), C (m/z = 344.10), D (m/z = 344.09), E (m/z = 330.10), F (m/z = 330.36), G (m/z = 316.11), H (m/z = 316.11), I (m/z = 302.06), J (m/z = 288.07), a (m/z = 269.05), b (m/z = 255.06), c (m/z = 240.92), d (m/z = 240.98), e (m/z = 226.84), f (m/z = 213.06), α (m/z = 138.16), β (m/z = 124.03), and γ (m/z = 110.14). These species correspond to three pairs of isomeric molecules with two to four less methyl groups than CV. For example, B is formed by the removal of a methyl group from two different sides of the CV molecule, while the other corresponding isomer in this pair, D, is produced by the removal of two methyl groups from the same side of the CV structure. In the second pair of isomers, E is formed by the removal of three methyl groups from each side of the CV molecule, while the other isomer in this pair, F, is produced by the removal of two methyl groups from one side of the CV structure while one more methyl group was removed from the other side of the CV structure. In the third pair of isomers, H is formed by the removal of two methyl groups from two different sides of the CV molecule, while the other isomer in this pair, G, is produced by the removal of two methyl groups from the same side of the CV structure and by removal of a methyl group from the remaining two sides of the CV structure. Because the polarities of the D, F and H species are greater than those of the C, E and G intermediates, the latter were eluted after the D, F and H species, respectively. Since the two N-methyl groups are stronger auxochromic moieties than the N,N-dimethyl groups and the amino group, the maximal absorption of the D, F and H intermediates was anticipated to occur at wavelengths longer than the band position of the C, E and G species, respectively (Table S3 of the ESI†). On the other hand, the oxidation process is initiated by hydroxylation of the central carbon of CV. For example, the intermediates 4-(N,N-dimethylamino)-4′-(N′,N’-dimethylamino)benzophenone (a) and 4-(N,N-dimethylamino)phenol (α) are formed by an ˙OH radical attack on the conjugated structure, yielding a carbon-centered radical, which is subsequently attacked by molecular oxygen, leading to cleavage of the CV conjugated chromophore structure. Afterwards, compounds a and α were possibly formed by a series of N-demethylated intermediates in a stepwise manner. This was also reported for the photocatalytic reaction of CV.6,11,12
![]() | ||
Fig. 14 Proposed degradation mechanism of CV under photocatalytic processes, followed by the identification of several intermediates by HPLC-ESI mass spectral techniques. |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra12482h |
This journal is © The Royal Society of Chemistry 2016 |