Yan Zhou,
Sushan Ren,
Qimei Dong,
Yingying Li and
Hanming Ding*
School of Chemistry and Molecular Engineering, East China Normal University, 500 Dongchuan Road, Shanghai 200241, China. E-mail: hmding@chem.ecnu.edu.cn
First published on 13th October 2016
A novel nanocomposite, Bi nanorods and Bi2WO6 nanosheets supported on reduced graphene oxide (Bi/Bi2WO6/rGO), was synthesized via a facile one-pot solvothermal method. In the reaction process, Bi2WO6 nanosheets and Bi nanorods were grown in situ on the rGO sheets, which were simultaneously achieved by the reduction of GO. Such a synthetic strategy can form effective close interfacial contacts and strong interactions among Bi2WO6, Bi and rGO, leading to efficient separation and transfer of photogenerated electron–hole pairs. As a result, the ternary plasmonic photocatalyst exhibits a much higher photocatalytic activity than pure Bi2WO6 and the binary composites in the photocatalytic degradation of rhodamine B and p-chlorophenol under visible light irradiation, which could be ascribed to the synergic effects of the improved electron–hole pair separation efficiency, enhanced visible-light harvesting and the good adsorptive capacity toward dye molecules.
As a new visible light-responding non-titania-based photocatalyst, Bi2WO6 with a unique layered structure has attracted increasing interest recently.9–11 However, there are two main drawbacks that limit the application of the Bi2WO6 photocatalyst. First, pure Bi2WO6 can only absorb light with wavelengths shorter than 420 nm.12,13 Secondly, the high recombination of photogenerated charges after light absorption severely limits the light energy conversion efficiency.14,15 It was suggested that if a high optical absorption could be maintained while charge recombination could be significantly suppressed, a good photocatalytic activity would be anticipated. To solve these problems, one way is to fabricate heterojunction composite photocatalysts by coupling of another material with the appropriate Fermi energy. The recombination of photogenerated electron–hole pairs can thus be effectively suppressed, resulting in increased probability of the photogenerated charges being available for photocatalytic reactions.16,17 Reduced graphene oxide (rGO) has recently attracted tremendous attention owing to its fascinating electrical and chemical properties.18,19 When rGO is hybridized with a photocatalyst, the excellent electronic conductivity of rGO will promote charge transfer through its conjugated structure, inhibiting the recombination of the photoexcited electron–hole pairs20,21 Furthermore, rGO nanosheets can act as adsorptive centers due to their large π-conjugation system and 2D planar structure, which make dye molecules adsorb easily onto their surface via strong π–π interactions between rGO and the dye molecules.22
Lately, bismuth (Bi) has been observed to exhibit surface plasmon resonance (SPR) properties.23 Compared to the general noble metals such as Au and Ag, Bi is much cheaper and easily prepared. It is documented that several Bi-modified photocatalysts, such as Bi/g-C3N4 and Bi/Bi2O2CO3 exhibited highly promoted photocatalytic performance, which is ascribed to the SPR effect of Bi.24,25 Considering the salient characteristics of Bi, Bi2WO6, and rGO, it prompts us to consider whether we can use a facile method to prepare a Bi/Bi2WO6/rGO triple nanocomposite, in which synergic effects could improve the photocatalytic activity.
In this study, we have developed a one-pot hydrothermal method to prepare a Bi/Bi2WO6/rGO nanocomposite, in which the formation of Bi2WO6, reduction of graphene oxide (GO) into rGO, and reduction of Bi3+ ions to Bi metal are simultaneously accomplished in the same synthetic system. Notably, besides being a solvent, ethylene glycol (EG) also acts as a mild reductant for the reduction of GO and Bi3+ ions, and therefore no additional reductants are required. The Bi/Bi2WO6/rGO nanocomposite exhibits a high photocatalytic activity to degrade rhodamine B (RhB) solutions under visible light irradiation. And the photocatalytic mechanism has been proposed as well.
![]() | ||
Fig. 1 XRD patterns of (a) GO, (b) Bi, (c) Bi2WO6, (d) Bi/Bi2WO6, (e) Bi2WO6/rGO, and (f) Bi/Bi2WO6/rGO-3. Triangles and solid dots refer to Bi and Bi2WO6, respectively. |
As shown in Fig. 2, the Raman peak at 303 cm−1 is assigned to the translation modes involving simultaneous motions of Bi3+ and WO66−.28 These two peaks at 763 and 817 cm−1 are attributed to antisymmetric and symmetric Ag modes of the O–W–O group terminus, respectively.29 Compared to Bi2WO6, the intensity of Raman peaks of Bi/Bi2WO6 composites are slightly enhanced. This phenomenon could be attributed to the surface-enhanced Raman scattering effect of metallic Bi.30 As for rGO, there are two typical Raman bands at ∼1611 cm−1 (G band) and at ∼1359 cm−1 (D band) for the graphitized structure.31 In the Raman spectrum of Bi/Bi2WO6/rGO-3, these two characteristic bands were also observed. In comparison with GO (ID/IG = 0.98), an increased D/G intensity ratio (ID/IG = 1.03) was observed, which revealed that GO has been partially reduced after the solvothermal reaction.32 Furthermore, the positions of D band and G band in the ternary composites were slightly shifted (inset of the Fig. 2), due to the reduction of GO and the combination of Bi and Bi2WO6 with rGO.33,34
![]() | ||
Fig. 2 Raman spectra of (a) GO, (b) Bi2WO6, (c) Bi/Bi2WO6 and (d) Bi/Bi2WO6/rGO-3. Inset: the enlarged Raman spectra of GO and Bi/Bi2WO6/rGO-3. |
The chemical states and the surface composition of the Bi/Bi2WO6/rGO-3 sample has been analyzed by XPS (Fig. 3). In the full XPS spectrum of Fig. 3A, different binding energies were assigned to W 4f, Bi 4f, W 4d, Bi 4d, C 1s, O 1s, and Bi 4p state emission peaks. These peaks at 34.9 and 35.1 eV in Bi/Bi2WO6/rGO-3 correspond to W 4f5/2 and W 4f7/2 (Fig. 3B), both of which can be assigned to the W6+ oxidation state.35 The O 1s peaks for Bi/Bi2WO6/rGO-3 (Fig. 3C) can be deconvoluted into two bands at 529.8 and 530.7 eV, which are associated with hydroxyl (OH) groups and oxygen species in the lattice oxygen, respectively.36 Two main peaks with binding energies of 158.8 eV and 164.1 eV can be observed in Fig. 3D, which can be ascribed to the Bi3+ 4f7/2 and Bi3+ 4f5/2 binding energies, respectively.37 Moreover, apart from the two peaks discussed above, there are two peaks at 162.5 eV and 157.2 eV with low intensity, which can be ascribed to metallic Bi.38 As shown in Fig. 3E, the peak at 288.7 eV for the CO bonding exists, but its peak intensity is very low. To illustrate the reduction degree of the GO sheets in the Bi/Bi2WO6/rGO-3 sample, the peak area ratios of C
O bonds to the total area are calculated to be 0.16 by the XPS C 1s peak area analysis, proving the transformation from GO to rGO.39,40 Therefore, the results confirm the successful reduction of Bi3+ and GO into metal Bi and rGO via the facile hydrothermal process.
![]() | ||
Fig. 3 XPS spectra of the Bi/Bi2WO6/rGO-3 sample: (A) full spectrum, (B) W 4f, (C) O 1s, (D) Bi 4f, and (E) C 1s. |
TEM images showing the morphology and microstructure of the as-prepared Bi/Bi2WO6/rGO-3 are presented in Fig. 4. It can be seen that the rGO sheets exhibit a typical rippled and crumpled structure and the Bi/Bi2WO6 composites spread over the rGO surface (Fig. 4a). The apparent contrast between the darker nanorods and brighter nanosheets, as illustrated in Fig. 4b, suggests the formation of Bi nanorods and Bi2WO6 nanosheets on the surface of rGO. The Bi nanorods appear distinctly darker than the Bi2WO6 due to the higher electron density, which could be further confirmed by HRTEM (Fig. 4c).41–43 The Bi nanorods with an average length of 50 nm have been observed in the TEM images (Fig. 4b and S6a†). As can be observed in Fig. 4c, the fringe interval of 0.315 nm is in accordance with the interplanar spacing of the (131) plane of Bi2WO6, whereas the periodic fringe spacing of 0.325 nm can be indexed as the (102) plane of the hexagonal Bi. To further obtain information about the Bi structure, the special region of Bi nanorods was analyzed (Fig. 4d). The spacing of 0.227 nm between neighboring lattice fringes corresponds to the distance between two (110) planes, indicating [100] as the growth direction for the Bi nanorods.44 These results agree well with the XRD results. Moreover, the stacking width is calculated to be 0.37 nm for rGO layers corresponding to the spacing of the (002) lattice planes.45,46 This spacing is very close to pristine graphite, indicating that GO has been well reduced to graphene (rGO).
Fig. 5 shows the nitrogen adsorption–desorption isotherms for samples of Bi2WO6, Bi/Bi2WO6, Bi2WO6/rGO and Bi/Bi2WO6/rGO-3. The typical type IV isotherms and hysteresis loops of the sample are characteristic of mesoporous materials.37,38 The BET surface areas of the as-prepared Bi2WO6, Bi/Bi2WO6, Bi2WO6/rGO, and Bi/Bi2WO6/rGO-3 are 27.87, 39.12, 67.42, and 67.67 m2 g−1, respectively. Notably, the higher surface area of Bi/Bi2WO6/rGO-3 is attributed to the presence of rGO nanosheets in the nanocomposite. Moreover, the pore volume determined by the Barrett–Joyner–Halenda (BJH) model indicated broad distributions with pore volumes of 24.65, 35.14, 68.43, and 69.26 cm3 g−1 for Bi2WO6, Bi/Bi2WO6, Bi2WO6/rGO, and Bi/Bi2WO6/rGO-3, respectively. The pore size distribution of Bi/Bi2WO6/rGO-3 is shown in the inset in Fig. 5, which exhibits an average pore diameter of 20 nm. Owing to its high surface area and large pore volume, the mesoporous photocatalyst will provide efficient adsorption sites and enhance transportation of charge carriers, which would lead to an improvement of the photocatalytic activity.
![]() | ||
Fig. 5 Nitrogen adsorption–desorption isotherms of various samples. Inset: the pore size distribution of Bi/Bi2WO6/rGO-3. |
Bi(NO3)3 → Bi3+ + 3NO3− | (1) |
Na2WO6 → 2Na+ + WO42− | (2) |
2Bi3+ + WO42− + 2H2O → 4H+ + Bi2WO6 | (3) |
![]() | (4) |
![]() | (5) |
2HOCH2CH2OH → 2CH3CHO + 2H2O | (6) |
6CH3CHO + 2Bi3+ → 3CH3COCOCH3 + Bi + 6H+ | (7) |
The overall synthesis can be divided into four consecutive stages. At the early stage of the reaction process, when Bi(NO)3 is dissolved in the GO solution (eqn (1)), the positively charged Bi3+ could be uniformly adsorbed onto the surfaces of the GO sheets, which are negatively charged due to the existence of OH and CO functional groups (Scheme 1A and B). After Bi3+ ions have been adsorbed, the reaction between Bi3+ and EG takes place on the surface of GO, forming EG–Bi3+ complexes via the coordination between Bi3+ ions and OH groups (eqn (4) and Scheme 1C).47 Then, in the following solvothermal treatment at high operating temperature, the complex is gradually decomposed to release the Bi3+ ions, which then react with WO42− ions to form Bi2WO6 nanosheets on the surface of the graphene components (eqn (5)).48 Moreover, metal Bi and rGO can be formed simultaneously (eqn (6) and (7), Scheme 1D) during the hydrothermal treatment. When Bi(NO3)3 is in excess in the reaction, there are still many residual Bi3+ ions absorbed on the surface of GO after the formation of Bi2WO6. These residual Bi3+ ions can be reduced into Bi nanorods by the reductant EG in the presence of PVP. Therefore, the content of Bi nanorods can be easily tuned just by varying the precursor ratios. It has been reported that only urchin-like bismuth structures were obtained by self-assembly of Bi nanorods in pure EG solution, in which a typical coordinative reduction course was suggested.49 It is believed that the anisotropic effects of crystal and the face inhibitor function of PVP play crucially important roles in the formation of 1D nanostructures.50 In the solvothermal reaction process, PVP molecules could preferentially adsorb onto the primary surfaces of Bi crystals through O–Bi bonding, as reported in the previous work.44 The polymer molecules adsorbed on some surfaces of the Bi crystals could significantly decrease their growth rates and lead to highly anisotropic growth.51 Therefore, the isolated Bi nanorods were formed in the presence of PVP under appropriate reaction conditions.
UV-Vis diffuse reflectance spectra (DRS) of the as-prepared samples are shown in Fig. 6. Compared with that of pure Bi2WO6, the light-harvesting ability of Bi/Bi2WO6 is apparently enhanced, ranging from the UV to visible light region. This enhancement is attributed to the SPR effect of plasmonic Bi.52 In addition, the absorption of Bi/Bi2WO6 is gradually intensified with increasing content of Bi (Fig. S3†). For Bi/Bi2WO6/rGO, the continuous background absorption from 400–800 nm is enhanced due to the black body effect of graphene.53 Moreover, this observation also clearly indicates that more visible light can be absorbed in the prepared Bi/Bi2WO6/rGO nanocomposite to generate more electron–hole pairs, and thus its photocatalytic activity will be improved under visible light irradiation. This hypothesis was confirmed by measuring the degradation of RhB over the obtained samples under the same conditions. With the increasing amount of rGO, the Bi/Bi2WO6/rGO composites show a continuously enhanced visible-light absorption in the range of 400–800 nm, which is in agreement with the previous report.54 Therefore, the incorporation of Bi nanorods and rGO sheets increased the light absorption of the ternary composites in the visible region.
![]() | ||
Fig. 6 UV-Vis diffuse reflectance spectra of (a) Bi2WO6, (b) Bi/Bi2WO6, (c) Bi2WO6/rGO, (d) Bi/Bi2WO6/rGO-1, (e) Bi/Bi2WO6/rGO-2, (f) Bi/Bi2WO6/rGO-3, and (g) Bi/Bi2WO6/rGO-4. |
The degree of charge recombination can be evaluated by photoluminescence (PL) spectroscopy (Fig. 7). The lower PL intensity reflects the lower recombination rate of photogenerated electrons and holes.55 Evidently, Bi/Bi2WO6 shows a significantly diminished PL intensity in comparison with pure Bi2WO6, which suggests that the deposition of Bi results in a remarkable decline in the recombination rate of photogenerated electron–hole pairs.25 In Bi2WO6/rGO, photogenerated electrons could be transferred efficiently from Bi2WO6 to rGO due to the high charge carrier mobility of the rGO sheets, leading to a low charge recombination rate. The PL results demonstrated that the rGO layers with a two-dimensional π-conjugated structure could serve as an effective electron-accepting material.56 A more efficient separation of electron–hole pairs was observed in Bi/Bi2WO6/rGO-3, which may be due to the charge transfer between Bi/Bi2WO6 and rGO. These results clearly indicate that the recombination of photogenerated charge carriers of Bi2WO6 is greatly restrained by the coupling of Bi and rGO.
Fig. 8 shows the transient photocurrent responses of the obtained samples. The photocurrents were stable and reversible at light-on and light-off for all samples. In comparison with Bi2WO6, Bi/Bi2WO6, and Bi2WO6/rGO, Bi/Bi2WO6/rGO-3 composite exhibited an increased current density, about 1.6, 1.2, and 1.15 times that of Bi2WO6, Bi/Bi2WO6, and Bi2WO6/rGO, respectively. The improvement of photocurrent indicated an enhanced photogenerated electron–hole separation efficiency, which is in good agreement with the PL measurements.
![]() | ||
Fig. 8 Photocurrent responses of (a) Bi2WO6, (b) Bi/Bi2WO6, (c) Bi2WO6/rGO, and (d) Bi/Bi2WO6/rGO-3. |
The photocatalytic activities of the samples were evaluated by the degradation of RhB under visible light irradiation (λ > 420 nm), as shown in Fig. 9. A blank control test (without photocatalyst) confirmed that the self-degradation of RhB can be regarded as negligible. The Bi/Bi2WO6 composite has a higher photocatalytic activity than that of Bi2WO6, which is due to the SPR effect of Bi. In addition, the photocatalytic activity was further enhanced by increasing the Bi amount in the composites (Fig. S4†), which is consistent with the results of DRS. Compared with pure Bi2WO6 and Bi/Bi2WO6, the Bi/Bi2WO6/rGO photocatalysts exhibited much higher photocatalytic activity (Fig. 9A). About 99.9% of RhB molecules were decomposed in 120 min when Bi/Bi2WO6/rGO-3 was used as the photocatalyst, whereas only 49% and 83% of the RhB molecules were degraded when pure Bi2WO6 and Bi/Bi2WO6 were employed, respectively. Moreover, the photocatalytic activity of the Bi/Bi2WO6/rGO composites increases with increasing the rGO content, and Bi/Bi2WO6/rGO-3 shows the highest degradation efficiency. However, when more GO was added, the photocatalytic activity of the Bi/Bi2WO6/rGO photocatalyst was found to decrease. It is probably because the dense black layers of excessive rGO shield the light irradiation on Bi and Bi2WO6, which can weaken the photocatalytic activity of the composite. Furthermore, excessive rGO also can act as a recombination center, which then decreases the efficiency of charge separation.57
In order to directly show the photocatalytic activity enhancement of Bi/Bi2WO6/rGO, the Langmuir–Hinshelwood kinetics model was applied to calculate the apparent pseudo-first order rate constant k.58 As can be seen in Fig. 9B, the rate constants corresponding to Bi/Bi2WO6, Bi2WO6/rGO, Bi/Bi2WO6/rGO-1, Bi/Bi2WO6/rGO-2, Bi/Bi2WO6/rGO-3 and Bi/Bi2WO6/rGO-4 were estimated to be 0.0111, 0.0099, 0.0115, 0.0203, 0.0355 and 0.0147 min−1, which are about 2.27, 2.02, 2.35, 4.14, 7.24 and 3 times as much, respectively, as that over Bi2WO6 (0.0049 min−1). These results clearly indicated that introducing the appropriate amount of Bi and rGO into the Bi2WO6 photocatalyst system can greatly enhance its photocatalytic activity.
It is usually accepted that the degradation of RhB over the photocatalyst would occur partially via a photosensitization pathway. In order to eliminate the photosensitization effect, colorless p-chlorophenol (4-CP) was used as the model pollutant. The photocatalytic degradation results are shown in Fig. 10. In contrast to Bi2WO6, Bi/Bi2WO6, and Bi2WO6/rGO, an enhanced photocatalytic activity was observed for the Bi/Bi2WO6/rGO-3 composite. As shown in Fig. 10A, a similar trend to the photocatalytic degradation of RhB was observed under visible light irradiation. The Bi/Bi2WO6/rGO-3 composite exhibits the highest photocatalytic performance among all the tested samples for RhB and 4-CP removal. Fig. S5† shows the evolution of UV-Vis spectra for RhB and 4-CP solution in the presence of Bi/Bi2WO6/rGO-3 under visible light irradiation. The UV-Vis peak intensity of RhB at 553 nm decreased gradually with irradiation time, whereas its position remained unchanged, suggesting that the RhB molecules were directly destroyed through the cleavage of the chromophore ring structure rather than a de-ethylation process.59 The UV-Vis intensity of 4-CP also decreased with irradiation time, but slower as compared with that of RhB. Only 52% 4-CP was removed, whereas almost 99% RhB was degraded after 120 min. Thus, the degradation efficiency for 4-CP is lower than that for RhB (Fig. 10B), which implies that photosensitization plays a role in the RhB degradation.
As is known, various active species (˙OH, ˙O2−, and h+) generated in the photocatalytic processes play key roles. Therefore, trapping experiments were carried out to explore the main active species. Isopropyl alcohol (IPA), p-benzoquinone (BQ), and triethanolamine (TEOA) were adopted as the traps for ˙OH, ˙O2−, and h+, respectively.60 Fig. 11 shows the effect of different scavengers on the photocatalytic degradation of RhB and 4-CP over Bi/Bi2WO6/rGO-3. It can be easily seen that the addition of IPA does not cause deactivation of the Bi/Bi2WO6/rGO-3 photocatalyst in either case, indicating that ˙OH was not the major active species involved in the photocatalytic processes. However, the photocatalytic performance of Bi/Bi2WO6/rGO-3 significantly decreases on the addition of TEOA and BQ. Thus, ˙O2− radicals and h+ were proven to be the dominant active species in the photocatalytic degradation of RhB and 4-CP over the Bi/Bi2WO6/rGO-3 composite photocatalyst under visible light irradiation. According to the previous report, the CB electrons of Bi2WO6 possess a poor reducing power owing to their more positive potential (+0.3 V, vs. SHE) than the one-electron oxygen reduction (O2 + H+ ± e− = HO2, −0.046 V vs. SHE).61 It is plausible that the conduction band potential of Bi2WO6 is incapable of reducing the dissolved O2 to ˙O2− radicals. However, with the addition of rGO, the electronic interaction and charge equilibrium between Bi2WO6 and rGO lead to the negative shift of the Fermi level and a change in the energy levels of the conduction band and valence band, indicating that the photocatalytic performance of reduction reactions is enhanced.62,63
![]() | ||
Fig. 11 Effects of different scavengers on (A) RhB and (B) 4-CP degradation over Bi/Bi2WO6/rGO-3 under visible light irradiation. |
In addition to photocatalytic efficiency, the stability of photocatalysts is another important issue for their practical application. To study the photocatalytic stability and reusability of Bi/Bi2WO6/rGO-3, the photocatalysis process was repeated five times under the same conditions. As shown in Fig. 12A, the Bi/Bi2WO6/rGO-3 composite had a high and stable activity for photocatalytic degradation of RhB under visible light irradiation. After five cycling runs, the photocatalyst incurs no detectable loss of its photocatalytic activity. XRD patterns of Bi/Bi2WO6/rGO-3 before and after the photocatalytic reaction were comparable, and are shown in Fig. 12B. There was no apparent change observed in the XRD patterns. TEM and XPS analysis also proved that there was no apparent change in composition and morphology for Bi/Bi2WO6/rGO-3 during the photocatalytic process (Fig. S6 and S7†), which also confirmed the stability of the structure. It can be concluded that Bi/Bi2WO6/rGO possesses an efficient photocatalytic activity and can be easily separated for reuse.
The enhanced photocatalytic activities of the Bi/Bi2WO6/rGO composites could be ascribed to the synergic effects of the following factors. Firstly, the introduction of Bi nanorods in the composite enhanced the light absorption in the visible region, which could be ascribed to the SPR effect of metallic Bi. Secondly, due to their matching energy levels, the interfacial charge transfer could effectively enhance the separation efficiency of photogenerated electrons and holes. Finally, the presence of rGO provides more active adsorption sites on the surface of rGO in the photocatalytic reaction. A proposed mechanism for the separation and transportation of photogenerated electron–hole pairs at the Bi/Bi2WO6/rGO interface is shown in Scheme 2. Electrons and holes could be generated from both photo-excited Bi2WO6 and SPR-excited Bi under visible light irradiation. Since the Fermi energy level of Bi (Ef = −0.17 eV vs. SHE) is more negative than the conduction band (CB) of Bi2WO6 (ECB = 0.24 eV vs. SHE),64,65 it is energetically favorable for the SPR-excited electrons to transfer from Bi to the CB of Bi2WO6. Subsequently, these electrons can transfer from the CB of Bi2WO6 to the surface of rGO nanosheets due to their two-dimensional conjugated π-structure and ultrahigh charge carrier mobility. Concomitantly, the holes transfer to the Bi2WO6 and Bi surfaces, which corresponds to oxidation of dye molecules into CO2. Thus, in the case of Bi/Bi2WO6/rGO, the rGO sheets served as an acceptor of the photogenerated electrons of Bi2WO6 and Bi, which suppressed the recombination of photogenerated electron–hole pairs. The electrons in rGO can be captured by the dissolved O2 molecules to generate superoxide anion radicals (˙O2−) and then react with RhB (or 4-CP). In the meantime, the holes in the composite can directly oxidize the pollutants into CO2 and H2O, and other intermediates.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra20316g |
This journal is © The Royal Society of Chemistry 2016 |