Wee-Jun Onga,
Jia-Jun Yeonga,
Lling-Lling Tana,
Boon Tong Gohb,
Siek-Ting Yonga and
Siang-Piao Chai*a
aMultidisciplinary Platform of Advanced Engineering, Chemical Engineering Discipline, School of Engineering, Monash University, Jalan Lagoon Selatan, Bandar Sunway, 47500 Selangor, Malaysia. E-mail: chai.siang.piao@monash.edu; Fax: +60-3-551-46207; Tel: +60-3-551-46234
bLow Dimensional Materials Research Centre, Department of Physics, Faculty of Science, University of Malaya, 50603 Kuala Lumpur, Malaysia
First published on 3rd November 2014
In this study, a series of reduced graphene oxide (RGO)–Zn0.5Cd0.5S nanocomposites was synthesized via an improved one-step co-precipitation-hydrothermal strategy using thiourea as an organic S source. The experimental results demonstrated that thiourea facilitated heterogeneous nucleation of Zn0.5Cd0.5S and in situ growth of Zn0.5Cd0.5S nanocrystals on the RGO sheets via electrostatic attraction. Moreover, the addition of NaOH as a precipitating agent in the reaction environment was found to reduce the aggregation of Zn0.5Cd0.5S on the RGO sheets. Such an intimate interfacial contact between Zn0.5Cd0.5S and RGO resulted in well-dispersed nanoparticles decorated on RGO sheets. Photocatalytic performances of the RGO–Zn0.5Cd0.5S were evaluated by the degradation of Reactive Black 5 (RB5) under a low-power 15 W energy-saving daylight bulb at ambient conditions. Compared with pristine Zn0.5Cd0.5S, 20RGO–Zn0.5Cd0.5S (20 wt% of RGO) displayed an enhanced RB5 degradation of 97.4% with a rate constant of 0.0553 min−1 after 60 min of visible light irradiation. 20RGO–Zn0.5Cd0.5S exemplified a 1.3-fold enhancement after RGO incorporation relative to that for pristine Zn0.5Cd0.5S. The remarkable photocatalytic performance was ascribed to the efficient migration efficiency of the photoinduced electrons from Zn0.5Cd0.5S to RGO to inhibit the charge carrier recombination. Additionally, to systematically verify the role of each active species in the degradation of RB5, trapping experiments for radicals and holes were individually explored. It is confirmed that photogenerated ˙O2−, ˙OH and h+ were responsible for the degradation of RB5 in the 20RGO–Zn0.5Cd0.5S system. Lastly, a postulated visible-light photocatalytic mechanism for the RB5 degradation was discussed.
Dealing with pollutants demands the development of effective and sustainable decontamination technologies. Advanced oxidation process (AOP), particularly heterogeneous photocatalysis, has been shown to be a promising solution to the remediation of environmental pollution due to the fact that it can be conducted at ambient conditions resulting in desired complete mineralization of dye degradation.2 Heterogeneous photocatalysis for environmental decontamination is considered to be a green strategy without creating harmful by-products. In the past few years, a number of remarkable progresses on semiconductor photocatalysis have been achieved in the environmental degradation.3–7 TiO2 and ZnO have been widely used to degrade various organic pollutants in water and air due to their high stability, resistant to corrosion and relatively inexpensive.8–10 Although the sun provides an abundant source of photons, they are effective only under UV light with moderate performance due to their wide band gap energies, which comprises a small fraction (5%) of solar energy that reaches the Earth's surface as compared with visible light (45%).11,12 In addition to wide band gap, high recombination rate of charge carriers imposes severe limitations on the photocatalytic efficiency. Therefore, considering energy saving and utilization, engineering of highly efficient visible-light-induced photocatalyst is an urgent and challenging matter hitherto.
Metal chalcogenides are emerged as good candidates for visible-light photocatalysts due to their suitable band positions and improved catalytic functions. Among them, CdS has received immersed attention due to the narrow band gap resulting in more responsive to the visible light region.13 However, it suffers from high charge recombination rate and photocorrosion under visible light irradiation, inhibiting the application of CdS to a large extent.14 On the other hand, the band gap of ZnS is ca. 3.66 eV, which exhibits remarkable ability in redox and durability, but fails in utilizing the visible light effectively.15 To overcome the aforementioned problems, engineering ZnxCd1−xS photocatalyst is a viable approach because CdS possesses similar coordination mode with ZnS and metal atoms could be mutually substituted in the same crystal lattice.16,17 The fabrication of ZnxCd1−xS could circumvent the weak oxidation ability and photocorrosion of CdS as well as shift the absorption edge of ZnS into visible region. As a result, the band gap energy of ZnxCd1−xS can be flexibly tuned by changing the molar ratio of Zn and Cd precursors.
To further enhance the photocatalytic performance of ZnxCd1−xS, carbonaceous nanomaterials are employed as the supports of catalysts for the application in the photocatalytic fields.13,18 In particular, graphene, which is known as a single atomic layer of graphite arranged in six-membered rings of carbon atoms, offers a two-dimensional (2D) superior electrical conductivity, making it an excellent electron transport matrix.19 This favors the effective transfer of the photogenerated electrons from the surface of photocatalyst to graphene to prolong the lifetime of charge carriers. Recently, there have been several studies on the development of reduced graphene oxide (RGO)–ZnxCd1−xS via different synthetic routes as effective visible-light photocatalysts for H2 production from water splitting. For instance, Yan et al.20 developed RGO–ZnxCd1−xS by a facile one-pot co-precipitation reaction using H2S as a S source and reducing agent. Besides that, Zhang et al.18 synthesized RGO–Zn0.8Cd0.2S nanocomposites by a one-pot co-precipitation-hydrothermal approach using Na2S for enhanced H2 production compared to pure Zn0.8Cd0.2S. Nevertheless, the influence of different S sources on the synthesis of RGO–ZnxCd1−xS has recently been explored by Li et al.21 It was found that Na2S- or H2S-based reaction promoted rapid formation of ZnxCd1−xS, which resulted in large grain size due to quick aggregation, and thus poor contact between RGO and ZnxCd1−xS. Very recently, Min et al.15 prepared RGO–ZnxCd1−xS by a two-step co-precipitation-hydrothermal method using thiourea as a precursor. However, it was found that the Zn0.5Cd0.5S nanospheres were decorated on the RGO sheets with a large particle size of ca. 100–200 nm. It is known that a large particle size results in an inefficient harvesting of light and higher electron–hole recombination rate due to increased migration distances en route to the surface reaction sites.22
In addition, to effectively suppress the recombination of photogenerated charge carriers, the excellent interfacial contact between RGO and ZnxCd1−xS is of paramount significance. Hence, inspired by previous works, we envision for the first time that, an improved one-pot co-precipitation-hydrothermal technique with thiourea as the organic S source in the synthesis of RGO–Zn0.5Cd0.5S was employed for the photodecolorization of dye, in which limited attention has paid to in the past literature. Besides the low toxicity, the diamine structure of thiourea could reduce graphene oxide (GO) to RGO.21 Our results demonstrated that among pure ZnxCd1−xS with different stoichiometries, Zn0.5Cd0.5S exhibited the greatest degradation of Reactive Black 5 (RB5) under a low-power 15 W energy-saving daylight bulb. Thus, Zn0.5Cd0.5S was used in the following study. The relative quantity of RGO to Zn0.5Cd0.5S can prominently affect the visible-light-responsive property and photocatalytic activity. Thus, it is indispensable to optimize the relative mass for the heterogeneous growth of Zn0.5Cd0.5S on the RGO sheets for effective visible light response and attaining the maximum photocatalytic efficiency. Also, mechanism of the enhanced activity over RGO–Zn0.5Cd0.5S is still unclear till now. As a result, the role of active species for the degradation of RB5 was thoroughly investigated due to the lack in evidence on confirming the main oxidative and reductive species (holes, electrons or radicals), which have been relatively scarce and still in its infancy in the RGO–Zn0.5Cd0.5S system.
The durability and stability tests of the as-synthesized RGO–Zn0.5Cd0.5S nanocomposites were also conducted using similar procedure. The product underwent three consecutive cycles with each lasting for 1 h. In order to make up for the loss of catalyst during the recycling process, two first runs were conducted separately and its remaining catalysts were mixed to obtain 30 mg of photocatalysts before conducting the 2nd then 3rd stability tests. After each cycle, the photocatalyst was centrifuged (13500 rpm, 30 min) and washed thoroughly with DI water, and then it was added to fresh RB5 solution. A minimum of two duplicate runs were conducted under the same conditions and the results were averaged to ensure the reproducibility of photodegradation results. The experimental error was found to be in the range of ±5%.
Moreover, the addition of NaOH as a precipitating agent in the reaction medium was found to further reduce the nucleation rate by inhibiting the tendency of quick growth and aggregation of Zn0.5Cd0.5S. This could be evidenced by a drastic decrease in the mean particle size of Zn0.5Cd0.5S from 303 nm (without NaOH in the reaction medium) to 31.6 nm (with NaOH in the reaction medium) in the 20RGO–Zn0.5Cd0.5S hybrid system as shown in Fig. S3 in ESI.† Therefore, a small grain size was attained due to reduced surface energy of Zn0.5Cd0.5S nanoparticles via a one-pot co-precipitation-hydrothermal process. The in situ generation of Zn0.5Cd0.5S nanoparticles on the RGO sheets improved the interfacial close contact between the two phases, which agreed well with the TEM images, as discussed in the latter section. Meanwhile, GO was simultaneously reduced to RGO since ethanol is a good reducing agent under hydrothermal process.26 Overall, the RGO sheets provided a fertile platform for the heterogeneous nucleation of Zn0.5Cd0.5S to take place. This is in accordance with previous studies on the role of RGO with tunable surface characteristics as a desirable scaffold for the growth of guest materials.27–29 The whole synthetic protocol of the RGO–Zn0.5Cd0.5S hybrid nanocomposites is depicted in Fig. 1.
![]() | ||
Fig. 1 Schematic illustration of the synthesis process of RGO–Zn0.5Cd0.5S nanocomposites (for clarity, sizes are not represented proportionally.). |
The XRD patterns of the as-prepared graphite oxide, pure Zn0.5Cd0.5S and xRGO–Zn0.5Cd0.5S are illustrated in Fig. 4. For graphite oxide, the XRD peak at 9.5° corresponded to the (001) interlayer spacing of 0.93 nm, which indicated the presence of oxygen-containing functional groups.31,32 Interestingly, no apparent peak of graphite oxide at 2θ = 9.5° was observed in all the xRGO–Zn0.5Cd0.5S hybrid nanocomposites. This could be attributed to the exfoliation of graphite oxide in the nanocomposites and reduction of GO to RGO with significantly fewer oxygen-containing functionalities after the one-pot synthesis process, which agreed well with the previous studies.33,34 In addition, no characteristic diffraction peak of graphene (2θ = 24.5°) was observed due to the low amount and relatively low diffraction intensity of graphene.35 The reduction of GO to RGO could be further evidenced by FTIR analysis, as discussed later. Meanwhile, all diffraction peaks of xRGO–Zn0.5Cd0.5S could be well-indexed to the hexagonal wurtzite of the pristine Zn0.5Cd0.5S, with the distinct peaks corresponding to (100), (002), (101), (110), (103) and (112) planes.16,36 No peak shift was shown when RGO was hybridized with Zn0.5Cd0.5S, manifesting that RGO had a negligible effect on the crystal phase of Zn0.5Cd0.5S during the formation of the composites. By comparing 20RGO–Zn0.5Cd0.5S with 20RGO–ZnS and 20RGO–CdS, all these composites demonstrated an apparent shift in the XRD patterns (Fig. S6 in ESI†). When Zn was incorporated into the CdS crystal, the diffraction peaks of 20RGO–Zn0.5Cd0.5S presented an obvious shift towards the higher angle. This implied that Zn2+ incorporated into the lattice of CdS crystal and decreased the fringe lattice distance of the CdS due to smaller radius of Zn2+ ion (0.74 Å) than that of Cd2+ (0.97 Å).37,38 This deduced that the Zn atoms in the CdS crystal influenced the position of Cd atoms and subsequently varied the lattice structure of CdS during the fabrication of ZnxCd1−xS. The calculated lattice constants a and c of 20RGO–ZnS, 20RGO–Zn0.5Cd0.5S and 20RGO–CdS are summarized in Table S1 in ESI.† The changes in lattice constants for these nanocomposites proved that a solid solution of Zn0.5Cd0.5S was formed instead of physical coupling between ZnS and CdS crystals.
Since the oxygen-containing functional groups are active in the IR region, FTIR analysis was employed to qualitatively investigate the degree of deoxygenating in the xRGO–Zn0.5Cd0.5S hybrid nanocomposites. As indicated in Fig. 5, apart from the CC skeletal vibration of sp2 unoxidized graphitic domains at 1625 cm−1, the FTIR spectrum of graphite oxide divulged the presence of oxygenated functional groups at 853 (O–C
O stretching), 1054 (C–O–C stretching), 1204 (phenolic C–OH stretching), 1728 (C
O stretching vibrations in carboxyl or carbonyl groups) and 3415 cm−1 (structural O–H groups on the RGO).23 The peak at 2360 cm−1 was clearly recorded for all the samples, which was ascribed to the physically adsorbed CO2 from the atmosphere. Compared with graphite oxide, for all the xRGO–Zn0.5Cd0.5S hybrid nanocomposites, the vibration of O–C
O, C–O–C and C
O peaks disappeared, whereas the intensity of the C–OH and O–H peaks considerably reduced after the one-pot co-precipitation-hydrothermal route, signifying that most of the oxygen-containing functionalities in GO have been effectively reduced.
UV-Vis diffuse reflectance spectra (DRS) of the as-prepared photocatalysts are presented in Fig. 6A. For all the xRGO–Zn0.5Cd0.5S samples, a substantial increase in the absorption at wavelengths shorter than 530 nm was attributed to the intrinsic band gap of Zn0.5Cd0.5S. According to the Kubelka–Munk (KM) function, the band gap energy of pristine Zn0.5Cd0.5S was estimated to be ca. 2.53 eV by extrapolating the linear region of the [F(R)hv]2 vs. photon energy (Fig. 6B). Furthermore, when a small amount of RGO was incorporated with Zn0.5Cd0.5S, a background absorption in the visible light region (550–800 nm) was enhanced with increasing RGO content for the xRGO–Zn0.5Cd0.5S. Evidently, the corresponding colour of the samples became darker, which was from yellow to dark green, with increasing RGO content (inset of Fig. 6A). This decreased the reflection of light and hence improved the light absorption ranging from 500 to 800 nm. The surface electric charge of the Zn0.5Cd0.5S was enhanced due to RGO hybridization, resulting in the electronic transition between excited Zn0.5Cd0.5S and RGO. Further observation implied that all the xRGO–Zn0.5Cd0.5S nanocomposites exhibited almost similar absorption edge as that of the pure Zn0.5Cd0.5S, claiming that the layered-structured RGO in the xRGO–Zn0.5Cd0.5S system only served as a scaffold for the immobilization of Zn0.5Cd0.5S nanocrystals.17 Hence, the estimated band gap energies of the xRGO–Zn0.5Cd0.5S nanocomposites were found to be very similar (ca. 2.50–2.53 eV) (Fig. 6B). However, too high RGO content in the nanocomposite would impose a negative effect on the photocatalytic performance because RGO will absorb most of the light, resulting in a reduced light utilization by Zn0.5Cd0.5S. Thus, an appropriate RGO loading in the xRGO–Zn0.5Cd0.5S samples is of paramount significance for maximizing light utilization and photocatalytic activity.
![]() | ||
Fig. 6 (A) UV-Vis DRS (inset shows the colour of the photocatalysts) and (B) plot of transformed KM function vs. hv for the Zn0.5Cd0.5S and xRGO–Zn0.5Cd0.5S (x = 1, 5, 10, 15, 20, 25 and 30). |
With the hybridization of RGO to the Zn0.5Cd0.5S, the photocatalytic activity of RGO–Zn0.5Cd0.5S was markedly enhanced, elucidating that the addition of an appropriate amount of RGO could noticeably improve the overall photocatalytic efficiency. Among all the RGO–Zn0.5Cd0.5S photocatalysts, 20RGO–Zn0.5Cd0.5S presented the highest photodegradation efficiency of 97.4% relative to the pure Zn0.5Cd0.5S (92.0%). This was accredited to the unique features of RGO in improving the lifetime of photoinduced electron–hole pairs from Zn0.5Cd0.5S and thus boosting the separation of charge carriers.39 This is due to the fact that RGO possesses a remarkably high charge carrier mobility of more than 200000 cm2 V−1 s−1 at room temperature and endows a superior 2D π-conjugation structure to efficiently retard the charge recombination.40–42 This statement was well-supported by PL analysis in which a strong fluorescence emission peak at ca. 530 nm was detected, which was attributed to the intrinsic band gap energy of Zn0.5Cd0.5S (Fig. 7C). In comparison to pure Zn0.5Cd0.5S, the remarkably decreased PL emission intensity of the 20RGO–Zn0.5Cd0.5S photocatalyst represented a lower recombination probability of photogenerated charge carriers with more efficient interfacial charge transfer due to the interaction between excited Zn0.5Cd0.5S and RGO sheets. Thus, the RGO sheet in the RGO–Zn0.5Cd0.5S nanocomposite acted as the separation center of the charge carriers due to its highly conductive electron transport “highway”. More importantly, the incorporation of RGO enhanced the adsorptivity of RGO–Zn0.5Cd0.5S nanocomposites via π–π conjugation between RB5 and aromatic regions of RGO.43 Analogues features were also observed in the case of RGO–TiO2 and RGO–ZnO for the degradation of dye.44,45
In addition, the absorption intensity of the RGO–Zn0.5Cd0.5S enhanced with increasing RGO content because RGO is known to be a good light-harvesting material as supported by UV-Vis analysis. However, with the excess addition of RGO content, the photocatalytic activity was deteriorated, revealing that a synergistic interaction between RGO and Zn0.5Cd0.5S is necessary to improve the photoactivity. Evidently, when the RGO content in the RGO–Zn0.5Cd0.5S was beyond 20 wt%, a decrease in photocatalytic activity was observed for 25RGO–Zn0.5Cd0.5S and 30RGO–Zn0.5Cd0.5S nanocomposites. This was due to the presence of excess RGO, resulting in an increase in the opacity and light scattering. In other words, the introduction of excess RGO into the Zn0.5Cd0.5S matrix led to shielding of the active sites on the catalyst surface and decreased the light intensity through the depth of the RB5 solution. Despite the increase of absorption intensity in the visible region, the total amount of separated electrons from Zn0.5Cd0.5S reduced although the addition of RGO was excellent for electron storage. Hence, this concludes that high addition amount of RGO is not favourable for the enhanced photoactivity, which is generally a universal problem for the RGO-based composites.46
The time-dependent absorption spectra of RB5 solution in the presence of 20RGO–Zn0.5Cd0.5S was shown in Fig. 7D. The intensity of the peak at 597 nm, which was the characteristic UV-Vis absorption of RB5 molecule, reduced significantly with time and diminished after 60 min of light irradiation, confirming the complete decomposition of RB5 molecule. This could be further evidenced by the colour change from blue to colourless after 60 min of light irradiation as shown in the inset of Fig. 7D. From Fig. 7E, there was a linear relationship between ln(Co/C) and irradiation time, confirming that the photocatalytic degradation was indeed pseudo first order kinetics based on the Langmuir–Hinshelwood model: ln(Co/C) = kt, where Co is the equilibrium concentration of RB5, C is the concentration at time t and k is the apparent first order reaction rate constant.25,47 The calculated k value for 20RGO–Zn0.5Cd0.5S was 0.0553 min−1, which exhibited a 1.3-fold enhancement after RGO hybridization relative to that for pristine Zn0.5Cd0.5S. The order of k value was found to be 0.0553 min−1 (20RGO–Zn0.5Cd0.5S) > 0.0487 min−1 (15RGO–Zn0.5Cd0.5S) > 0.0454 min−1 (10RGO–Zn0.5Cd0.5S) > 0.0448 min−1 (25RGO–Zn0.5Cd0.5S) > 0.0444 min−1 (5RGO–Zn0.5Cd0.5S) > 0.0439 min−1 (1RGO–Zn0.5Cd0.5S) > 0.0433 min−1 (Zn0.5Cd0.5S) > 0.0396 min−1 (30RGO–Zn0.5Cd0.5S) > 0.0002 min−1 (GO), which was concordant with the results shown in Fig. 7A and B.
The formation of ˙OH could be further confirmed by the PL technique using TA as a probe molecule. The experiment of measuring the active ˙OH could avoid the self-photosensitization effect of dye. As depicted in Fig. 8B, a gradual increase in PL intensity at 425 nm was observed with increasing irradiation time, manifesting that the ˙OH was formed in the photocatalytic degradation process, which was in agreement with the results of tert-butanol and isopropanol quenching. In the absence of light or 20RGO–Zn0.5Cd0.5S photocatalyst, the PL intensity did not increase. This result signified that the fluorescence was excited by the 2-hydroxyterephthalic acid from the chemical reactions between TA and ˙OH formed on the illuminated 20RGO–Zn0.5Cd0.5S surface. Therefore, ˙OH was originated from both pathways, namely (1) from the reaction of photogenerated electrons through multistep reduction of O2 and (2) from the oxidation of water via holes. Furthermore, Fig. 8C illustrates a comparison of the PL intensity for pure Zn0.5Cd0.5S and 20RGO–Zn0.5Cd0.5S. It is clear that the amount of ˙OH produced from the 20RGO–Zn0.5Cd0.5S was higher than that of pure Zn0.5Cd0.5S, inferring that the former exhibited higher photocatalytic activity than the latter owing to the inhibition of charge carrier recombination by the RGO sheets, which verified well with the photocatalytic results and PL emission spectra in Fig. 7A–C. Overall, these results confirmed the evidence of ˙OH formation and participated in the photodegradation process.
Based on the above information, a synergetic effect photocatalytic mechanism of 20RGO–Zn0.5Cd0.5S on the enhancement of visible light activity was demonstrated in Fig. 8D. It is evident that the interaction between Zn0.5Cd0.5S and RGO was responsible for the efficient charge separation under visible light irradiation. Firstly, upon light irradiation, Zn0.5Cd0.5S with a band gap energy of 2.53 eV was excited by visible light and induced the formation of electron–hole pairs. Secondly, the photoexcited electrons were transferred from conduction band (CB) of Zn0.5Cd0.5S (−0.42 eV vs. NHE) to RGO (−0.08 eV vs. NHE) via a percolation mechanism,25,49 which has a lower band-edge position to achieve charge equilibrium and stabilization between these two components. It has been widely reported that the delocalized π structure of RGO facilitates the transfer of electrons and serves as a superior electron acceptor.34,50 As a result, the transfer of electrons to RGO significantly reduced the electron trapping in the lattice of Zn0.5Cd0.5S to suppress the recombination of electron–hole pairs. The accumulated electrons on the RGO sheets reacted with molecular O2 in solution to form reactive ˙O2− followed by the formation of oxidative ˙OH species (eqn (2) and (3)). The RB5 molecules were adsorbed with offset face-to-face orientation via π–π conjugation between RB5 and the aromatic regions of RGO,43 leading to mineralization of RB5 to other degraded products. On the other hand, the holes in the valence band (VB) of Zn0.5Cd0.5S (2.11 eV vs. NHE) have strong oxidizing power to oxidize RB5 directly as evidenced by the results of triethanolamine quenching. At the same time, holes could oxidize OH− or H2O to ˙OH since the VB position of Zn0.5Cd0.5S was more positive than the standard redox potential of ˙OH/OH− (1.99 eV vs. NHE).51,52 The adsorbed RB5 was finally oxidized by the ˙OH radicals to produce mineralized products, which was in line with our results of tert-butanol and isopropanol quenching. The plausible photocatalytic reaction process was shown in eqn (1)–(6).
![]() | (1) |
RGO (e−) + O2 → RGO + ˙O2− | (2) |
˙O2− + 2H+ + 2e− → ˙OH + OH− | (3) |
h+ + H2O → H+ + ˙OH | (4) |
h+ + OH− → ˙OH | (5) |
RB5 + h+(˙OH, ˙O2−) → mineralized products | (6) |
Footnote |
† Electronic supplementary information (ESI) available: Synthesis procedures of GO, photocatalytic degradation of RB5 using pristine ZnxCd1−xS, FESEM images, particle size analysis and XRD patterns of the studied photocatalysts as control experiments, XRD patterns of pristine ZnxCd1−xS, adsorption–desorption equilibrium of RB5 on the photocatalysts in the dark and electronegativity calculation method. See DOI: 10.1039/c4ra10467f |
This journal is © The Royal Society of Chemistry 2014 |