Zhigang Xiong and
X. S. Zhao*
School of Chemical Engineering, The University of Queensland, St. Lucia, Brisbane, QLD 4072, Australia. E-mail: george.zhao@uq.edu.au; Fax: +61-733654199; Tel: +61-733469997
First published on 10th November 2014
Composite materials consisting of layered titanate and cadmium sulfide (CdS) particles were prepared via consecutive ion exchange and precipitation processes. Samples with different CdS loadings were characterized using X-ray diffraction, physical adsorption/desorption of nitrogen, scanning/transmission electron microscopy and ultraviolet-visible diffuse reflectance spectroscopy techniques. The photocatalytic properties of the samples in degrading methylene blue (MB) under visible light irradiation were evaluated. Results showed that CdS nanoparticles (NPs) were intercalated between titanate layers. The composite materials showed an improved photocatalytic activity as compared to both bulk CdS and a mixture of CdS and titanate. The intimate contact between titanate and CdS particles benefited interparticle electronic coupling. The observed improvement in photocatalytic activity is a combination of a number of contributions, including CdS photocatalysis, photosensitized degradation of MB over K2Ti4O9, and the synergistic effect between CdS and K2Ti4O9. The synergy indeed contributed to MB degradation and was found to depend on the content of CdS in the composite with an optimized content of 19 wt%, but it is not a dominant factor determining the overall photocatalytic activity of the composite.
To solve the problems associated with CdS photocatalyst, embedding CdS in a solid matrix has been demonstrated to be a promising approach.28–31 Chen et al.30 showed that CdS-implanted layered double hydroxide (LDH) displayed an excellent performance in the degradation of rhodamine B under both UV and visible light irradiations. Shangguan and co-workers31 reported that CdS-intercalated metal oxides were superior to both CdS and a mixture of CdS and the metal oxide with respect to photocatalytic activity for hydrogen evolution. We also observed32 an enhanced photocatalytic performance of a CdS–graphene composite in the degradation of non-biodegradable dyes under visible light irradiation. The improvement is generally attributed to the fast transfer of photogenerated electrons from CdS to the second material, such as metal oxide or graphene.1 Such a synergistic effect between CdS and the second material has been interpreted by comparing their band potentials.12 However, due to the complicated nature of photocatalytic reactions,33 they require an understanding of the role of each component in the composite including CdS, the second semiconducting material, the CdS/support interface and the interaction of these with chemical reactants during photocatalytic processes. To what extent does the synergistic effect between the two components determine the overall photocatalytic performance also needs further analysis? The study could have significant impact on understanding the underlying mechanism and designing advanced composite photocatalytic materials.
Here, we report the preparation of CdS–pillared titanate composite materials and their photocatalytic properties in degradation of methylene blue (MB) under visible light irradiation. Titanate was selected to couple with CdS because of the following reasons. First, the layered structure of titanate34 could stabilize CdS particles. Second, titanate can be prepared as large particles,35 making it easy to separate from the reaction mixture by for example sedimentation or filtration. Third, as has been demonstrated previously,11 titanate possesses an excellent ability of accepting photogenerated electrons, thus having a potential to minimize the electron–hole recombination.36 In the present work, titanate–CdS composites (designated as K2Ti4O9–CdS) with different loadings of CdS were prepared and characterized. The role of each semiconductor component in the photocatalytic degradation of MB and the effect of CdS content on the synergistic contribution were investigated.
![]() | ||
Fig. 1 XRD patterns of different samples. The standard XRD pattern of CdS is included for reference. The data on the top left are the basal spacing in nanometres. |
In the preparation of CdS–titanate composite materials, Cd2+ ions were firstly loaded onto K2Ti4O9 via an ion exchange process, followed by precipitation in the presence of S2− as illustrated in Scheme 1. Briefly, 50 mL of K2Ti4O9 suspension (20 g L−1) was mixed with 20 mL of Cd(NO3)2 (Sigma-Aldrich) solutions with different concentrations (0.24 M for K2Ti4O9–CdS-1 and 0.81 M for K2Ti4O9–CdS-2) under stirring. The mixture was stirred at 60 °C for 24 h. Then, a Na2S solution (0.49 M, Sigma-Aldrich) was added dropwise. After continuous stirring at 60 °C for 3 h, yellow precipitates were collected by filtration, washed several times with deionized water, and dried in an oven at 70 °C. The samples are denoted as K2Ti4O9–CdSx%, where x stands for the weight percentage of CdS in a composite. The content of CdS was determined by the energy dispersive X-ray (EDS) analysis.
For comparison purpose, a mixture containing K2Ti4O9 and 19 wt% of CdS was prepared. This sample is denoted as (K2Ti4O9–CdS)mix.
The stability of the photocatalyst was evaluated as follows. 0.05 g of K2Ti4O9–CdS47% was suspended in 100 mL of 2.0 × 10−5 M MB solution. After a given time interval, 5 mL aliquots were withdrawn, filtered and analyzed on the UV-Vis spectrophotometer. After MB was completely degraded, the suspension was further irradiated for 20 min. Then, 20 mL stock solution (1 × 10−4 M) of MB was supplied to restore the initial concentration of MB to 2.0 × 10−5 M. The suspension was equilibrated again for 30 min in the dark before being irradiated by visible light.
Fig. 2 shows the FESEM and HRTEM images of sample K2Ti4O9 and K2Ti4O9–CdS-1. As can be seen, titanate K2Ti4O9 has a rodlike morphology with a diameter of approximately 500 nm and lengths up to several micrometres (Fig. 2A). The surface of sample K2Ti4O–CdS-1 was covered by small aggregates with sizes ranging from 10 nm to 500 nm (Fig. 2B). The EDX data (Fig. 2C and D) confirmed the presence of CdS in the composite. The HRTEM images (Fig. 2E–G) further confirmed the rodlike morphology of titanate K2Ti4O9, as well as the presence of CdS aggregates deposited on the surface of the rod. The dark regions on the K2Ti4O9 background were probably due to the intercalated CdS particles. The lattice fringe indicates the highly crystalline nature of CdS. The interplanar spacing of the adjacent lattice fringes was estimated to be 0.34 nm, attributed to the (111) plane of CdS.
![]() | ||
Fig. 2 FESEM images of K2Ti4O9 (A) and K2Ti4O9–CdS-1 (B). SEM image (C), EDX spectrum (D), and HRTEM images (E–G) of K2Ti4O9–CdS-1. |
Table 1 shows the elemental contents in two composite samples. It can be seen that when the initial [Cd2+] increased from 0.24 M to 0.81 M, the atomic ratio of K/Ti decreased from 0.15 to 0.07 with the ratio of Cd/Ti contradictorily increased from 0.39 to 1.27, indicating the displacement of K+ by Cd2+ during the ion exchange process. The result is slightly different from that determined by XPS analysis (Fig. S2†), in which element K was unable to be detected, and the atomic percentage of Cd was 12.0% in composite K2Ti4O9–CdS-1, much larger than 5.4% determined by EDX. As XPS is an effective tool for surface analysis, K+ ions on the external surface of K2Ti4O9 is thus believed to be fully displaced by Cd2+ ions, with some Cd ions intercalating into the interlayers as evidenced by the XRD data (Fig. 1). When the content of CdS increased from sample K2Ti4O–CdS-1 to sample K2Ti4O–CdS-2, the atomic percentage of Cd on the surface increased about 1.6 times, while the overall Cd content determined by EDX increased about 2.3 times, suggesting that most Cd ions were intercalated into the interlayers of K2Ti4O9, resulting in an expansion of basal spacing as confirmed by the XRD data (Fig. 1). It is also noted that the atomic concentration of Cd determined by XPS is much larger than that of S, but they are comparable in the EDX analysis. The results suggest that there still have some Cd2+ ions adsorbed on the surface of K2Ti4O9 that cannot be precipitated by S2− ions. The resulting weight percentages of CdS in composites K2Ti4O9–CdS are thus calculated by using that of the element S (Table 1), and are determined to be 19 wt% and 47 wt% for sample K2Ti4O9–CdS-1 and K2Ti4O9–CdS-2, respectively.
For (K2Ti4O9–CdS)mix, the FESEM image (see Fig. S3†) indicates that most CdS particles existed as aggregates in the mixture instead of depositing on the surface of K2Ti4O9, which is in good accordance with the XRD analysis (Fig. 1).
The nitrogen adsorption–desorption isotherms and the pore size distribution curves of samples K2Ti4O9, K2Ti4O9–CdS19% and K2Ti4O9–CdS47% are shown in Fig. S4.† It is seen that the K2Ti4O9–CdS composites displayed a type IV isotherm, with a H3-type hysteresis loop, indicating the aggregation of plate-like particles giving rise to slit-shaped pores.40 The Brunauer–Emmett–Teller (BET) specific surface area of samples K2Ti4O9–CdS19% and K2Ti4O9–CdS47% were calculated to be about 33.8 and 77.0 m2 g−1, respectively. The BET surface area of sample K2Ti4O9 was only 2.2 m2 g−1. Similar findings have also been demonstrated in iron oxide41 and zinc oxide42 pillared titanate. The pore size distribution curves computed from the desorption branches using the Barrett–Joyner–Halenda (BJH) model clearly showed a narrow distribution with the main peak centred at about 3.5 nm and a weak shoulder peak at around 1.8 nm for both composite samples. On the basis of the data of both the pore size distribution curves and the basal spacing of the composite materials, the formation of large pores was due to the stacking of titanate platelets and the small pores to the basal spacing of layered K2Ti4O9 upon intercalation of CdS particles. The porous structure of the composite materials is believed to favour the uptake of organic pollutants from solution to the catalyst surface for subsequent reactions.
Fig. 3 shows the UV-visible reflectance spectra of samples K2Ti4O9, CdS, K2Ti4O9–CdS, and (K2Ti4O9–CdS)mix. Layered titanate K2Ti4O9 as a typical photocatalyst has a strong absorption in the UV region with the absorption edge at around 351 nm.11,35,43 Composite K2Ti4O9–CdS however showed a strong absorption in the visible light region with the absorption edge at around 539 nm for K2Ti4O9–CdS19% and 579 nm for K2Ti4O9–CdS47%. The bandgap thus can be estimated to be about 2.30 eV and 2.14 eV, respectively. Compared to that of bulk CdS (∼2.05 eV), the CdS NPs in K2Ti4O9–CdS displayed a larger bandgap.31 As the bandgap strongly depends on the crystallite dimension of a semiconductor,4 the bandgap increases and the band edge shifts to yield larger redox potentials when the crystallite dimension decreases. The result suggests that the crystallite size of CdS particles in composite K2Ti4O9–CdS are smaller than that of bulk CdS,4,44 probably due to the confining effect of the K2Ti4O9 interlayers. Moreover, the absorption intensity of composite K2Ti4O9–CdS19% was much lower than that of (K2Ti4O9–CdS)mix, even though both samples had similar content of CdS, attributing to the intercalation of CdS into K2Ti4O9 interlayers and the low extinction coefficient of K2Ti4O9 in the visible region (Fig. 3). Nonetheless, the loading of CdS enabled the composite K2Ti4O9–CdS to be active under visible light irradiation, whereas the intimate contact between CdS and K2Ti4O9 and the increased redox potential of CdS will benefit the interparticle electronic coupling and charge transfer.
![]() | ||
Fig. 3 UV-visible reflectance spectra of (a) K2Ti4O9, (b) CdS, (c) (K2Ti4O9–CdS)mix, (d) K2Ti4O9–CdS19%, and K2Ti4O9–CdS47% (e). |
The photocatalytic performance of the samples in the degradation of methylene blue (MB) under visible light irradiation is shown in Fig. 4. It can be seen that MB was very stable in the absence of a photocatalyst, or in the presence of photocatalysts without visible light irradiation. The concentration of MB was however slightly decreased in the presence of K2Ti4O9 or (K2Ti4O9–CdS)mix in the dark. This concentration decrease is probably due to ion exchange between cationic MB dye and K+ in K2Ti4O9.45 As K2Ti4O9 is only photocatalytically active under UV light irradiation, MB underwent moderate degradation in the K2Ti4O9–MB system under visible light irradiation. With regard to photocatalyst CdS, it displayed a significantly higher degradation rate than that of K2Ti4O9. Composite K2Ti4O9–CdS47% (or K2Ti4O9–CdS-2) gives the highest degree of MB degradation under the visible light irradiation. The performance of K2Ti4O9–CdS47% is even higher than that of P25, which can degrade dye pollutants under visible light via a photosensitization mechanism.11 The absorption of MB in the visible light region gradually decreased with irradiation time (Fig. S5†), indicating the complete destruction of aromatic structures. The degradation was well-fitted with a pseudo-first-order kinetics model ln(Ct/C0) = −kt, where Ct and C0 are concentrations of MB at time t and 0 min, respectively, and k is the rate constant, which were calculated to be 1.75 × 10−2, 1.47 × 10−2 and 3.48 × 10−3 min−1 for samples K2Ti4O9–CdS47%, CdS and K2Ti4O9, respectively. The k gradually decreased when the initial concentration of MB increased (Fig. S6†). It is also seen that the degradation rate of MB in the presence of composite K2Ti4O9–CdS19% was slightly slower than that of K2Ti4O9–CdS47%, but faster than that of (K2Ti4O9–CdS)mix with the rate constants of about 1.29 × 10−2 min−1 for K2Ti4O9–CdS19% sample and 8.24 × 10−3 min−1 for (K2Ti4O9–CdS)mix sample. To further identify the visible light activity of K2Ti4O9–CdS, benzoic acid that has no absorption in the visible light region was used as a probe molecule, which underwent gradual degradation under visible light irradiation (Fig. S7†), but it was hardly degraded by P25 or K2Ti4O9 under the same experimental conditions.
Fig. S8† shows the stability of photocatalyst K2Ti4O9–CdS47%. It can be seen that the material possessed an excellent stability after four times of reuse. The slightly declined photocatalytic activity is believed to be due to the loss of catalyst during sampling for the analysis of MB concentration.8
Considering that both CdS and titanate K2Ti4O9 are semiconductor photocatalysts,28,35 and only CdS can be excited by visible light irradiation. After bandgap excitation, the electron on the VB of CdS was injected into its CB (Fig. 5), followed by reacting with dissolved oxygen to form ROS such as HO˙, O2˙−, HOO˙ and H2O2.4,12 The valence hole together with the ROS attacked MB present in both solution and on the catalyst surface, producing various N-demethylated products, such as thionine and formaldehyde,46 which were further mineralized to CO2 and H2O under prolonged light irradiation.4 For photocatalyst K2Ti4O9, it cannot be excited by visible light due to its wide bandgap energy, which can only absorb UV light (Fig. 3). However, as titanate is semiconductive,35,47 its CB can accept electrons generated from excited dye molecules as has been observed previously.45 The electron in the CB reacts with surface-adsorbed O2 to produce ROS for dye degradation. Such a photosensitization-assisted degradation over K2Ti4O9 process is different from semiconductor photocatalysis,8,11,45,48 and has proven to be an effective route for the removal of non-biodegradable dye pollutants.48
The possibility of electron transfer from excited MB to K2Ti4O9 can be estimated using the standard redox potentials versus normal hydrogen electrode (NHE) (ENHE), which can be estimated from the work functions (E) using ENHE (V) = −4.5 − E (eV).13 Considering that the redox potentials of MB and excited MB (MB*) versus NHE are 1.17 V and −0.69 V,49 and the potential energies of the CB and VB of K2Ti4O9 are −4.02 and −7.56 eV,11 respectively, the electron transfer from excited MB to K2Ti4O9 is thus energetically feasible (see Fig. 5), accounting for the observed moderate degradation of MB over K2Ti4O9 (Fig. 4). Such a degradation process has also been observed in other semiconducting systems including TiO2,48 graphene,13,50 SnO2,51 etc. However, considering the redox potentials of CB (−3.78 eV) and VB (−5.84 eV) of CdS,32 the degradation of MB over semiconductor CdS via the photosensitization process is unlikely due to the mismatched redox potentials (Fig. 5). Takizawa and co-workers46 compared the degradation of MB and rhodamine B (RhB) in the presence of CdS photocatalyst, and also concluded that the degradation of MB over CdS was merely due to the photocatalysis of CdS rather than the photosensitized electron transfer from excited MB to the CB of CdS.
For composite K2Ti4O9–CdS, the MB degradation is much more complicated. The observed overall performance consists of at least three processes: CdS photocatalysis, the photosensitized degradation of MB over K2Ti4O9, and the synergistic contribution of the two semiconductors. Coupling K2Ti4O9 with CdS could alleviate the charge carrier recombination in CdS and realize a vectorial transfer of photogenerated charge carriers from one component to the other due to their unique structural properties.12 First, K2Ti4O9 is a semiconductor with an anisotropic structure,11,43 indicating a low surface and bulk irregularities and facilitating a high transport mobility of charge carriers in the bulk phase.52 Second, the redox potential of the CB of K2Ti4O9 is much higher than that of CdS (Fig. 5), thus thermodynamically favouring direct electron transfer from CdS to K2Ti4O9. We have previously observed that titanate can indeed accept electrons in a titanate–anatase core–shell composite through an interparticle charge carrier transition mechanism.11,36 The higher potential gap between CdS and titanate K2Ti4O9 (Fig. 5) than that between anatase and K2Ti4O911 could impose a stronger driving force, enabling the charge transfer from the CB of CdS to that of K2Ti4O9. The photoluminescence spectra of CdS, K2Ti4O9–CdS19% and K2Ti4O9–CdS47% shown in Fig. S9† confirmed the charge transfer process. It is seen that CdS has an intense fluorescence emission maximum at 516 nm due to the electron–hole recombination. After being combined with K2Ti4O9, the matched band potentials (Fig. 5) drives the charge transfer from one particle to its neighbour to form a spatial separation between electrons and holes, leading to decreased luminescent intensity of CdS in composite K2Ti4O9–CdS. This retardation improved the utilization of excitons and enhanced the production of ROS for MB degradation (Fig. 5).
It is noted that the observed concentration decrease of MB in the presence of composite K2Ti4O9–CdS may also be partially due to ion exchange between cationic MB and K+ in K2Ti4O9. This contribution was however insignificant (Fig. 4), because most of the ion exchange sites of titanate had already been occupied by Cd2+ during the preparation process. While the removal of MB by ion exchange between MB and Cd2+ can be neglected as is evidenced by the minor change of the MB concentration in the presence of K2Ti4O9–CdS19% in the dark (Fig. 4).
Fig. 6 shows the rate constants for MB degradation over photocatalysts with different CdS loadings. It is seen that coupling CdS with semiconductor K2Ti4O9 indeed enhanced the degradation efficiency with the optimum amount of CdS at about 47 wt%. From Fig. 6 and Table 2 it is then possible to normalize the rate constant with respect to CdS loaded onto K2Ti4O9. We will focus on the interparticle interaction between CdS and K2Ti4O9 support as the contribution of synergistic effect to the overall degradation requires further analysis. It is clear from the last column of Table 2 that the normalized rate constant of sample K2Ti4O9–CdS19% is larger than that of sample K2Ti4O9–CdS47%, suggesting the exertion of larger photocatalytic activity of CdS NPs in sample K2Ti4O9–CdS19%. This may be due to the fact that a low loading of CdS NPs at K2Ti4O9 surface benefits the dispersion of CdS (Fig. 2B) and the contact between the two semiconductors, thus enhancing the interparticle charge separation and depressing the charge recombination in excited CdS NPs. On contrast, a high loading of CdS may result in severe aggregation of CdS particles (Fig. S10†), weakening the interparticle interaction and the charge separation between excited CdS and K2Ti4O9. In addition, the high photoactivity of sample K2Ti4O9–CdS47% as compared to K2Ti4O9–CdS19% suggests that the synergistic effect between coupled semiconductors with matched band potentials is not always a dominant factor determining the overall photocatalytic performance. Other factors, such as CdS photocatalysis and the photosensitization-assisted degradation over K2Ti4O9 as observed in this work should also be considered when designing novel composite photocatalysts for different applications.
Samples | EDX CdS/Tia | k (min−1) | Normalized kb |
---|---|---|---|
a Atomic ratio.b Rate constant normalized with respect to CdS content is obtained by dividing the rate constant k (column 3) by the CdS/Ti ratio (column 2). | |||
CdS | — | 1.47 × 10−2 | — |
K2Ti4O9 | — | 2.55 × 10−3 | — |
K2Ti4O9–CdS | |||
19 wt% | 0.52 | 1.29 × 10−2 | 2.48 × 10−2 |
47 wt% | 2.22 | 1.75 × 10−2 | 1.12 × 10−2 |
Footnote |
† Electronic supplementary information (ESI) available: FESEM images of sample K2Ti4O9–CdS47% and (K2Ti4O9–CdS)mix, N2 sorption–desorption isotherms and pore size distribution curves, photoluminescence spectra of CdS and composite K2Ti4O9–CdS, and the recycled degradation of MB. See DOI: 10.1039/c4ra09692d |
This journal is © The Royal Society of Chemistry 2014 |