Kula Kamal
Senapati
,
Chandan
Borgohain
and
Prodeep
Phukan
*
Department of Chemistry, Gauhati University, Guwahati-781014, Assam, India. Fax: (+)91-361-2700311; E-mail: pphukan@yahoo.com
First published on 20th June 2012
A new synthesis of a CoFe2O4–ZnS magnetic nanocomposite has been achieved and used as an efficient photocatalyst in the degradation of methyl orange in water under UV irradiation. The photocatalyst is magnetically recoverable from the reaction medium and suitable for multiple uses with sustained catalytic activity.
Semiconductor nanocatalysts having high surface area, strong UV absorption and long life-span have been widely used for the degradation of organic pollutants in water.3–5 Among these, ZnS (bandgap energy: 3.6–3.8 eV) has served as one of the most efficient photocatalysts for wastewater treatment such as in the photodegradation of dyes, p-nitrophenol, and halogenated benzene derivatives, reduction of heavy metals and water-splitting for H2 evolution.6–11 Recently, Suslick developed a nanostructured Ni2+-doped ZnS photocatalyst for the evolution of H2 under visible light.12 ZnS nanoparticles (NPs) are a good photocatalyst due to the rapid generation of electron–hole (e−/h+) pairs by photoexcitation and high negative reduction potentials of excited electrons. However, they have the serious limitation of easy recombination of e−/h+ pairs which in turn reduce the photocatalytic activity significantly.13,14 Moreover, these types of suspended photocatalysts suffer a noticeable disadvantage in ease of separation, recovery, recycling and high cost in large scale production.15 In addition, they tend to agglomerate in aqueous solution and hinder the penetration of lights.
Consequently, design and development of magnetic NCs with high catalytic activity and easy separation can heighten their potential for practical applications. Magnetic materials such as Fe3O4,16,17 BaFe2O4,18,19 NiFe2O420,21 and γ-Fe2O322 have recently been utilized as core materials in making magnetic NCs with TiO2 or ZnO semiconductor shells, which can be easily separated from a solution by applying a magnetic field.
It is also noteworthy that the fabrication of a magnetic-nanocomposite with a uniform coating over the core surface in nano-scale dimensions is a challenging task due to difficulties in complete dispersion of the magnetic core in solution. Moreover, the commonly used core materials such as Fe3O4 or γ-Fe2O3 are not chemically or thermally stable.23 Therefore, a new approach for preparing a stable and highly dispersible magnetic core is desirable for effective coating with catalytically active species in nano-scale dimensions.
Recently, we have reported a new magnetic nanocomposite using ferromagnetic CoFe2O4 as the core material with multi-functional properties.24–27 CoFe2O4 is well-known to have large magnetic anisotropy, moderate saturation magnetization, remarkable chemical stability and mechanical hardness,28 which makes it a good candidate as a core material for the fabrication of NCs. Zhang reported a new CoFe2O4-ZnO nanocomposite with multiple properties.29 Very recently, Mourão reported the synthesis of a CoFe2O4-TiO2 nanocomposite for photodegradation of dye in water.30 The photocatalytic activity of these NCs might be related to the electronic and absorption properties of the core materials which could restrain the recombination of e−/h+ pairs by absorbing photo-generated electrons from the conduction band of the semiconductor. Thus, the avenue of CoFe2O4 as a core material can be explored for the development of a conjugate of CoFe2O4 and ZnS NPs. However, no study on photocatalytic activity of CoFe2O4–ZnS NCs has been reported so far. Therefore, a study on such conjugates might pave the way for developing recyclable photocatalysts for the degradation of organic contaminants in water and environmental remediation.
![]() | (1) |
The Powder XRD measurement was done individually for the ZnS, CoFe2O4 and CoFe2O4–ZnS NPs to characterize the phase and crystallization (Fig. 1a–c). It can be seen that cubic ZnS NPs (JCPDS card 5-0566)33 have been prepared and the broadening of diffraction peaks is in accordance with the nanocrystalline nature of the ZnS particles. The diffraction pattern of the as-synthesized CoFe2O4–ZnS particles (Fig. 1a) shows that all the peaks are in good agreement with the cubic structure of spinel CoFe2O4 (JCPDS cards: 3-864 and 22-1086)34,35 and of ZnS NPs. These results reveal that the nanocomposite is composed of cubic-structured CoFe2O4 and ZnS NPs. The average size of the crystalline particles for the primary CoFe2O4 NPs and ZnS NPs was estimated to be 30 ± 2 nm and 7 ± 1 nm, respectively, based on the Debye–Scherrer formula.36
![]() | ||
Fig. 1 XRD pattern of (a) CoFe2O4–ZnS (*contribution from ZnS NPs), (b) CoFe2O4 and (c) ZnS NPs. |
Fig. 2 shows the representative TEM image, with the SAED pattern in the inset, of the as-prepared ZnS NPs showing the uniform size-distribution with the size in between 5–10 nm, which is in good agreement with the size determined from the Debye–Scherrer relation.36 The diffraction rings have been identified as the (111), (220) and (311) planes of the cubic ZnS phase.33 The morphology of the CoFe2O4–ZnS NCs was analyzed by FESEM (Fig. S1 in ESI†) and TEM measurement as shown in Fig. 2 (c, d) which show the uniform size distribution of NPs with cubic morphology. The corresponding SAED pattern of the composite is shown in the inset of Fig. 2(e). The crystalline sizes were found to be about 30 ± 5 nm, which were close to the average crystallite size obtained from the Debye–Scherrer formula based on XRD patterns. A close observation of the FESEM micrographs indicates that the primary CoFe2O4 nanocrystals are covered with ZnS NPs. The TEM image of CoFe2O4–ZnS NCs shows the core-shell nanostructure, where the dark particles of CoFe2O4 are shielded by a thin layer of bright ZnS particles (Fig. 2c–d).
![]() | ||
Fig. 2 TEM image of (a) ZnS NPs and (c, d) CoFe2O4–ZnS NCs with the corresponding SAED patterns (b, e). |
The lattice image of the nanocomposite (Fig. S3–S5 of ESI†) marked with the (111) lattice plane of spinel CoFe2O4 and the (220) lattice plane of cubic ZnS NPs revealed the highly crystalline nature of the NCs. The high resolution TEM (HRTEM) image allowed us, by means of Gatan Digital MicrographTM software, to obtain the fast Fourier transform (FFT) from which lattice spacings of 4.59 A° and 2.7 A° were obtained (experimental error, <5%) for the CoFe2O4 and ZnS NPs, respectively.
The as-synthesized NPs were further characterized by laser micro Raman spectroscopy which is a useful tool for investigating NPs and their surrounding media.37 Strong evidence for the core–shell structure of CoFe2O4–ZnS NCs was also observed in the Raman spectra. The inset in Fig. 3 shows the Raman spectra of ZnS NPs with the characteristic peaks at around 259 and 345 cm−1 which indicates the cubic structure of ZnS NPs.38 The spinel CoFe2O4 shows five Raman active modes (Ag + Eg + 3E2g).39 Three modes of CoFe2O4 NPs at 458 cm−1, 600 cm−1 and 665 cm−1 can be seen in Fig. 3(b) along with the 258 cm−1 and 345 cm−1 modes from ZnS NPs. We did not observe any significant shifts in bands of Raman spectra, which indicates that there is no substitution in the CoFe2O4–ZnS nanocomposite.40 Therefore, it may be inferred that the core–shell structured CoFe2O4–ZnS NC has been effectively synthesized.
![]() | ||
Fig. 3 Raman spectra of (a) CoFe2O4 NPs and (b) CoFe2O4–ZnS NCs and the inset is of ZnS NPs (*Raman scattering peaks from ZnS NPs). |
We further evaluated the magnetic properties of the CoFe2O4 core and the CoFe2O4–ZnS NCs by VSM measurement at room temperature as shown in Fig. 4. The NCs give the coercivity (Hc) and saturation magnetization (Ms) of 1.2 kOe and 17.5 emu g−1 respectively which is only a negligible change from the CoFe2O4 core (Hc: 1.6 kOe and Ms: 18 emu g−1) and that is an indication of the formation of nano-scale coating at the core-surface. The band-gap energy, Eg of the as-synthesized nanocomposite was determined based on the Tauc plot41 and was found to be 3.2 eV from the absorbance spectra (Fig. S6 in ESI†).
![]() | ||
Fig. 4 Magnetic hysteresis loops for (a) CoFe2O4 NPs and (b) CoFe2O4–ZnS NCs. |
The specific surface area and porosity were measured by the application of the BET equation and the BJH method. Fig. 5 shows the N2 adsorption–desorption isotherms of the CoFe2O4–ZnS NC. The BET surface area of the NCs was found to be 30.269 m2 g−1 as calculated by the linear part of the BET plot, which is much higher than that of the CoFe2O4 NPs prepared by the co-precipitation method (17.97 m2 g−1).42 The total pore volume at P/Po = 0.98 is 0.18 cm3 g−1. The BET isotherm is of type II and H3 hysteresis loop (BDDT/IUPAC classification), characteristic of mesoporous adsorbents.43
![]() | ||
Fig. 5 BET isotherm and pore size distribution (inset) of the as-synthesized CoFe2O4–ZnS magnetic NPs. |
To investigate the potential application of the magnetic NC as a recyclable photocatalyst for the removal of contaminants from wastewater, we performed the degradation of MO in an aqueous medium driven by UV-irradiation as a probe reaction. Fig. 6 (I) and (II) show the absorption spectra of an aqueous solution of MO irradiated under UV-light for different time intervals in the presence of CoFe2O4–ZnS NCs and ZnS NPs, respectively. The typical absorption peak at 460 nm gradually disappears as the UV-irradiation time increases, and completely diminishes after 20 min, meaning complete photodegradation of MO by the CoFe2O4–ZnS NCs. The photocatalytic degradations of MO over different time intervals using the as-synthesized samples as the catalyst and with no catalyst are shown in Fig. 6 (III). We have observed that the CoFe2O4–ZnS NCs exhibit high photocatalytic efficiency with a degradation of 98% in 20 min of UV light exposure. The photocatalytic degradation of MO was further investigated by LCMS analysis as described in the supplementary information (Fig. S10–13 in ESI†) which revealed its appreciable degradation under the photocatalytic conditions applied. Thus, a NC with enhanced photocatalytic activity has been achieved that is capable of faster degradation of MO than any other reported method.44–46
![]() | ||
Fig. 6 Changes of UV-Vis spectra of MO with irradiation time in the presence of (I) CoFe2O4–ZnS NCs and (II) ZnS NPs and (III) degradation curves of MO with (a) CoFe2O4–ZnS NCs, (b) no catalyst, (c) ZnS NPs, (d) CoFe2O4 NPs and (e) TiO2 NPs under UV light irradiation. |
For comparison, we also checked the photocatalytic activity of the as-synthesized CoFe2O4 core and ZnS NPs shell employing entirely the same photocatalytic conditions. While keeping all the reaction conditions the same, the ZnS NPs showed low photocatalytic activity and no apparent photocatalytic activity was also detected for CoFe2O4 NPs under the identical UV-irradiation. We also tested the photocatalytic activity of TiO2 NPs (TEM and XRD are shown in Fig. S7 and S8, respectively, in ESI†) under the same reaction conditions and observed very low degradation of MO in 20 min of UV light exposure. The good photocatalytic activity of the CoFe2O4–ZnS NCs with a larger rate (rate constant: 0.0718 min−1) in spite of much larger particle sizes as compared to ZnS NPs, which have considerably low photocatalytic activity (rate constant: 0.0041 min−1), may be associated with small crystallites and poor crystallinity of small NPs. In addition, the high photocatalytic efficiency of the NCs as compared to ZnS NPs is due to the electrostatic interaction between the core and shell components of the conjugate nanostructure of CoFe2O4 and ZnS NPs.
The corresponding mechanism can be interpreted as follows: under UV light irradiation, electrons in the valence band (VB) of ZnS are excited to its conduction band (CB) leaving the same amount of holes in the VB forming the e−/h+ pairs. Due to the strong interfacial connection between ZnS and CoFe2O4, the excited electrons of the CB of ZnS can transfer to that of CoFe2O4. So, the e−/h+ pairs are separated at the CoFe2O4–ZnS interfaces, which decreased their recombination probability. Meanwhile, the O2 absorbed on the CoFe2O4 surface can accept electrons and enhance the efficiency of generating hydroxyl radicals, which are very reactive towards the degradation of organic molecules and so enhanced photocatalytic activity is observed. A plausible mechanistic scheme of the charge separation and the photocatalytic activity for the photocatalyst is shown in Scheme 1.
![]() | ||
Scheme 1 Illustration of photocatalysis by CoFe2O4–ZnS NCs. |
The photocatalytic decomposition of MO by CoFe2O4–ZnS NCs follow a pseudo first-order kinetic law, and can be expressed as
![]() | (2) |
![]() | ||
Fig. 7 Kinetics of MO degradation in the presence of ZnS NPs and CoFe2O4–ZnS NCs. |
It can be seen that there exists a linear relationship between ln(Co/C) and irradiation time. The pseudo-first-order rate constant and linear regression coefficients (R) for photo-degradation of MO with CoFe2O4–ZnS NCs and ZnS NPs are mentioned in the graph. It has been observed that the photo-degradation rate by CoFe2O4–ZnS NCs (k = 0.0718 min−1) is much faster than that by ZnS NPs (k = 0.0041 min−1).
To extend the scope of the nanocomposite for photodegradation of other dyes, we checked the degradation of methylene blue and phenolphthalein using the CoFe2O4–ZnS photocatalyst under the same reaction conditions (Fig. S16 and S17 in ESI†). The degradation of methylene blue was comparable to that of MO. However, in case of phenolphthalein, the degradation was lower (49% in 60 min) under the identical reaction conditions. This difference in degradation might be due to the difference in redox potentials of different dyes along with their difference in adsorption capacities.
We further observed that the magnetic photocatalyst could be easily separated in the liquid-phase reaction with the aid of an external magnet. To ensure complete removal of the catalyst we measured the energy-dispersive X-ray spectrum of the water residue after centrifugation of the treated water at very high speed and no elemental contribution present in the nanocomposite was detected. The recyclability of the photocatalyst was tested upon separation of the photocatalyst magnetically and then performance of photocatalysis for multiple cycles (Fig. 8). We have not observed any significant loss of activity up to five catalytic cycles, which indicates that the as prepared magnetic photocatalyst is stable and very effective for the removal of organic contaminants from water.
![]() | ||
Fig. 8 Recyclability of photocatalytic degradation of MO. |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c2cy20400b |
This journal is © The Royal Society of Chemistry 2012 |