S. Kaviya and
Edamana Prasad*
Department of Chemistry, Indian Institute of Technology Madras, Chennai-600 036, TN, India. E-mail: pre@iitm.ac.in; Fax: +91-44-2257-4202; Tel: +91-44-2257-4232
First published on 29th January 2015
Photocatalytic degradation of pollutants will be attractive if the degradation process is under direct sunlight in the presence of a cost effective catalyst. We have synthesized a novel, monodispersed ZnO–Ag nano custard apple (NCA) using pomegranate peel extract as reducing and stabilizing agent. The ZnO–Ag NCA was characterized by various spectroscopic and microscopic methods. The photocatalytic performance of ZnO–Ag NCA was evaluated using the degradation of methylene blue dye in the presence of direct sunlight. It was observed that NCA show ∼40% and 24% enhanced photocatalytic activity compared to commercial ZnO nanoparticles and TiO2 (P25). Moreover, the performance of the photocatalyst NCA has been examined in the presence of oxidative reagents such as peroxomonosulfate (PMS), peroxodisulfate (PDS) and hydrogen peroxide (H2O2). The mineralization data of methylene blue, performed by total organic carbon (TOC) analysis, revealed that 95% and 70% mineralization was achieved in 3 h using a NCA–PMS combination and NCA, respectively. The catalyst did not lose its catalytic activity for up to five cycles. Overall, this system is relatively inexpensive, reproducible, extremely stable and efficient for complete degradation of methylene blue in aqueous solution.
Reports have shown that the photocatalytic efficiency of ZnO nanocomposites can be significantly affected by morphologies.14,15 Various synthetic strategies have been developed to modify the photocatalytic performance of ZnO–Ag photo catalyst by different morphologies such as spherical,16 core–shell,17 rod,18 plate,19 wire20 and flower.21 In this article, we report a novel morphology of ZnO–Ag catalyst, achieved via one-pot, inexpensive, template free, rapid biosynthesis. The shape of the nano-cluster is identical with that of a custard apple and hence we name the monodispersed ZnO–Ag catalyst as a Nano Custard Apple (NCA). To the best of our knowledge, this is the first successful synthesis of monodispersed ZnO–Ag nano custard apple (NCA). While researchers have been using various synthetic methods to prepare ZnO nanoparticles (NPs) such as flame reactor,7 wet-phase,7 laser ablation,22 hydrothermal,23 seed layer,24 sol–gel,25 sputtering,26 chemical vapour deposition (CVD),27 molecular beam epitaxy (MBE),28 we have adopted a greener approach in this work where pomegranate peal extract is used as reducing and stabilizing agent for the nanosystem. Biological approach29–31 using plant or plant fractions for synthesizing nano-structures are attractive over other synthetic methodologies7,22–28 due to their simple, cost effective and eco-friendly nature. The NCA has been utilized for photocatalytic degradation of methylene blue (MB) under sunlight irradiation. The effect of the photocatalytic efficiency of ZnO–Ag NCA in presence of other oxidants such as peroxomonosulfate (PMS), peroxodisulfate (PDS) and hydrogen peroxide (H2O2) has been examined. We have also investigated the complete mineralization of MB and reusability of the catalyst.
![]() | ||
Fig. 1 XRD pattern of ZnO–Ag nanoparticles with different loading of Ag (0.2–2.0 wt%) calcined at 300 °C. |
The size and morphology of ZnO–Ag NPs are studied by SEM and TEM analysis. The SEM images of NPs formed at different loading of Ag (wt%) are shown in Fig. 2 and S1.† The morphology of ZnO–Ag NP formed in presence of 0.2 wt% of Ag indicates the tendency of the system to form custard apple kind of structures (Fig. S1a†). However, the particles formed at 0.6 wt% loading of Ag have clearly shown custard-apple morphology with the diameter of about ∼100 nm and the assemblies are monodispersed in nature (Fig. 2a and b). TEM image of ZnO–Ag (0.6 wt%) NPs also show similar morphology (Fig. 2c). The high resolution TEM image displayed the distance of two adjacent lattice fringe spacing in ZnO and AgNPs and the values were about 0.24 and 0.23 nm, respectively (Fig. 2d and S2†). It also confirms the well dispersed nature of AgNP on ZnO surface in ZnO–Ag NCA. The lattice fringe distance corresponds to (101) and (111) planes in the ZnO and Ag crystal, respectively.35 As the concentration of Ag is increased from 1.0 to 2 wt%, the NCA structure is collapsed and irregular morphologies were obtained (Fig. S1b–d†). This suggests that excess amount of silver leads to the aggregation of particles. While it is known that calcination temperature (300 °C) has some influence on the shape of the particles, the above results strongly suggest that silver play a major contribution towards the size, shape and morphology of ZnO–Ag NP formation. The presence of Ag on ZnO–Ag nano clustered apple was further confirmed by EDAX analysis (Fig. S3†). The elemental mapping obtained from EDS analysis shows that Zn, O and Ag are distributed in the ZnO–Ag NCA (Fig. S4†).
DRS peaks of silver loaded (0.2 to 2.0 wt%) ZnO are shown in Fig. 3. The pure commercial ZnO NPs exhibits absorption only in UV region indicating its incapability of light harvesting at visible region. The spectra of ZnO–Ag NPs contain both ZnO peak around 350 nm and the surface plasmon resonance peak of silver NPs around 430 nm. This also confirms the formation of AgNPs in the reaction mixture during the synthesis. While increasing the concentration of Ag from 0.2 wt% to 0.6 wt%, the peak intensity of ZnO is decreased and enhancement in the SPR peak (420 nm) of AgNPs has been observed. The peak at 420 nm indicates the formation of well dispersed spherical AgNPs in NCA. The SPR peak position was relatively stable between 1.4 wt% and 2.0 wt% of Ag, indicating that the surface saturation of ZnO NPs by AgNPs. The SPR peak is broadened (400–750 nm) as the concentration of Ag is further increased (1 wt% to 2 wt%). This indicates the increment in the particle size of AgNPs on ZnO, due to higher order of surface deposition of AgNPs.29 The band gap of ZnO–Ag NCA (3.0 eV) calculated by Tauc's plot and the value is significantly lower than pure ZnO (Fig. S5†). The result clearly indicates that the absorption properties of ZnO were altered by the formation of ZnO–Ag NCA.
Photocatalytic property of the NPs highly depends on the PL intensity and recombination rate of the excited electrons and holes.36 Fig. S6† shows the room temperature photo luminescent (PL) spectra (excitation at 325 nm) recorded from the bare ZnO and ZnO–Ag nano particles. Each spectrum consists of two bands, one is in the UV region (383 nm) and another is in a visible region (550 nm).37 The peak at 383 nm is attributed to the near-band edge (NBE) emission of ZnO NP, which is mainly due to the recombination of electron–hole in ZnO.37 A broad green emission band centred at 550 nm is due to recombination of photogenerated hole and the singly ionized-charge oxygen vacancy.38 The reduced intensity at NBE of ZnO–Ag (0.6 wt%) NCA indicates strong exchange interaction of electrons between Ag and ZnO.37 Conversely, the PL intensity at 550 nm is decreased as the concentration of silver is increased to 0.6 wt% of Ag, presumably because of dispersion of AgNPs on ZnO reaches a maximum at the particular Ag concentration. The decreased emission intensity at NBE indicates a lower electron–hole recombination rate and hence a longer life time of the photogenerated carries caused by AgNPs.36 The results taken together revealed that, charge transfer from ZnO to Ag is rapid in the case of ZnO–Ag NCA, at 0.6 wt% of Ag.
To investigate the molecular vibration modes in ZnO–Ag nano custard apple, FT-Raman measurement was carried out at room temperature by exciting the sample with a laser source λ = 1064 nm. Fig. S7† shows the Raman spectrum of ZnO–Ag nano custard apple. A pure hexagonal wurtzite phase of ZnO in NCA shows Raman peaks at 583.1, 435.3, 382.0, 331.6 and 99 cm−1 which are corresponding to the E1 (LO-longitudinal optical), E2 (high), A1 (TO-transverse optical), A1 and E2 (low) vibrational modes.39,40 ZnO–Ag nano custard apple shows dominant Raman peaks at 436 and 99 cm−1 which correspond to E2 (high) and E2 (low), respectively.41 The second order Raman peak at 333 cm−1 was originated from the zone boundary phonons.41 The less intense peak at 580 cm−1 is associated with the oxygen vacancies and surface defects caused by the addition of Ag on ZnO nanostructure. This is also consistent with the PL spectra at 550 nm (Fig. S6†). Further, the presence of Ag on ZnO–Ag nano custard apple is confirmed by observing the peaks at 1036, 1102, 1348 and 1583 cm−1 which are assigned to Ag nanoparticles.42,43
The surface chemical composition and chemical states of elements present in ZnO–Ag NCA was investigated by XPS measurement. Fig. S8† shows the XPS spectra of ZnO–Ag NCA which indicates the presence of Zn, O, and Ag. The strong peak at 1021.8 eV is ascribed to Zn 2p3/2 and peak at 1044.2 eV for Zn 2p½, indicating the oxidation state of Zn as +2.44 The binding energy of O1s peak (528.5 to 530.14 eV) was attributed to lattice oxygen present in ZnO, whereas, the peak at 531.57 eV was assigned to surface hydroxyl group in ZnO–Ag NCA. The presence of surface hydroxyl group facilitates the trapping of photoinduced electron and hole, leading to enhanced photocatalytic degradation process.44,45 Ag 3d spectrum of AgNP in ZnO–Ag NCA contains two peaks at 368.3 and 374.1 eV which are assigned to Ag 3d5/2 and Ag 3d3/2, respectively.45
![]() | ||
Fig. 4 Photocatalytic degradation of methylene blue in the presence of different photocatalysts ([catalyst] = 0.5 g L−1, [MB] = 0.02 mM). |
The dye degradation rate is highly depended on the morphology and crystallinity of the sample.47 The enhanced photocatalytic activity of biosynthesized ZnO–Ag NPs is due to the formation of nano heterojunctions compared to ZnO NPs and TiO2 (P25).48 We observed a better catalytic activity for ZnO–Ag NCA, presumably due to the uniform morphology and optimum loading of Ag, which results in the strong electronic interaction between Ag and ZnO NPs which retards the recombination of charge carriers (Fig. S6†). Since the rate of charge recombination is less, an electric field at the ZnO–Ag nano-junction will be built in, upon sunlight irradiation.49 The presence of AgNPs in the catalyst acts as nano-antenna for light trapping due to the characteristic localized surface plasmon resonance (LSPR) of AgNPs which is converted into locally excited electric field as a result of photo excitation of LSPR.48 While shining sunlight into ZnO–Ag NCA the electron in the valance band of ZnO will transfer into the conduction band. ZnO in ZnO–Ag NCA will allow the charge transfer from ZnO to AgNPs due to LSPR and accelerates the separation of photogenerated electron–hole pairs. The holes in the valence band will react with water and generate ˙OH radicals and electron in the conduction band of AgNPs will produce superoxide anion radicals (O2˙−) during the photocatalytic degradation process. The formation of these active radicals results in the efficient photocatalytic degradation of MB under direct sunlight.
Alternately, sunlight can excite the adsorbed methylene blue on the surface of the catalyst, which injects electrons into the conduction band of ZnO nano particles.50 The electrons on the ZnO surface can be trapped by the AgNPs and thereby prevent the recombination process. Further, the electrons are consumed by the dissolved oxygen in the solution to generate reactive oxidative species to degrade the dye molecules.45,51 The overall mechanistic picture of the photocatalytic degradation of MB in presence of ZnO–Ag NCA under sunlight irradiation is shown schematically in Fig. 5. The various reactions involved in the process can be summarized as follows.45,52,53
![]() | ||
Fig. 5 Proposed mechanism for the electron transfer events occurred in ZnO–Ag NCA during the degradation of MB upon sunlight irradiation: (a) band gap excitation and (b) sensitization. |
The amount of catalyst required for the photocatalytic reaction is important since it can strongly influence the photocatalytic degradation of MB. As we found ZnO–Ag NCA shows a better photocatalytic activity towards the degradation of MB, we optimized the amount of catalyst required for a better degradation rate of MB. We varied the catalyst amount from 0.1 to 0.6 g L−1 and found that the degradation rate of MB was increased up to 0.5 g L−1 of ZnO–Ag NCA (Fig. 6). The higher amount of the catalyst (0.5 g L−1) offers more active surface area for the photocatalytic degradation of MB. A decrease in the degradation rate of ZnO–Ag NCA at 0.6 g L−1 is presumably due to emergence of light scattering by excess amount of catalyst present in the reaction medium.50
![]() | ||
Fig. 6 Photocatalytic degradation of methylene blue at various concentration of ZnO–Ag NCA under sunlight irradiation ([catalyst] = 0.5 g L−1, [MB] = 0.02 mM). |
The generation of ˙OH during the photocatalytic reaction in the presence of catalyst was monitored using terephthalic acid (TA) as a probe molecule. The fluorescent intensity of TA is directly proportional to the amount of ˙OH production during the photocatalytic reaction due to the formation of hydroxy terephthalic acid.49 Fig. 7a shows the fluorescent intensity of hydroxy terephthalic acid formed in presence of different catalyst during the photocatalytic reaction under sunlight irradiation for 60 min. The results revealed that there is no formation of ˙OH in the dark reaction condition in presence of ZnO–Ag NCA. The fluorescent intensity is enhanced with time in presence of catalyst upon sunlight irradiation. We observed that ZnO–Ag NCA generates significantly higher amount of ˙OH than other catalyst. We have monitored the PL intensity of terephthalic acid in presence of ZnO–Ag NCA with different irradiation time (Fig. S10†). The maximum intensity was obtained at 60 min of sunlight irradiation. The results indicates that plasmonic effect of well dispersed and smaller sized AgNPs in ZnO–Ag NCA could lead efficient trapping of higher number of photogenerated charge carriers and generate more number of ˙OH during the reaction.
We investigated the effects of oxidant such as peroxomonosulfate (PMS), peroxodisulfate (PDS) and hydrogen peroxide (H2O2) on the photocatalytic degradation of MB in presence of ZnO–Ag NCA under sunlight irradiation. We observed that the photocatalytic degradation of MB by ZnO–Ag NCA is higher in presence of PMS than PDS or H2O2 (Fig. 7b). An efficient charge separation at the interface of ZnO/Ag in ZnO–Ag NCA, leads higher surface tapped electrons. PMS oxidise faster and simultaneously produce ˙OH and SO4˙− during the reaction.54 Hence the overall rate of degradation of MB was increased in ZnO–Ag NCA with PMS combination.
The total organic carbon (TOC) analysis was performed to evaluate the mineralization efficiency of ZnO–Ag NCA in presence and absence of other oxidants (Fig. 8a). Generally, complete mineralization of MB takes place in two steps. The initial step involves the ring opening of MB, which is followed by subsequent oxidation of the fragments. Hence, mineralization will take more time than decolorization of dye molecule in presence of photo catalyst. The results revealed that, in the absence of oxidants, the mineralization rate of ZnO–Ag NCA is significantly low. The higher rate of mineralization was observed in the case of ZnO–Ag NCA with PMS. The complete mineralization of MB in presence of ZnO–Ag NCA with PMS under sunlight irradiation was taken place in 3 h.
Finally, the reusability of biogenic ZnO–Ag NCA during the photo catalytic degradation of MB was investigated (Fig. 8b). The results indicate that, even five consecutive cycles the NCA did not lose the photocatalytic activity towards the degradation of MB. It is noteworthy that the NCA based photo catalyst was stable under sunlight irradiation. The catalytic activity is decreased after the use of fifth cycle. The reason for this observation is the leaching of AgNPs from the surface of ZnO which can be seen from the XRD pattern (Fig. S11a†). Corroborating to this, the corresponding SEM images (Fig. S11b†) of the sample showed the collapsed structure of ZnO–Ag NCA.
Footnote |
† Electronic supplementary information (ESI) available: See DOI: 10.1039/c4ra15293j |
This journal is © The Royal Society of Chemistry 2015 |