DOI:
10.1039/C5RA04608D
(Paper)
RSC Adv., 2015,
5, 55499-55512
Photocatalytic degradation of 4-chlorophenol over Ag/MFe2O4 (M = Co, Zn, Cu, and Ni) prepared by a modified chemical co-precipitation method: a comparative study
Received
16th March 2015
, Accepted 8th June 2015
First published on 8th June 2015
Abstract
In this study, MFe2O4 (M = Zn, Cu, Co and Ni) nanoparticles were prepared by a modified chemical co-precipitation method followed by annealing. Then, Ag/MFe2O4 nanoparticles were synthesised via a wetness impregnation method. The structural properties of the samples were systematically investigated by XRD, SEM, TEM, EDX, DRS, BET, SPV and FT-IR techniques. The photocatalytic performances of the prepared MFe2O4 and Ag/MFe2O4 samples were comparatively studied by the photodegradation of 4-chlorophenol under xenon lamp illumination in a FTIR reaction cell. The photodegradation rate of 4-chlorophenol was 90% for the Ag/CoFe2O4 sample. The cheap cost and easy synthesis of the photocatalysts indicate the potential for a better thermal disposal of chlorophenol pollutants at lower temperatures.
1. Introduction
Over the past decades, chlorophenols have been an important organic pollutant of water pollutants.1 They are listed as one of 129 priority pollutants by the U.S. EPA.2 In the 1950s, 4-chlorophenol (4-CP) was a typical pollutant among these chlorophenols, and was used in disinfectants, antiseptics and herbicides, and was also applied in the large-scale disinfection of drinking water.2–6 It can result in negative effects such as carcinogenic, teratogenic, mutagenic and acute toxicity. It is difficult to dispose of these pollutants by usual treatment methods. Compared to the conventional multi-treatment processes, photocatalysis is an effective method which transforms chlorophenol organic compounds into small inorganic molecules.
TiO2 is known as an efficient photocatalyst because of its nontoxicity, low-cost, and photostability, which has led to many promising applications in photovoltaics,7 photocatalysis,8 photo-electrochromics and sensors.9 J. Theurich et al. reported that TiO2-based materials showed preeminent photocatalytic activities in the photodegradation of 4-CP under UV-irradiation.10 P. Rangsunvigit et al. reported that the addition of Pt in TiO2 can improve the catalytic activity and the highest activity was obtained with 1% Pt/TiO2. For TiO2–SiO2, the highest activity was achieved with 10% SiO2–TiO2 which has the highest adsorption capacity.11 However, TiO2 nanomaterials are normally transparent in the visible light region. By developing new visible light-induced catalysts, it is possible to satisfy sustainable energy demands.
Recently, some composite oxides such as spinel AB2O4 have shown better selectivity and sensitivity to certain gases than single-metal oxides. Nanoparticles of spinel ferrites MFe2O4 (M = Zn, Co, Ni, Cu etc.) are an emerging new type of catalyst which people have shown great interest in by addressing the relationship between their physical chemistry properties and their crystal structure.12 Some synthetic methods to prepare MFe2O4 ferrites such as sol–gel,13 micro-emulsion,14,15 citrate gel,16 precipitation,17 precursor techniques18–20 host template,21 and mechanical alloying22 have been developed. A chemical co-precipitation method is more suitable to prepare nanostructured oxides because of its simplicity and good control of grain size. B. Veronica et al. have reported superparamagnetism and interparticle interactions in ZnFe2O4 nanocrystals.23 The method could also be extended to synthesize CoFe2O4, CuFe2O4 and NiFe2O4 nanocrystals. J. Jiang et al.24 reported nanostructured NiFe2O4 via a refluxing route which exhibited typical superparamagnetic behavior at room temperature, with a present finite coercivity of 353 Oe at 5 K. S. Saadi et al.25 and H. Yang et al.26 studied photoassisted hydrogen evolution over CuFe2O4, and they found that CuFe2O4 exhibited promising photocatalytic activity. S. Li et al.27 have studied the fabrication and characterization of TiO2/BaAl2O4. J. Okal et al.28 have reported the catalytic combustion of methane over Ru/ZnAl2O4 nanoparticles. C. Ragupathi et al.29 have studied the photocatalytic activities of ZnAl2O4 which was prepared by a microwave method.
Supported noble metal catalysts have better activities and handling properties. Moreover, the synthesis conditions are mild. Noble metal/semiconductor oxide composites are very attractive because of their enhanced efficiency of photocatalytic activity. It was reported that the addition of the noble metals Pt, Au, Rh to TiO2 can enhance the photocatalytic efficiency.30 An efficient support, including activated carbon,31 silica,32 alumina33 and zeolites,34 could improve the activity, recycling, and selectivity of Ag catalyst systems. Recently, Liu Yang et al.35 reported that Ag nanoparticles in mesoporous TiO2 were not only an antimicrobial but also an intensifier of photocatalysis. Jia Ren et al.36 reported Ag-loaded Bi2WO6 photocatalysts that exhibited a significant increase in the photocatalytic activity of inactivating E. coli, a Gram-negative bacterium, and S. epidermidis, a Gram-positive bacterium under visible light irradiation (λ > 420 nm). Recently, spinel type oxide catalysts have been used as supports for Ru and Pd and a distinct metal–support interaction has been found.37,38 However, the photocatalytic properties and the activities of these spinel type oxides for the degradation of 4-CP have not been reported.
In this work, MFe2O4 (M = Zn, Co, Ni and Cu) nanoparticles with a brick-like spinel superstructure were prepared by a modified chemical co-precipitation method and post calcinations at 673 K, 773 K and 873 K, and then 1% wt noble metal Ag loading on the MFe2O4 (M = Zn, Co, Ni and Cu) (673 K) by a wetness impregnation method, which produced the Ag/MFe2O4 nanomaterials. These MFe2O4 and Ag/MFe2O4 catalysts were studied for their photocatalytic degradation of 4-CP under xenon lamp irradiation. The photocatalytic mechanism is discussed, thereby showing the method of environmental purification and the need for new active catalysts.
2. Experimental section
2.1 Materials and methods
The MFe2O4 (M = Zn, Co, Ni and Cu) catalysts with brick-like superstructures were prepared by a co-precipitation method. The starting materials were FeSO4·7H2O and MSO4·nH2O (M = Zn, Co, Ni and Cu). A 0.1 M (100 mL) solution of FeSO4·7H2O and 0.2 M (100 mL) MSO4·nH2O was mixed in deionized water. A 0.3 M (25 mL) solution of Na2C2O4 was prepared and added to the salt solution. The solution was continuously stirred at 353 K and allowed to cool slowly to room temperature. A ferrioxalate precipitate appeared. Then the precipitate was washed twice with distilled water and ethanol, respectively. The separated precipitate was dried overnight at 373 K. The precursor was divided into three portions and calcined at 673 K, 773 K and 873 K for 2 h, respectively. Consequently, MFe2O4 (M = Zn, Co, Ni and Cu) samples were produced. The samples are named in Table 1.
Table 1 The crystallite size, calcination temperature and photodegradation ratio for 4-CP of the synthesized samples
Abbreviated name of synthesized samples |
Prepared samples |
Calcination temperature (K) |
Average crystallite size (nm) |
Photocatalytic degradation ratio of 4-CP with the sample (%) |
Ag/Zn |
Ag/ZnFe2O4 |
673 |
13 |
88 |
Zn1 |
ZnFe2O4 |
673 |
14 |
81 |
Zn2 |
ZnFe2O4 |
773 |
15 |
77 |
Zn3 |
ZnFe2O4 |
873 |
18 |
65 |
Ag/Co |
Ag/CoFe2O4 |
673 |
11 |
90 |
Co1 |
CoFe2O4 |
673 |
12 |
79 |
Co2 |
CoFe2O4 |
773 |
14 |
70 |
Co3 |
CoFe2O4 |
873 |
16 |
60 |
Ag/Ni |
Ag/NiFe2O4 |
673 |
8 |
86 |
Ni1 |
NiFe2O4 |
673 |
8 |
78 |
Ni2 |
NiFe2O4 |
773 |
10 |
68 |
Ni3 |
NiFe2O4 |
873 |
14 |
55 |
Ag/Cu |
Ag/CuFe2O4 |
673 |
25 |
81 |
Cu1 |
CuFe2O4 |
673 |
26 |
62 |
Cu2 |
CuFe2O4 |
773 |
22 |
58 |
Cu3 |
CuFe2O4 |
873 |
21 |
52 |
Without catalyst |
|
|
|
7.5 |
The catalysts of MFe2O4 (M = Zn, Co, Ni and Cu) (calcined at 673 K) with Ag supports were prepared by an impregnation strategy.39 AgNO3 solution was used as the precursor. The silver loading was 1 wt%. After evaporation, the catalyst was then dried at 100 °C for 12 h and calcined at 673 K in air for 6 h. Finally, the catalyst was slowly cooled to room temperature. The Ag/MFe2O4 (M = Zn, Co, Ni and Cu) samples are named in Table 1.
All chemicals in this work were analytical grade reagents and used as starting materials without further purification.
2.2 Characterization of the catalysts
The crystal structures of the MFe2O4 and Ag/MFe2O4 (M = Zn, Co, Ni and Cu) samples were examined by X-ray diffraction (XRD, Rigaku D/max) with Cu Kα radiation (λ = 0.15418 nm) over the 2θ range of 20–80°. The surface area was measured by Brunauer–Emmett–Teller (BET), and N2 gas adsorption–desorption isotherms were determined using a Micromeritics ASAP-2000. Pore volume was calculated by the BJH model. The light absorption properties were measured using a UV-Vis diffuse reflectance spectrophotometer (JASCO, UV-550) with a wavelength range of 200–800 nm. The morphology of these prepared samples was characterized by scanning electronic microscopy (SEM) with a JSM-6700 LV and by transmission electron microscopy (TEM, Hitachi 800 system at 200 kV). Energy dispersive X-ray analysis (EDX, Horiba 7593 H) was performed to determine the elemental concentration distribution of the samples. Fourier transform infrared (FTIR, BRUKER VERTEX 70) spectra were recorded in the range 4000–450 cm−1 with a 4 cm−1 resolution. XPS data were obtained using a Perkin-Elmer PHI 5600 electron spectrometer which used achromatic Al Kα radiation (1486.6 eV) with Ar+ sputtering to remove the surface layer of the sample.
The characteristics of photogenerated charge carriers in the visible light spectrum were examined by a lock-in-based surface photovoltage (SPV) measurement system. The system consists of monochromatic light from a monochromator (Omni-λ3005) and a xenon lamp (500 W), a photovoltaic sample cell and a lock-in amplifier (SR830-DSP) with an optical chopper (model SR540). The AC photovoltage signal of the sample was obtained using a capacitor structure. The highest intensity monochromatic light in the spectrum was less than 80 μW cm−2. The effective overlapping area of the two electrodes was about 1 cm2 for testing. The corresponding phase spectra were taken with the SPV spectra by a computer. SPV measurements were carried out in an ambient environment and the raw data were used directly.
2.3 Measurement of photocatalytic activities
The photodegradation of 4-chlorophenol was performed under a xenon lamp (XQ-500 W) with an intensity of 40 mW cm−2. The reaction system was a self-made photochemical reactor, demonstrated in ref. 40. In the test, the initial concentration of the 4-CP (C6H4ClO) solution was 20 mg L−1. Reaction suspensions were prepared by adding 0.01 g catalyst powder into the 4-CP solution (100 mL). The reaction suspension was stirred in a magnetic stirrer in the dark for 30 minutes to reach an adsorption/desorption equilibrium. Then the test was performed under the xenon lamp irradiation. After a given irradiation time, 4.0 mL samples were withdrawn, and the collected samples are centrifuged at 6000 rpm for 5 minutes in order to separate the photocatalyst from the 4-CP solution. The absorption spectra of 4-CP were measured by a UV-Vis spectrophotometer.
3. Results and discussion
3.1 XRD analysis
From Fig. 1a, the undoped NiFe2O4 samples Ni1, Ni2, Ni3 and the Ag-doped NiFe2O4 sample Ag/Ni can be assigned to the inverse spinel structure, according to the standard XRD patterns (JCPDS patterns no. 21-1272). The peaks with 2θ values of 30.6°, 36.0°, 44.0°, 57.8° and 63.2° correspond to the crystal planes (220), (311), (400), (511), and (440) of crystalline NiFe2O4, respectively.
 |
| Fig. 1 (a) The XRD patterns of the NiFe2O4 (Ni1, Ni2 and Ni3) samples and the Ag/NiFe2O4 (Ag/Ni) sample, (b) the CoFe2O4 (Co1, Co2 and Co3) samples and the Ag/CoFe2O4 (Ag/Co) sample, (c) the CuFe2O4 (Cu1, Cu2 and Cu3) samples and the Ag/CuFe2O4 (Ag/Cu) sample, and (d) the as-synthesized precursor, the ZnFe2O4 (Zn1, Zn2 and Zn3) samples and the Ag/ZnFe2O4 (Ag/Zn) sample. | |
The crystallite size was calculated by the Scherrer equation:
D = 0.89λ/β cos θ |
where
D is the average crystal size,
λ is the wavelength of Cu Kα (0.15406 nm),
β is the full width at half maximum, and
θ is the diffraction angle.
41
The average crystallite size is calculated from the (311) peak of the crystal plane using Scherrer's equation. In this case the average crystallite sizes were 8 nm for Ni1, 10 nm for Ni2, 14 nm for Ni3 and 8 nm for Ag/Ni (see Table 1).
Fig. 1b shows the XRD patterns of the undoped CoFe2O4 samples Co1, Co2, Co3 and the Ag-doped CoFe2O4 sample Ag/Co. The results indicate that the final product is CoFe2O4 with the expected inverse spinel structure. The peaks with 2θ values of 30.7° (with index number of 220), 36.3° (311), 43.6° (400), 57.9° (511), and 63.2° (440) reveal the formation of CoFe2O4. The average crystallite sizes of these samples are about 12 nm (Co1), 14 nm (Co2), 16 nm (Co3), and 11 nm (Ag/Co).
The XRD patterns of the undoped CuFe2O4 samples Cu1, Cu2, Cu3 and the Ag-doped CuFe2O4 sample Ag/Cu are shown in Fig. 1c. In the XRD pattern of Cu1 and Ag/Cu, a phase of copper ferrite appears, and the diffraction angle and intensity of the characteristic peaks are consistent with that of the standard JCPDS card no. 34-0425. Copper ferrite with a spinel crystal structure phase can be directly formed after calcination (673 K), and it does not require other synthetic techniques.42 There exists a small amount of CuO impurity phase in the as-obtained CuFe2O4 when the calcination temperature is increased (773 K and 873 K). From the XRD data in Fig. 1c, the average sizes of these samples are about 26 nm (Cu1), 22 nm (Cu2), 21 nm (Cu3), and 21 nm (Ag/Cu).
Fig. 1d shows the XRD patterns of the as-synthesized precursor for ZnFe2O4 and the Zn1, Zn2, Zn3 and Ag/Zn samples. It was noted that the diffraction peaks of the as-synthesized precursor did not match the phase of zinc ferrite. The observed diffraction peaks from the (220), (311), (400), (422), (511) and (440) crystal planes are well matched with the cubic spinel structure of the ZnFe2O4 phase (JCPDS card no. 22-1012). From the XRD data of Fig. 1d, the average sizes of these samples are about 14 nm (Zn1), 15 nm (Zn2), 18 nm (Zn3), and 13 nm (Ag/Zn). The average crystallite size of the Ag/MFe2O4 sample seems smaller than the undoped MFe2O4. This phenomenon is probably due to the reaction between MFe2O4 and HNO3 where some of the MFe2O4 particles were destroyed into smaller ones.43
The crystallinity will increase as the calcining temperature rises. Despite this, a little of the CuO and Fe2O3 phases existed in the powders which were calcined below 773 K.44 The augmentation in the calcination temperature of the CuFe2O4 sample led to a decrease in the degree of ordering and crystallite size of both the CuO and Fe2O3 phases. This decrease may be due to the solid-state reaction between CuO and Fe2O3 solids yielding a nanocrystalline CuFe2O4 catalyst.45
From Fig. 1, the peaks due to Ag species cannot be observed due to the low doping content of Ag in the samples. As the calcination temperature increases, the peak intensity of the MFe2O4 samples also increases. This indicates that the calcination temperature plays a role in the formation of the spinel crystal structure and morphology. The average crystallite size of Ag/MFe2O4 samples seems smaller than undoped MFe2O4 samples. The results are consistent with the description reported by X. Li et al.46
The specific surface areas of the MFe2O4 and Ag/MFe2O4 samples are shown in Table 2. The surface areas of Ag/Co, Ag/Zn, Ag/Ni, Ag/Cu are 187, 166, 153.5 and 142.3 m2 g−1, respectively. The mesoporous Ag/CoFe2O4 has the highest specific surface area. The pore volumes of Ag/Co, Ag/Zn, Ag/Ni, Ag/Cu are 0.12, 0.162, 0.181 and 0.20 cm3 g−1, respectively. The results show that the specific surface areas of the samples with silver loading are greater than the other samples. With the increase in the specific surface area, the pore volume will be reduced.
Table 2 The specific surface area and pore volume of the MFe2O4 (M = Zn, Cu, Co and Ni) and Ag/MFe2O4 samples
Abbreviation name |
Prepared sample |
BET surface area/m2 g−1 |
Pore volume/cm3 g−1 |
Ag/Zn |
Ag/ZnFe2O4 |
166 |
0.162 |
Zn1 |
ZnFe2O4 |
138 |
0.22 |
Ag/Co |
Ag/CoFe2O4 |
187 |
0.12 |
Co1 |
CoFe2O4 |
136 |
0.257 |
Ag/Ni |
Ag/NiFe2O4 |
153.5 |
0.181 |
Ni1 |
NiFe2O4 |
129 |
0.29 |
Ag/Cu |
Ag/CuFe2O4 |
142.3 |
0.20 |
Cu1 |
CuFe2O4 |
108 |
0.34 |
3.2 SEM and TEM analysis
The SEM images in Fig. 2 show the morphology of Co1 (Fig. 2a), Co2 (Fig. 2c), Co3 (Fig. 2b), and Ag/Co (Fig. 2d). The SEM images in Fig. 3 show the morphology of Ni1 (Fig. 3a), Ni2 (Fig. 3c), Ni3 (Fig. 3b) and Ag/Ni (Fig. 3d). The CoFe2O4 and NiFe2O4 powders calcined at different temperatures are completely composed of “timber”. Fig. 2e shows the TEM micrograph of Ag/Co. Fig. 3e and f show the TEM micrographs of Ag/Ni. Both the porous blocks (Ag/Ni and Ag/Co) were formed through the agglomeration of numerous Ag/Ni and Ag/Co nanoparticles, respectively.
 |
| Fig. 2 The SEM images of the calcined CoFe2O4 samples at 673 K (Co1) (a), 773 K (Co2) (c), 873 K (Co3) (b) and the Ag/CoFe2O4 (Ag/Co) sample at 673 K (d), and (e) the TEM image of the calcined Ag/CoFe2O4 sample (Ag/Co) at 673 K. | |
 |
| Fig. 3 The SEM images of the calcined NiFe2O4 samples at 673 K (Ni1) (a), 773 K (Ni2) (c), 873 K (Ni3) (b) and the Ag/NiFe2O4 (Ag/Ni) sample at 673 K (d), and (e and f) the TEM images of the calcined Ag/NiFe2O4 sample (Ag/Ni) at 673 K. | |
Fig. 4 presents the TEM images of the Cu1 and Ag/Cu samples. It is found that blocks (Fig. 4a) are formed through the agglomeration of numerous nanocrystalline CuFe2O4 particles. Fig. 4b shows the TEM micrograph of the Ag/Cu sample. The nanoparticles distinctly exhibit a narrow size distribution with sizes ranging from 20 to 30 nm. Fig. 5 presents the TEM images of the Zn1 and Ag/Zn samples. It is found that blocks (Fig. 5a) are formed through the agglomeration of numerous nanocrystalline ZnFe2O4 particles. Fig. 5b shows the TEM micrograph of the Ag/Zn sample. The nanoparticles distinctly exhibit a narrow size distribution with sizes of 20 nm, approximately. These results are in good agreement with the XRD analysis.
 |
| Fig. 4 The TEM images of the CuFe2O4 (Cu1) (a) and the Ag/CuFe2O4 (Ag/Cu) samples (b), and the EDS pattern of the CuFe2O4 (Cu1) sample (c). | |
 |
| Fig. 5 The TEM images of the ZnFe2O4 (Zn1) (a) and the Ag/ZnFe2O4 (Ag/Zn) samples (b), and the EDS pattern of the Ag/ZnFe2O4 (Ag/Zn) sample (c). | |
From the above images, we cannot distinguish the Ag species in the samples. So the EDS patterns of the prepared Cu1 and Ag/Zn samples were measured. The EDS spectrum in Fig. 4c exhibits Cu, Fe and O peaks and also reflects the atomic ratio of Cu, Fe and O in the prepared Cu1 sample (Table 3). The peaks due to metallic silver are identified in the EDS pattern (Fig. 5c) of the Ag/Zn sample. Table 4 exhibits the atomic ratio of the elements in the prepared Ag/Zn sample.
Table 3 The atomic ratio of the elements in the CuFe2O4 (Cu1) sample
Element |
Weight ratio (%) |
Atomic ratio (%) |
C |
29.34 |
44.34 |
O |
40.99 |
46.46 |
Fe |
19.03 |
6.16 |
Cu |
10.64 |
3.04 |
Total |
100.00 |
100.00 |
Table 4 The atomic ratio of the elements in the Ag/ZnFe2O4 (Ag/Zn) sample
Element |
Weight ratio (%) |
Atomic ratio (%) |
C |
31.25 |
47.59 |
O |
37.46 |
42.79 |
Fe |
19.44 |
6.34 |
Zn |
11.61 |
3.24 |
Ag |
0.24 |
0.04 |
Total |
100.00 |
100.00 |
3.3 DRS analysis
The UV-Vis spectra of the MFe2O4 and Ag/MFe2O4 (M = Zn, Co, Ni and Cu) samples are shown in Fig. 6. The MFe2O4 and Ag/MFe2O4 samples have strong absorption in the range of the whole spectrum (200–800 nm). As shown in Fig. 6, we clearly observed the characteristic bands of absorption at 500–700 nm. Wider and stronger absorption spectra are found for Ag/MFe2O4 samples.
 |
| Fig. 6 (a) The UV-Vis absorption spectra of the NiFe2O4 (Ni1, Ni2 and Ni3) samples and the Ag/NiFe2O4 (Ag/Ni) sample, (b) the CoFe2O4 (Co1, Co2 and Co3) samples and the Ag/CoFe2O4 (Ag/Co) sample, (c) the CuFe2O4 (Cu1, Cu2 and Cu3) samples and the Ag/CuFe2O4 (Ag/Cu) sample, and (d) the ZnFe2O4 (Zn1, Zn2 and Zn3) samples and the Ag/ZnFe2O4 (Ag/Zn) sample. | |
Due to the mesoporous structure, the photo-generated electron–hole pairs in the Ag/MFe2O4 nanocrystal are separated mainly via diffusion. With the improved separation efficiency of these photo-induced electron–hole pairs into charge carriers, the SPV response and photocatalytic activity of Ag/MFe2O4 are higher than the MFe2O4 samples. Moreover, because the Fermi level of Ag (ca. 4.3 eV)47 is much lower in electronic energy than the conduction band edge of the MFe2O4 catalysts,48 the Ag species probably adsorbed onto the surface of MFe2O4 plays a key role in trapping and transporting the photo-generated electrons. Therefore, the Ag/MFe2O4 sample exhibits a higher photocatalytic activity than MFe2O4.
For the Ag/CoFe2O4 sample, the absorption intensity is visibly stronger than the other samples. So the Ag/CoFe2O4 sample may have a better absorption capability to visible light.
3.4 FTIR analysis
The FTIR spectra of the MFe2O4 and Ag/MFe2O4 (M = Zn, Co, Ni and Cu) samples are shown in Fig. 7. The main absorption bands γ1 and γ2 around 600 and 400 cm−1 are attributed to the stretching vibrations of the tetrahedral and octahedral sites, respectively, which are responsible for the formation of MFe2O4.49 The band at 580 cm−1 is attributed to the Fe–O bond vibration of the samples.50,51
 |
| Fig. 7 (a) The FTIR spectra of the NiFe2O4 (Ni1, Ni2 and Ni3) samples and the Ag/NiFe2O4 (Ag/Ni) sample, (b) the CoFe2O4 (Co1, Co2 and Co3) samples and the Ag/CoFe2O4 (Ag/Co) sample, (c) the CuFe2O4 (Cu1, Cu2 and Cu3) samples and the Ag/CuFe2O4 (Ag/Cu) sample, and (d) the ZnFe2O4 (Zn1, Zn2 and Zn3) samples and the Ag/ZnFe2O4 (Ag/Zn) sample. | |
The peaks at 3455 cm−1, 1340 cm−1 and 1610 cm−1 can be assigned to the stretching vibrations of hydrogen-bonded surface water molecules and hydroxyl groups.52 The peak at 1185 cm−1 is attributed to the C–O stretching vibration mode.53 The peaks around 1000 cm−1 (930 and 1030 cm−1) in Fig. 7c and d correspond to the C–H stretching vibrations in C
C bonds.54,55
3.5 XPS analysis
The XPS analysis of the MFe2O4 (M = Zn, Cu, Co and Ni) and Ag/MFe2O4 samples are shown in Fig. 8–11. Fig. 8 shows the high resolution spectra of the ZnFe2O4 and Ag/ZnFe2O4 samples. Fig. 8a shows the Zn peaks of Zn 2p3/2 and Zn 2p1/2 at 1023.4 eV and 1047.9 eV, respectively. Fe peaks of Fe 2p1/2 and Fe 2p3/2 at 728.9 eV and 715 eV, respectively, and the O 1s peak at 534.1 eV are shown in Fig. 8b and c, which indicate the existence of ZnFe2O4 in the sample. Fig. 8d shows the spectra of the Ag peak with Ag 3d3/2 at 372.1 eV. No contaminating species were observed within the sensitivity of this technique.
 |
| Fig. 8 The XPS spectra of the ZnFe2O4 and Ag/ZnFe2O4 samples, the Zn peaks (a), the Fe peaks (b), the O peak (c) and the Ag peak (d). | |
Fig. 9 shows the high resolution XPS spectra of Co 2p, Fe 2p, O 1s and Ag 3d. In Fig. 9a, two peaks appear at 780.31 eV and 788.2 eV, which are attributable to Co 2p3/2. From Fig. 9b, the peaks of Fe 2p3/2 and Fe 2p1/2 are located at 710.98 eV and 724.25 eV, respectively. Fig. 9c shows the XPS spectra of the C 1s peak at 530.5 eV. These peaks indicate the existence of CoFe2O4 in the sample. Fig. 9d shows the XPS spectra of the Ag peak with Ag 3d5/2 at 368.8 eV.
 |
| Fig. 9 The XPS spectra of the CoFe2O4 and Ag/CoFe2O4 samples, the Co peaks (a), the Fe peaks (b), the O peak (c) and the Ag peak (d). | |
Fig. 10 shows the XPS spectra of the NiFe2O4 and Ag/NiFe2O4 samples. The peak located at 860.2 eV is attributed to the Ni 2p3/2 as shown in Fig. 10a. From Fig. 10b, the peaks of Fe 2p3/2 and Fe 2p1/2 are located at 710.9 eV and 725.2 eV, respectively. Fig. 10c represents the XPS spectra of O 1s with a single peak at 530 eV. The spectrum shows that the Ag 3d3/2 is approximately at 372.6 eV in Fig. 10d.
 |
| Fig. 10 The XPS spectra of the NiFe2O4 and Ag/NiFe2O4 samples, the Ni peaks (a), the Fe peaks (b), the O peak (c) and the Ag peak (d). | |
Fig. 11 shows the XPS spectra of the CuFe2O4 and Ag/CuFe2O4 samples. Fig. 11a shows the peaks of Cu with Cu 2p3/2 at 936.7 eV. The Fe 2p shows two peaks at binding energies of 728.9 and 715.7 eV which are attributable to Fe 2p1/2 and Fe 2p3/2, respectively. Fig. 11c shows the O peak of O 1s at 534.5 eV. Fig. 11d shows the XPS spectra of the Ag peak with Ag 3d3/2 at 372.8 eV.
 |
| Fig. 11 The XPS spectra of the CuFe2O4 and Ag/CuFe2O4 samples, the Cu peaks (a), the Fe peaks (b), the O peak (c) and the Ag peak (d). | |
3.6 SPV analysis
Fig. 12 shows the SPV amplitudes and phase spectra of the MFe2O4 and Ag/MFe2O4 (M = Zn, Co, Ni and Cu) samples. The Ag/MFe2O4 samples exhibit a larger SPV response than the MFe2O4 samples. The higher SPV signal may imply the higher separation rate of the photo-generated charge carriers.56 From Fig. 12b, d, f and h, the SPV phase values for the Ag/MFe2O4 samples below 600 nm are about 90°. This indicates that the photo-generated electrons accumulate on the surface of the samples.57 It could be observed by comparison that the MFe2O4 samples have a smaller phase retardation. The electronic band gaps of the dispersed nanocrystals in the “timber-like” aggregation of Ag/MFe2O4 would improve the redox activity of their photo-generated electrons and holes. In Fig. 12g, the SPV spectrum of the Ag/ZnFe2O4 sample has a 50 nm blue shift with respect to ZnFe2O4. Moreover, the SPV signal of the Ag/CoFe2O4 sample is significantly larger than CoFe2O4 and the other samples (Fig. 12a). For Ag/CoFe2O4, the efficiency of charge separation is higher than the other samples, so that Ag/CoFe2O4 has higher photocatalytic activity.
 |
| Fig. 12 The specific surface photovoltage spectra (a and c) and corresponding phase spectra (b and d) of the MFe2O4 (M = Co1, Co2, Co3 and Ni1, Ni2, Ni3) samples and the Ag/MFe2O4 (M = Co, Ni) samples, and the specific surface photovoltage spectra (e and g) and corresponding phase spectra (f and h) of the MFe2O4 (M = Cu1 and Zn1) samples and the Ag/MFe2O4 (M = Cu and Zn) samples. | |
3.7 Comparison of the photocatalytic activities of the MFe2O4 and Ag/MFe2O4 samples
Fig. 13 shows the comparison of the photocatalytic activities of the MFe2O4 and Ag/MFe2O4 (M = Co, Zn, Ni and Cu) samples, studied by the photodegradation of a solution of 4-CP under xenon lamp illumination. After 120 minutes, the 4-CP solution cannot be photodegraded under xenon lamp irradiation without any catalyst, while the 4-CP solution can be photodegraded by the MFe2O4 and Ag/MFe2O4 samples. The degradation ratios of the 4-CP solution with the MFe2O4 and Ag/MFe2O4 samples are shown in Table 1 and Fig. 13a. The results show that the photodegradation ratio of the 4-CP solution with the Ag/CoFe2O4 sample is up to 90% and that the photodegradation ratios with the Ag/ZnFe2O4, Ag/NiFe2O4 and Ag/CuFe2O4 samples are 88%, 86% and 81%, respectively. This implies that the Ag/CoFe2O4 sample has a higher photocatalytic activity than the other samples.
 |
| Fig. 13 (a) The photocatalytic degradation ratio of 4-CP with the MFe2O4 and the Ag/MFe2O4 (M = Co, Ni, Cu and Zn) samples, and (b) the concentration change of 4-CP over 120 min under xenon light illumination. | |
There are many influential factors which have an impact on the decomposition rate, such as the following: (1) MFe2O4 nanoparticles absorb photon energy greater than the band gap energy and generate photo-generated electron–hole pairs in the bulk; (2) photoexcited carriers separate and move to the surface without recombination; (3) water molecules (or H+) are reduced and oxidized by the photo-generated electrons and holes to produce H2 and O2, respectively;58 (4) redox reactions with adsorbed reactants.59 Metal properties have some influence, but the above factors are clearly more important than the nature of the metal on the degradation efficiency.
According to XRD studies, the crystallite size of Ag/MFe2O4 samples is smaller than that of the MFe2O4 samples. The Ag/MFe2O4 samples have more hydrogen-bonded, surface water molecules and hydroxyl groups adsorbed at the surface due to the smaller crystallite size. Radical groups have been shown to be capable of contributing to the oxidation process of organic substances. With the decrease in particle size, the catalytic activity of the Ag/CoFe2O4 sample is enhanced because of the widened band gap. Moreover, from UV-vis-DRS analysis of the MFe2O4 and Ag/MFe2O4 (M = Co, Zn, Ni and Cu) samples, a Ag dopant enhances the absorbance of visible light. Furthermore, small particles of the semiconductor became almost totally depleted upon immersion in the electrolyte (allowing for large photovoltages).60
We learned by reading the literature that there are many loading quantities, such as 0.05 wt%,61 0.98 wt%, 1.3 wt%,62 1.5 wt%,63 3.2 wt%64 and 12 wt%.65 It was concluded through experiments that the sample with 1 wt% Ag loading has the strongest adsorption capacity. Samples with Ag loading are not only antimicrobial but also enhance photocatalysis. Three factors are ascribed to the excellent antibacterial abilities of the mesoporous Ag/MFe2O4 samples. First, their high specific surface area and small particle size, which would result in better activity. Second, Ag nanoparticles are an excellent antimicrobial material, even though only a tiny amount of Ag is doped in the pores of Ag/MFe2O4 (Ag/M < 0.01, mole ratio). Furthermore, the Ag nanoparticles could act as trapping centers for photo-generated electrons in the conduction band of MFe2O4, leaving holes in the valence band of MFe2O4 and inhibiting the recombination of electrons and holes. These factors will lead to an improved quantum efficiency for photochemical reactions.66,67 This implies that the Ag/CoFe2O4 sample calcined at 673 K shows a more enhanced photocatalytic activity than the other samples.
3.8 Stability and reusability of the MFe2O4 and Ag/MFe2O4 (M = Co, Zn, Cu, and Ni) catalysts
We also conducted stability experiments with the MFe2O4 and Ag/MFe2O4 samples over five reaction cycles under the same xenon lamp illumination test conditions. The photodegradation ratio of 4-CP over Ag/CoFe2O4 is shown in Fig. 14a and reaches 75.3% after the fifth repetition, which is slightly lower than that (90%) of the fresh Ag/CoFe2O4 catalyst. The photodegradation ratio of the Ag/ZnFe2O4, Ag/NiFe2O4 and Ag/CuFe2O4 samples after the fifth repetition are 71%, 68.1% and 65.2%, respectively. The results show that MFe2O4 and Ag/MFe2O4 have good stability and reusability as catalysts.
 |
| Fig. 14 The photocatalytic degradation ratio of 4-CP over the Ag/CoFe2O sample (a), Ag/ZnFe2O4 sample (b), Ag/NiFe2O4 sample (c), and Ag/CuFe2O4 sample (d) over five repetitions. | |
3.9 Intermediate analysis of the photocatalytic degradation of 4-CP
The main intermediates of the photocatalytic oxidation of 4-CP were hydroquinone (HQ), benzoquinone (BQ), and hydroxyhydroquinone (HHQ) as measured by HPLC.
Under dark conditions, the adsorption/desorption processes were equilibrated. 4-CP quickly disappeared during the first 30 min due to its adsorption on the sample surface.68 Intermediates produced by the photo-oxidation of 4-CP interact with the catalyst on the surface so that they can be released to the gas phase. Hydroxyl groups on the surface of the catalyst are able to react with the intermediates, which are then retained on the catalyst surface. Subsequently they can be converted into other products, like CO2 and H2O.69
4-CP was almost fully degraded, but 44% of the total organic carbon was still present in the solution after 120 minutes. We can learn from this phenomenon that the degradation efficiency was greater than the mineralization efficiency. Therefore, there are transient organic intermediates present in the photocatalytic system. The intermediate products were subsequently oxidized rapidly by the heterojunction anode.70 We can come to the conclusion that we should prolong the illumination time for complete mineralization.71
4. Conclusions
In summary, MFe2O4 (M = Zn, Cu, Co and Ni) nanocrystals were synthesized by a modified chemical co-precipitation method using different calcining temperatures (673 K, 773 K and 873 K). Then, based on the MFe2O4 (673 K) sample, we prepared the Ag/MFe2O4 (M = Zn, Cu, Co and Ni) catalysts through the traditional wetness impregnation strategy. Within the visible region of the spectra, these Ag/MFe2O4 nanocrystals possess attractive photovoltage responses, enhanced photocatalytic activities and good stabilities and reusabilities for the degradation of 4-CP. The Ag/CoFe2O4 catalyst with a spinel crystal structure annealed at 673 K promotes the adsorbates onto the active sites and results in a higher photo-induced reactivity. So the Ag/CoFe2O4 nanoparticles investigated here have an expected and remarkable photocatalytic activity and could be potentially applied in environmental purification in the future.
Acknowledgements
The project was supported by Open Research Fund of Key Laboratory of Subsurface Hydrology and Ecological Effect in Arid Region of Ministry of Education (no. 2014G1502033), Engineering Research Center of Groundwater and Eco-Environment of Shaanxi Province, National Nature Science Foundation of China (no. 20877013 and NSFC-RGC 21061160495), the National High Technology Research and Development Program of China (863 Program) (no. 2010AA064902), the Major State Basic Research Development Program of China (973 Program) (no. 2011CB936002), the Excellent Talents Program of Liaoning Provincial University (LR2010090) and the Key Laboratory of Industrial Ecology and Environmental Engineering, China Ministry of Education.
References
- L. H. Keith and W. A. Telliard, ES&T special report: priority pollutants: I-a perspective view, Environ. Sci. Technol., 1979, 13, 416–423 CrossRef.
- V. Janda and M. Svecova, By-products in drinking water disinfection, Chem. Listy, 2000, 94, 905–908 CAS.
- M. Czaplicka, Sources and transformation of chlorophenols in the natural environment, Sci. Total Environ., 2004, 322, 21–39 CrossRef CAS PubMed.
- R. Cheng, J. Wang and W. Zhang, Comparison of reductive of p-chlorophenol using FeO and nanosized FeO, J. Hazard. Mater., 2007, 144, 334–339 CrossRef CAS PubMed.
- Y. Wang, K. Chan, X. Li and S. K. So, Electrochemical degradation of 4-chlorophenol at nickel-antimony doped tin oxide electrode, Chemosphere, 2006, 65, 1087–1093 CrossRef CAS PubMed.
- M. Czaplicka and B. Kaczmarczyk, Infrared study of chlorophenols and products of their photodegradation, Talanta, 2006, 70, 940–949 CrossRef CAS PubMed.
- B. O'Regan and M. Gratzel, A low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films, Nature, 1991, 353, 737–740 CrossRef PubMed.
- X. Zhang, M. Jin, Z. Liu, S. Nishimoto, H. Saito, T. Murakami and A. Fujishima, Preparation and photocatalytic wettability conversion of TiO2-based superhydrophobic surfaces, Langmuir, 2006, 22, 9477–9479 CrossRef CAS PubMed.
- S. Liu and A. Chen, Coadsorption of horseradish peroxidase with thionine on TiO2 nanotubes for biosensing, Langmuir, 2005, 21, 8409–8413 CrossRef CAS PubMed.
- J. Theurich, M. Lindner and D. W. Bahnemann, Photocatalytic degradation of 4-chlorophenol in aerated aqueous titanium dioxide suspensions: a kinetic and mechanistic study, Langmuir, 1996, 12, 6368–6376 CrossRef CAS.
- P. Rangsunvigit, R. Tharathonpisutthikul, S. Chavadej and E. Gulari, Photocatalytic degradation of 4-chlorophenol: effects of Pt and SiO2, Chem. Lett., 2012, 41, 1371–1373 CrossRef CAS.
- S. Ayyappan, G. Paneerselvam, M. P. Antony and J. Philip, Structural stability of ZnFe2O4 nanoparticles under different annealing conditions, Mater. Chem. Phys., 2011, 128, 400–404 CrossRef CAS PubMed.
- S. G. Christoskova, M. Stoyanova and M. Georgieva, Low-temperature iron-modified cobalt oxide system part I. Preparation and characterization, Appl. Catal., A, 2001, 208, 235–242 CrossRef CAS.
- J. Fang, N. Shama, L. D. Tung, E. Y. Shin, C. J. O'Connor and K. L. Stokes, Ultrafine NiFe2O4 powder fabricated from reverse microemulsion process, J. Appl. Phys., 2003, 93, 7483–7485 CrossRef CAS PubMed.
- X. Li and C. Kutta, Synthesis and characterization of superparamagnetic CoxFe3−xO4 nanoparticles, J. Alloys Compd., 2003, 349, 264–268 CrossRef CAS.
- C. Yan, Z. Xu, F. Cheng, Z. Wang, L. Sum, C. Liao and J. Jia, Nanophased CoFe2O4 prepared by combustion method, Solid State Commun., 1999, 111, 287–291 CrossRef CAS.
- L. Cao, Q. Zhou, L. Gu, H. Shen, Q. Li, P. Lan and Y. Fang, Preparation and characterization of composite microspheres of nano zinc ferrite/poly(D,L-lactide-co-alanine), Trans. Nonferrous Met. Soc. China, 2012, 22, 360–365 CrossRef CAS.
- X. Li, W. Yang, F. Li, D. G. Evans and X. Duan, Stoichiometric synthesis of pure NiFe2O4 spinel from layered double hydroxide precursors for use as the anode material in lithium-ion batteries, J. Phys. Chem. Solids, 2006, 67, 1286–1290 CrossRef CAS PubMed.
- F. Li, J. Liu, D. G. Evans and X. Duan, Stoichiometric synthesis of pure MFe2O4 (M = Mg, Co and Ni) spinel ferrites from tailored layered double hydroxide (hydrotalcite-like) precursors, Chem. Mater., 2004, 16, 1597–1602 CrossRef CAS.
- G. N. Mahmoud, B. S. Elias, H. Mansor, H. S. Abdul and A. A. Hossein, Synthesis and characterization of zinc ferrite nanoparticles by a thermal treatment method, Solid State Commun., 2011, 151, 1031–1035 CrossRef PubMed.
- C. Pham-Huu, N. Keller, C. Estournès, G. Ehret, J. M. Grenèche and M. Ledoux, Microstructural investigation and magnetic properties of CoFe2O4 nanowires synthesized inside carbon nanotubes, Phys. Chem. Chem. Phys., 2003, 5, 3716–3723 RSC.
- Y. Shi, J. Ding and H. Yin, CoFe2O4 nanoparticles prepared by the mechanochemical method, J. Alloys Compd., 2000, 308, 290–295 CrossRef CAS.
- B. G. Veronica, S. P. Regino and J. T. F. Maria, Superparamagnetism and interparticle interactions in ZnFe2O4 nanocrystals, J. Mater. Chem., 2012, 22, 2992–3003 RSC.
- J. Jiang, Y. M. Yang and L. C. Li, Surfactant-assisted synthesis of nanostructured NiFe2O4 via a refluxing route, Mater. Lett., 2008, 62, 1973–1975 CrossRef CAS PubMed.
- S. Saadi, A. Bouguelia and M. Trari, Photoassisted hydrogen evolution over spinel CuM2O4 (M = Al, Cr, Mn, Fe and Co), Renewable Energy, 2006, 31, 2245–2256 CrossRef CAS PubMed.
- H. H. Yang, J. Yan, Z. Lu, X. Cheng and Y. Tang, Photocatalytic activity evaluation of tetragonal CuFe2O4 nanoparticles for the H2 evolution under visible light irradiation, J. Alloys Compd., 2009, 476, 715–719 CrossRef CAS PubMed.
- S. Li, W. C. Wang, Y. Q. Chen, L. J. Zhang, J. X. Guo and M. C. Gong, Fabrication and characterization of TiO2/BaAl2O4: Eu2+, Dy3+ and its photocatalytic performance towards oxidation of gaseous benzene, Catal. Commun., 2009, 10, 1048–1051 CrossRef CAS PubMed.
- J. Okal and M. Zawadzki, Catalytic combustion of methane over ruthenium supported on zinc aluminate spinel, Appl. Catal., A, 2013, 453, 349–357 CrossRef CAS PubMed.
- C. Ragupathi, J. J. Vijaya, S. Narayanan, L. J. Kennedy and S. Ramakrishna, Catalytic properties of nanosized zinc aluminates prepared by greenprocess using Opuntia dilenii haw plant extract, Chin. J. Catal., 2013, 34, 1951–1958 CrossRef CAS.
- M. R. Hoffmann, S. T. Martin, W. Choi and D. W. Bahnemann, Environmental Applications of Semiconductor Photocatalysis, Chem. Rev., 1995, 95, 69–96 CrossRef CAS.
- L. Chen, D. Ma, P. Barbara and X. Bao, Carbon-supported silver catalysts for CO selective oxidation in excess hydrogen, J. Nat. Gas Chem., 2006, 15, 181–190 CrossRef CAS.
- J. H. Park, J. Park and H. Shin, The preparation of Ag/mesoporous silica by direct silver reduction and Ag/functionalized mesoporous silica by in situ formation of adsorbed silver, Mater. Lett., 2007, 61, 156–159 CrossRef CAS PubMed.
- Y. Yu, H. He, Q. Feng, H. Gao and X. Yang, Mechanism of the selective catalytic reduction of NOx by C2H5OH over Ag/Al2O3, Appl. Catal., B, 2004, 49, 159–171 CrossRef CAS PubMed.
- C. Shi, M. Cheng, Z. Qu and X. Bao, Investigation on the catalytic roles of silver species in the selective catalytic reduction of NOx with methane, Appl. Catal., B, 2004, 51, 171–181 CrossRef CAS PubMed.
- L. Yang, X. Wang, Y. Fan and X. Yang, Excellent antimicrobial properties of mesoporous anatase TiO2 and Ag/TiO2 composite films, Microporous Mesoporous Mater., 2008, 114, 431–439 CrossRef PubMed.
- J. Ren, W. Z. Wang, S. M. Sun, L. Zhang and J. Chang, Enhanced photocatalytic activity of Bi2WO6 loaded with Ag nanoparticles under visible light irradiation, Appl. Catal., B, 2009, 92, 50–55 CrossRef CAS PubMed.
- D. Guin, B. Baruwati and S. V. Manorama, Pd on amine-terminated ferrite nanoparticles: a complete magnetically recoverable facile catalyst for hydrogenation reactions, Org.Lett., 2007, 9, 1419–1421 CrossRef CAS PubMed.
- B. Baruwati, V. Polshettiwar and R. S. Varma, Magnetically recoverable supported ruthenium catalyst for hydrogenation of alkynes and transfer hydrogenation of carbonyl compounds, Tetrahedron Lett., 2009, 50, 1215–1218 CrossRef CAS PubMed.
- D. S. Bae, K. S. Han and J. H. Adair, Synthesis and microstructure of Pd/SiO2 nanosized particles by reverse micelle and sol–gel processing, J. Mater. Chem., 2002, 12, 3117–3120 RSC.
- Z. Zhu, X. Li, Q. Zhao, Y. Shi, H. Li and G. Chen, Surface photovoltage properties and photocatalytic activities of nanocrystalline CoFe2O4 particles with porous superstructure fabricated by a modified chemical coprecipitation method, J. Nanopart. Res., 2011, 13, 2147–2155 CrossRef CAS.
- R. Jenkins and R. L. Snyder, Chemical Analysis: Introduction to X-ray Powder Diffractometry, Wiley, New York, 1996 Search PubMed.
- Z. Sun, L. Liu, D. Jia and W. Pan, Simple synthesis of CuFe2O4 nanoparticles as gas-sensing materials, Sens. Actuators, B, 2007, 125, 144–148 CrossRef CAS PubMed.
- Z. Zhu, X. Li, Q. Zhao, Y. Li, C. Sun and Y. Cao, Photocatalytic performances and activities of Ag-doped CuFe2O4 nanoparticles, Mater. Res. Bull., 2013, 48, 2927–2932 CrossRef CAS PubMed.
- W. Lv, B. Liu, Z. Luo, X. Ren and P. Zhang, XRD studies on the nanosized copper ferrite powders synthesized by sonochemical method, J. Alloys Compd., 2008, 465, 261–264 CrossRef CAS PubMed.
- N. Deraz and M. Hessien, Structural and magnetic properties of pure and doped nanocrystalline cadmium ferrite, J. Alloys Compd., 2009, 475, 832–839 CrossRef CAS PubMed.
- X. Li, G. E. Fryxell, C. Wang and M. H. Engelhard, The synthesis of Ag-doped mesoporous TiO2, Microporous Mesoporous Mater., 2008, 111, 639–642 CrossRef CAS PubMed.
- S. Ye, L. Fang and Y. Lu, Contribution of charge-transfer effect to surface-enhanced IR for Ag@PPy nanoparticles, Phys. Chem. Chem. Phys., 2009, 11, 2480–2484 RSC.
- Z. Szotek, W. Temmerman, D. Kodderitzsch, A. Svane, L. Petit and H. Winter, Electronic structures of normal and inverse spinel ferrites from first principles, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 74, 174431 CrossRef.
- N. M. Deraz and J. Anal, Production and characterization of pure and doped copper ferrite nanoparticles, J. Anal. Appl. Pyrolysis, 2008, 82, 212–222 CrossRef CAS PubMed.
- R. K. Selvan, V. Krishnan, C. O. Augustin, H. Bertagnolli, C. S. Kim and A. Gedanken, Investigations on the structural, morphological, electrical, and magnetic properties of CuFe2O4–NiO nanocomposites, Chem. Mater., 2008, 20, 429–439 CrossRef CAS.
- B. S. Randhawa, Preparation of ferrites from the thermolysis of transition metal ferrioxalate precursors, J. Mater. Chem., 2000, 10, 2847–2852 RSC.
- M. Wang, Z. Ai and L. Zhang, Generalized preparation of porous nanocrystalline ZnFe2O4 superstructures from zinc ferrioxalate precursor and its superparamagnetic property, J. Phys. Chem. C, 2008, 112, 13163–13170 CAS.
- S. M. Montemayor, L. A. G. Cerda and J. R. T. Lubian, Structural evolution during crystallization of β-BaB2O4 thin films fabricated by chemical solution deposition technique, Mater. Lett., 2005, 59, 1056–1061 CrossRef CAS PubMed.
- A. Setiabudi, M. Makkee and J. A. Moulijn, The role of NO2 and O2 in the accelerated combustion of soot in diesel exhaust gases, Appl. Catal., B, 2004, 50, 185–194 CrossRef CAS PubMed.
- J. O. Muller, D. S. Su, R. E. Jentoft, J. Krohnert, F. C. Jentoft and R. Schlogl, Morphology-controlled reactivity of carbonaceous materials towards oxidation, Catal. Today, 2005, 102, 259–265 CrossRef PubMed.
- Q. Zhai, S. Qiu, F. Xiao, Z. Zhang, C. Shao and Y. Han, Preparation, characterization, and optical properties of the host–guest nanocomposite material zeolite-silver iodide, Mater. Res. Bull., 2000, 35, 59–73 CrossRef CAS.
- V. Donchev, K. Kirilov, T. S. Ivanov and K. Germanova, Surface photovoltage phase spectroscopy – a handy tool for characterisation of bulk semiconductors and nanostructures, Mater. Sci. Eng., B, 2006, 129, 186–192 CrossRef CAS PubMed.
- K. Maeda, Z-Scheme Water Splitting Using Two Different Semiconductor Photocatalysts, ACS Catal., 2013, 3, 1486–1503 CrossRef CAS.
- M. Dahl, Y. Liu and Y. Yin, Composite Titanium Dioxide Nanomaterials, Chem. Rev., 2014, 114, 9853–9889 CrossRef CAS PubMed.
- X. B. Chen and S. S. Mao, Titanium dioxide nanomaterials: synthesis, properties, modifications, and applications, Chem. Rev., 2007, 107, 2891–2959 CrossRef CAS PubMed.
- T. Hayakawa, S. Selvan and M. Nogami, Field enhancement effect of small Ag particles on the fluorescence from Eu3+-doped SiO2 glass, Appl. Phys. Lett., 1999, 74, 1513–1515 CrossRef CAS PubMed.
- J. Chen, Z. Lu, Y. Wang, X. Chen and L. Zhang, Magnetically recyclable Ag/SiO2–CoFe2O4 nanocomposite as a highly active and reusable catalyst for H2 production, Int. J. Hydrogen Energy, 2015, 40, 4777–4785 CrossRef CAS PubMed.
- X. S. Li, G. E. Fryxell, C. Wang and M. Engelhard, The synthesis of Ag-doped mesoporous TiO2, Microporous Mesoporous Mater., 2008, 111, 639–642 CrossRef CAS PubMed.
- A. Chen, W. Zhang, Y. Liu, X. Han and X. Bao, In situ introduction of dispersed metallic Ag nanoparticles into the channels of mesoporous carbon CMK-3, Chin. Chem. Lett., 2007, 18, 1017–1020 CrossRef CAS PubMed.
- L. Chen, D. Ma, B. Pietruszka and X. Bao, Carbon-Supported Silver Catalysts for CO Selective Oxidation in Excess Hydrogen, J. Nat. Gas Chem., 2006, 15, 181–190 CrossRef CAS.
- M. Hoffmann, T. Martin, W. Choi and D. Bahnemann, Environmental applications of semiconductor photocatalysis, Chem. Rev., 1995, 95, 69–96 CrossRef CAS.
- X. Wang, J. Yu, H. Yip, L. Wu, P. Wong and S. Lai, A Mesoporous Pt/TiO2 Nanoarchitecture with Catalytic and Photocatalytic Functions, Chem.–Eur. J., 2005, 11, 2997–3004 CrossRef CAS PubMed.
- Z. Zhu, Q. Zhao, X. Li, Y. Li, C. Sun, G. Zhang and Y. Cao, Photocatalytic performances and activities of ZnAl2O4 nanorods loaded with Ag towards toluene, Chem. Eng. J., 2012, 203, 43–51 CrossRef CAS PubMed.
- G. Marci, M. Addamo, V. Augugliaro, S. Coluccia, E. García-López, V. Loddo, G. Martra, L. Palmisano and M. Schiavello, Photocatalytic oxidation of toluene on irradiated TiO2: comparison of degradation performance in humidified air, in water and in water containing a zwitterionic surfactant, J. Photochem. Photobiol., A, 2003, 160, 105–114 CrossRef CAS.
- H. Yu, X. Quan, S. Chen, H. Zhao and Y. Zhang, TiO2-carbon nanotube heterojunction arrays with a controllable thickness of TiO2 layer and their first application in photocatalysis, J. Photochem. Photobiol., A, 2008, 200, 301–306 CrossRef CAS PubMed.
- Y. Hou, X. Li, X. Zou, X. Quan and G. Chen, Photoeletrocatalytic activity of a Cu2O-loaded self-organized highly oriented TiO2 nanotube array electrode for 4-chlorophenol degradation, Environ. Sci. Technol., 2009, 43(3), 858–863 CrossRef CAS.
|
This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.