Next Article in Journal
Tunable Bandgap and Optical Properties of Black Phosphorene Nanotubes
Previous Article in Journal
Nanogels for Pharmaceutical and Biomedical Applications and Their Fabrication Using 3D Printing Technologies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Structural Property, Photophysical Property, Photocatalytic Property of Novel ZnBiErO4 under Visible Light Irradiation

1
School of Chemistry and Environmental Engineering, Changchun University of Science and Technology, Changchun 130600, China
2
State Key Laboratory of Pollution Control and Resource Reuse, School of the Environment, Nanjing University, Nanjing 210093, China
*
Author to whom correspondence should be addressed.
Materials 2018, 11(2), 303; https://doi.org/10.3390/ma11020303
Submission received: 29 December 2017 / Revised: 26 January 2018 / Accepted: 30 January 2018 / Published: 18 February 2018

Abstract

:
A novel photocatalyst ZnBiErO4 was firstly synthesized by solid-state reaction method and its structural and photocatalytic properties were analyzed by scanning electron microscopy, X-ray diffraction, X-ray photoelectron spectroscopy and UV-Vis diffuse reflectance. The results demonstrated that ZnBiErO4 crystallized with tetragonal crystal structure with space group I41/A. The lattice parameters for ZnBiErO4 were proved to be a = b = 10.255738 Å and c = 9.938888 Å. The band gap of ZnBiErO4 was estimated to be about 1.69 eV. Compared with nitrogen doped TiO2, ZnBiErO4 showed excellent photocatalytic activities for degrading methyl blue during visible light irradiation. The photocatalytic degradation of methyl blue with ZnBiErO4 or N-doped TiO2 as catalyst followed the first-order reaction kinetics. Moreover, the apparent first-order rate constant of ZnBiErO4 or N-doped TiO2 was 0.01607 min−1 or 0.00435 min−1. The reduction of total organic carbon, formation of inorganic products, such as SO42− and NO3 and the evolution of CO2 revealed the continuous mineralization of methyl blue during the photocatalytic process. ZnBiErO4 photocatalyst had great potential to purify textile industry wastewater.

1. Introduction

The organic pollutants of dye wastewater formatted by the industrial production, are toxicity, unacceptable color, high chemical oxygen demand and biological degradation, which caused the difficulty of self-degradation in natural conditions [1]. It will not only get undesirable result but also produce benzene, aniline and other secondary pollutants sometimes when we degrade it by traditional treatment methods [2,3,4,5,6,7]. In recent years, photocatalysis has caused general concern because of its strong oxidizability to degrade many organic pollutants in wastewater and effluents, especially for the dyes wastewater [8,9,10,11,12,13,14,15]. In recent years, as a semiconductor, TiO2 photocatalyst has attracted wide attention owing to the property of non-toxic, excellent stability and low cost [16,17,18]. However, due to the wide band gap (3.2 eV) of TiO2, the required wavelength is less than 387.5 nm, which near ultraviolet light and occupies only 4~5% of the solar energy and thus limits its application [19,20]. Although a variety of methods have been tried to modify TiO2, its quantum efficiency is still very low and the modification brings more shortcomings [21,22,23,24,25,26]. Based on the reason, it is necessary to develop new visible light responsive photocatalysts to utilize more visible light, which accounts for about 43% of the solar energy. Therefore, it is of grate important to develop the catalysts which can response to visible light. And many scientists have engaged in photocatalysis using visible light today [27,28,29,30,31,32,33,34,35].
AB2O4-type spinel composite oxides with narrow band gap have attracted extensive attention because of its excellent visible light response in textile wastewater treatment in recent years. They are capable to be used in adsorption, photocatalysis, hydrogen production and chemical sensors. It is demonstrated that NiCo2O4 and ZnFe2O4 exhibited a good photocatalytic activity for degrading methylene blue and other organic under visible light irradiation [36,37]. CoAl2O4 films were prepared in different structure, composition, morphology and optical properties by the method of chemical vapor deposition in 800–1200 °C by Carta et al. and its high photocatalytic activity were demonstrated [38]. The spinel oxide CaBi2O4 synthesized by Tang et al. could not only degrade gas phase acetaldehyde and MB aqueous solution under visible light irradiation but also keep high catalyst recovery [39]. ZnCo2O4 nanoparticles were synthesized by a co-precipitation decomposition method and the optical band gap of the nanoparticles was estimated to be 3.39 eV from the UV-Vis absorption spectrum. On the basis of the laboratory test results, the ZnCo2O4 nanoparticles performed a high photocatalytic activity for the degradation of methyl blue dye solution under visible light irradiation [40]. ZnBiGaO4 was prepared to photocatalytic decompose hydrogen sulfide (H2S) with two different methods under visible light irradiation, its band gap was estimated to be 2.8 eV [41]. MgFe2O4/PANI photocatalyst, which was synthesized by hydrothermal synthesis method, proved a high effective photodecolorization of both methylene blue and methyl orange dyes in aqueous solution and maintained a high removal ability after five repeated utilizations [42].
ZnBiErO4 is a material that has never been synthesized and the data about its structural and photocatalytic properties such as space group and lattice constants have not been found before. ZnBiErO4 seems to have a potential to improve the photocatalytic activity, because it has been proved that a slight modification of a semiconductor structure will result in a remarkable change in photocatalytic properties [43]. Herein, ZnBiErO4 was prepared by the solid-state reaction method for the first time. We studied its structure and photocatalytic properties through a series of characterizations. Photocatalytic performances of ZnBiErO4 in the degradation of methyl blue (MB) under visible light were also investigated. At the same time, we made a comparison between ZnBiErO4 and N-doped titanium dioxide to discover their differences in photocatalytic activity. The experimental results showed that the catalytic system with ZnBiErO4 revealed unusual catalytic effects for MB under visible light irradiation.

2. Results and Discussion

2.1. Characterization

Figure 1 is the scanning electron microscope-X-ray energy dispersion (SEM-EDS) spectrum of the as-synthesized ZnBiErO4. The composition was analyzed by EDS and the result was in line with that of XRD patterns. According to the SEM-EDS spectrum, the prepared ZnBiErO4 was composed of the elements of zinc, bismuth, Erbium and oxygen, furthermore, the mean atomic ratio of Zn:Bi:Er:O was estimated to be 1:1:1:4. Fortunately, other elements were not traced, which showed that our prepared ZnBiErO4 catalyst was pure phase without any other impurities. The SEM image shows the morphology of ZnBiErO4 particle and the irregular shape and uniform particle size of ZnBiErO4 were observed. Figure 2 shows a typical transmission electron microscopy (TEM) image of ZnBiErO4 catalyst, which reveals that the average particle diameter of ZnBiErO4 was about 350 nm. The corresponding selected-area electron diffraction (SAED) pattern is presented in Figure 3, showing discrete diffraction rings. It is suggested that the single sheet was comprised by multi crystals. Figure 4 is the diagram of particle size of ZnBiErO4 catalyst, which indicated that the particle size of ZnBiErO4 was mainly distributed in 326~387 nm. The result was in line with the TEM.
Figure 5 represents the X-ray powder diffraction patterns of ZnBiErO4, ZnO, Bi2O3 and Er2O3. Compared to the individual oxides used in synthesis, the synthesized photocatalyst was proved to be novel and it was not the mixture of individual oxides sintered at high temperature. Full-profile structure refinements of the collected X-ray diffraction data of ZnBiErO4 were continued to investigate by the RIETANTM program, which was based on Pawley analysis. The obtained XRD patterns of ZnBiErO4 samples are shown in Figure 6. It is demonstrated that the observed intensities were in according with the calculated intensities in a tetragonal crystal structure with space group I41/A. The lattice parameters for ZnBiErO4 were proved to be a = b = 10.255738 Å and c = 9.938888 Å. According to the lattice constants and the space group I41/A. All of the diffraction peaks for ZnBiErO4 could be successfully indexed and no other impurity phases are found, showing the single phase of the products. The result was in line with SEM-EDS spectrum. In order to simulate the structure of ZnBiErO4 accurately, Pawley refinements of the XRD spectrum and the three-dimensional space structure model of ZnBiErO4 photocatalyst were established by Material Studio software (Material Studio 2.0, Accelrys Corporation, USA). The atomic coordinates and structural parameters of ZnBiErO4 are listed in Table 1 and the atomic structure of ZnBiErO4 is shown in Figure 7.
In order to get the surface chemical compositions and the valence states of various elements of ZnBiErO4, the X-ray photoelectron spectrum (XPS) was carried out. The XPS full spectrum and detailed spectrum of ZnBiErO4 are shown in Figure 8. To see the results clearly, various elemental peaks with specific binding energies are provided in Table 2. According to the XPS full spectrum in Figure 8a, the synthesized ZnBiErO4 contained the elements of Zn, Bi, Er, O and C. The element of C was come from hydrocarbon for the sake of testing, so it was in agreement with the results of SEM-EDS. By fitting and separating the peak, we got the XPS detailed spectrum in Figure 8b–e. Zn2p1/2 and Zn2p3/2 peaks of Zn element were seen in Figure 8b, whose positions were 1043.6 eV and 1023.3 eV, respectively. Bi4f7/2 and Bi4f5/2 peaks of Bi element were observed in Figure 8c, whose positions were 157.7 eV and 163.0 eV, respectively. And the Er4d peak of Er element was presented in Figure 8d, whose position was 162.9 eV. It was obvious to find that there existed shoulders peaks at the XPS spectrum of O1s. By fitting and separating the peak, we got OLattice peak and OAbsorbed peak, which indicated that the catalyst absorbed some O2 in the process of calcination. Otherwise, the results showed that the oxidation states of Zn, Bi, Er and O ions from ZnBiErO4 were +2, +3, +3 and −2, respectively. Therefore, it could be deduced that the resulting material was highly pure under our preparation conditions. Notably, neither shoulders nor widening in the XPS peaks of ZnBiErO4 was observed, indicating the absence of any other phases.
The UV-vis absorption spectra of ZnBiErO4 and N-doped TiO2 are shown in Figure 9. In contrast to the N-doped TiO2, whose absorption edge was 414 nm, the novel synthesized ZnBiErO4 photocatalyst exhibited an obvious absorption in the visible region up to 690 nm. Furthermore, the band gap of ZnBiErO4 and N-doped TiO2 by extrapolating the linear portion of the curve to zero absorbance are 1.69 eV and 2.76 eV, respectively. It indicated that ZnBiErO4 could be excited by visible light and might exhibit a better visible light photocatalytic activity. Apparently, reflection and scattering did not cause a significant absorption.
For a crystalline semiconductor, the optical absorption near the band edge followed the equation [44]:
αhν = A(Eg)n
Here, A, α, Eg and ν are proportional constant, absorption coefficient, band gap and light frequency, respectively. Within this equation, n determines the character of the transition in a semiconductor. Eg and n can be calculated by the following steps: (i) plotting ln(αhν) versus ln(Eg) by assuming an approximate value of Eg (ii) deducing the value of n based on the slope in this graph. (iii) refining the value of Eg by plotting (αhν)1/n versus and extrapolating the plot to (αhν)1/n = 0. Figure 9a shows the plot of ln(αhν) versus ln(hν − Eg) for ZnBiErO4. Based on the Figure 9a, the slope of the graph was 1.49048961 and the value of n for ZnBiErO4 was calculated to be 2. Figure 9b shows the plot of (αhν)1/2 versus for ZnBiErO4. It was obvious to see that the value of Eg for ZnBiErO4 was calculated to be 1.69 eV, indicating that ZnBiErO4 possessed narrow band gap and the optical transition for ZnBiErO4 was directly allowed.

2.2. Photocatalytic Activity

Figure 10 (cited from [45]) shows the irradiance spectrum of the xenon lamp light. From the diagram, we can learn that the spectrum of xenon lamp is close to the visible light. Therefore, a 500 W xenon lamp was used as a light source with a 400 nm cutoff filter to provide visible light irradiation in this paper. Figure 11 shows temporal spectral changes of aqueous solutions of methyl blue (MB) due to visible light irradiation in the presence of ZnBiErO4. The result illustrated that ZnBiErO4 could photodegrade MB effectively under visible light irradiation and the color of MB solution shallowed gradually when irradiation time increased. Otherwise, it was also obvious to observe that there was a significant reduction in typical MB peaks in 591 nm from the results, demonstrating that the maximum absorption wavelength of MB was 591 nm. The result is close to the study by Cai Jiabia et al., in which the absorbance wavelength of methyl blue is 596 nm [46].
According to the principle of photocatalysis, electron-hole pairs within ZnBiErO4 resulted from the direct absorption of band-gap photons began to diffuse to the surface of ZnBiErO4, as a result, the photocatalytic activity of ZnBiErO4 for degradation of organic compounds might be enhanced. Figure 12 describes the photocatalytic degradation chart of MB under visible light irradiation in the presence of ZnBiErO4, N-doped TiO2 (N-TiO2) as well as in the absence of a photocatalyst. Based on the changes in the UV-Vis spectrum of decomposing MB in visible light (λ > 400 nm) by ZnBiErO4, N-doped TiO2, it could be seen that both ZnBiErO4 and N-doped TiO2 could photodegrade MB effectively under visible light irradiation. After visible light irradiation for 220 min, the removal rate of MB was estimated to be 96.99%, 62.39% or 3.31% with ZnBiErO4, N-doped TiO2 as catalyst, as well as in the absence of a photocatalyst, respectively. From the Figure 12, we can observe that the speed was slower of MB degradation by using ZnBiErO4 or N-doped TiO2 as a photocatalyst during the later reaction process compared with the speed of 0–60 min. According the results, we supposed that some intermediate products generated during MB degradation occupied active reaction sites on the photocatalytic surface, which could reduce the number of oxygen active radical and the reaction probability between the dye of MB and optical electron-hole pairs [47,48,49]. The photodegradation rate of MB was about 2.153 × 10−9 mol L−1 s−1 or 1.385 × 10−9 mol L−1 s−1 with ZnBiErO4 or N-doped TiO2 as catalyst under visible light irradiation for 220 min. The self-degradation rate of MB was 0.07348 × 10−9 mol L−1 s−1 without a catalyst. Furthermore, the photonic efficiency and the quantum yield was estimated to be 0.0452% (λ = 420 nm) and 0.6134% by using ZnBiErO4 as catalyst. Similarly, the photonic efficiency and the quantum yield was estimated to be 0.0291% (λ = 420 nm) and 0.3946% with N-doped TiO2 as catalyst. It could obviously tell that the photodegradation rate of MB, the photonic efficiency and quantum yield with ZnBiErO4 as catalyst were higher than those with N-doped TiO2 as catalyst, indicating that the photocatalytic degradation activity of ZnBiErO4 under visible light was higher than N-doped TiO2. One of the reason might be that a high photonic efficiency and quantum yield could lead to the production of a large number of electron-hole pairs, which were responsible for the photocatalytic degradation reaction directly or indirectly [50]. In addition, the band gap was another factor [51]. The band gap of ZnBiErO4 was calculated to be 1.69 eV and the band gap of N-doped TiO2 was 2.76 eV, demonstrated that ZnBiErO4 possessed narrower band gap compared with N-doped TiO2. Therefore, ZnBiErO4 could make use of more irradiation light energy than N-doped TiO2, which was more beneficial to generate separate photogenerated electrons and photogenerated holes. Subsequently, the ZnBiErO4 photocatalytic activities were improved. In conclusion, ZnBiErO4 possessed higher photocatalytic activity than N-doped TiO2.
The changes of total organic carbon (TOC) during photocatalytic degradation of MB in the presence of ZnBiErO4 and N-TiO2 as well as in the absence of a photocatalyst are shown in Figure 13, which was in accordance with the tendency shown in Figure 12. The measurements of TOC represented the change of organic carbon in MB solution which contained ZnBiErO4 catalyst, N-doped TiO2 catalyst or no catalyst under visible light irradiation. The gradual decrease of TOC represented the gradual disappearance of organic carbon and the reduce rate of TOC was estimated to be 94.87%, 59.42% or 3.09% after visible light irradiation for 220 min in the presence of ZnBiErO4 and N-TiO2 as well as in the absence of a photocatalyst. Moreover, the reaction stopped when the light was turned off in this experiment, indicating the obvious light response of ZnBiErO4 photocatalyst. It was illustrated that MB had been converted to other kinds of byproducts and the organic carbon in the MB had not been decomposed to CO2.
Figure 14 shows the first-order nature of the photocatalytic degradation kinetics with ZnBiErO4 or N-doped TiO2 as catalyst. A linear correlation between ln (C/Co) (or ln (TOC/TOCo)) and the irradiation time could be clearly seen for the photocatalytic degradation of MB by ZnBiErO4 or N-doped TiO2 under visible light irradiation. According to Figure 14, the first-order rate constant kC of MB concentration was estimated to be 0.01607 min−1 or 0.00435 min−1 with ZnBiErO4 or N-doped TiO2 as catalyst. Based on the result, it was more suitable for ZnBiErO4 to degrade MB under visible light irradiation compared with N-doped TiO2. Meanwhile, the first-order rate constant KTOC of TOC was estimated to be 0.01313 min−1 or 0.00419 min−1 with ZnBiErO4 or N-doped TiO2 as catalyst. The different value between kC and KTOC illustrated that the photodegradation intermediate products of MB probably appeared during the photocatalytic degradation of MB under visible light irradiation. At the same time, ZnBiErO4 showed a higher mineralization efficiency for MB degradation compared with N-doped TiO2.
Figure 15 reveals the amount variation of CO2 which was yielded during the photodegradation of MB with ZnBiErO4 or N-doped TiO2 as catalyst under visible light irradiation. The amount of CO2 increased gradually with increasing the irradiation time. After visible light irradiation for 220 min, the CO2 production of 0.30805 mmol with ZnBiErO4 as catalyst was higher than the CO2 production of 0.19189 mmol with N-doped TiO2 as catalyst. It could be distinctly seen from Figure 13 and Figure 15 that the amount of CO2 production was nearly equal to the amount of removed TOC and the quantity of CO2 production or the removed TOC was slightly lower than the quantity of reduced MB by using different catalysts. It was demonstrated that MB was mainly mineralized into some inorganic products including CO2 and H2O.
The nitrogen and sulfur atoms existed in MB were converted to some inorganic ions such as NO3 and SO42− as the end products. Figure 16 shows the concentration variation of NO3 ions and SO42− ions when MB was photodegraded by ZnBiErO4 under visible light irradiation. As the reaction time went by, it could be clearly seen that the concentration of NO3 and SO42− increased gradually during the photocatalytic degradation of MB with ZnBiErO4 as catalyst. After visible light irradiation for 220 min, the concentration of SO42− ions was 0.08445 mM with ZnBiErO4 as catalyst, indicating that 93.58% of sulfur from MB was converted into sulfate ions with ZnBiErO4 as catalyst after visible light irradiation for 220 min. Likewise, the concentration of NO3 ions was 0.08456 mM with ZnBiErO4 as catalyst during visible light irradiation for 220 min, indicating that 94.46% of nitrogen from MB was converted into nitrate ions with ZnBiErO4 as catalyst after visible light irradiation for 220 min. It was worth noting that the amount of SO42− ions released into the solution was lower than the value of stoichiometry. One possible reason might be that part of the sulfur was translated to sulfur-containing volatile compounds such as SO2. The second possible reason was that some SO42− ions were absorbed irreversibly on the surface of the photocatalyst, which had been observed by Lachheb et al. [52]. Regardless of the irreversible absorption of sulfate ions on the surface or not, it was important to stress that the evidence for restrained photocatalytic activity was not noticed. In contrast to the sulfate ions, the amount of nitrate ions at the end of the process was found to follow the expected stoichiometric ratio of three. The molecule of MB was converted to small organic species, which were subsequently mineralized into inorganic products such as SO42− ions, NO3 ions, CO2 and H2O ultimately.
Figure 17 describes the photocatalytic degradation chart of MB under visible light irradiation in the presence of ZnBiErO4, ZnO, Bi2O3, Er2O3 and mixture which contain ZnO, Bi2O3 and Er2O3. After visible light irradiation for 220 min, the removal rate of MB was estimated to be 96.99%, 59.35%, 57.44%, 54.30% or 58.53% with ZnBiErO4, ZnO, Bi2O3, Er2O3 or mixture as catalyst. It could be observed that the photodegradation rate of MB with ZnBiErO4 as catalyst was higher than those with ZnO, Bi2O3, Er2O3 or mixture as catalyst, indicating that the photocatalytic degradation activity of ZnBiErO4 under visible light was higher than those of ZnO, Bi2O3, Er2O3 or mixture. By investigating the photocatalytic degradation results of individual oxides and the synergistic effect of its mixture, we could learn that the photocatalytic activities were improved greatly when the new catalyst was formed, proving the justification of using ZnBiErO4 as the photocatalyst.
Figure 18 and Figure 19 represent the photocatalytic degradation chart of MB with ZnBiErO4 and commercial P25 as catalyst under visible light irradiation or sunlight irradiation. After visible light irradiation for 220 min, the removal rate of MB was estimated to be 96.99% or 12.01% with ZnBiErO4 or P25 as catalyst. After sunlight irradiation for 220 min, the removal rate of MB was estimated to be 97.85% or 24.51% with ZnBiErO4 or P25 as catalyst. It could be observed that the photodegradation rate of MB with ZnBiErO4 as catalyst was higher than those with P25 as catalyst whether under visible light irradiation or under sunlight irradiation, indicating that the photocatalytic degradation activity of ZnBiErO4 was higher than those of commercial P25.
Figure 20 shows the photocatalytic degradation of phenol with ZnBiErO4 as a photocatalyst under visible light irradiation. According to Figure 20, it was obvious to find that the photocatalytic activity was obtained when colorless phenol was selected as a contaminant model with ZnBiErO4 as photocatalyst. The photocatalytic degradation efficiency of phenol was estimated to be 98.80% when ZnBiErO4 was used as a photocatalyst under visible light irradiation after 220 min, demonstrating that ZnBiErO4 itself had photocatalytic activity and the photosensitive effect was not the main reason in the photodegradation process of MB by using ZnBiErO4 as a photocatalyst [53].

2.3. Photocatalytic Degradation Mechanism

As well as we have known, the organic pollutants can be degraded by the photogenerated reactive species during the photocatalytic reaction, including electrons (e), holes (h+), hydroxyl radicals (•OH) and superoxide radicals (•O2). In order to confirm the major reactive species for inducing the degradation of methyl blue with ZnBiErO4 as catalyst under visible light irradiation, radical scavenger experiments were performed by adding various scavengers into the system.
Superoxide dismutase (SOD) with the concentration of 66.7 mg L−1, sodium oxalate with the concentration of 10 mM, H2O2 with the concentration of 10 mM and tert-butyl alcohol (TBA) with the concentration of 10 mM were added into the system as scavengers to capture •O2, h+, e and •OH, respectively. Figure 21 illustrates photodegradation of methyl blue over ZnBiErO4 in the presence of various scavengers. As shown in Figure 21, after the addition of H2O2 as a scavenger for photogenerated electrons (e), the photodegradation efficiency of methyl blue remained almost the same compared with the system of no scavenger, demonstrating the minor role of e in this process. When tert-butyl alcohol (TBA) was used as the scavenger to capture •OH, the photodegradation activity declined somewhat. However, a significant reduction in the photocatalytic performance was observed in the presence of sodium oxalate and superoxide dismutase (SOD), which captured •O2 and h+ respectively. The results indicate that the photoreaction process in this system is dominated by h+ and •O2. In conclusion, h+ contribute most to the high activity of the ZnBiErO4 during the degradation of methyl blue. Additionally, •O2 radicals have a secondary importance during methyl blue degradation. As for •OH, it may participate in the photodegradation process. However, e exhibit almost no activity during methyl blue degradation in the presence of ZnBiErO4.
Based on above results, a possible mechanism of photocatalytic activity enhancement could be proposed as the following steps:
Scheme I:
ZnBiErO 4 Visible   light h + + e
e + O 2 · O 2
h + + OH · OH
h + + · OH + · O 2 + MB degradation products
Scheme II:
MB ( ads ) Visible   light MB ( ads ) *
MB ( ads ) * + ZnBiErO 4 ZnBiErO 4 ( e ) + MB ( ads ) +
ZnBiErO 4 ( e ) + O 2 ZnBiErO 4 + · O 2
· O 2 + MB degradation   products
Towards scheme I, the electron-hole pairs were generated on the ZnBiErO4 surface upon visible-light excitation (Reaction (R1)), the dissolved oxygen could be activated by the photogenerated electrons to produce superoxide anion radicals (Reaction (R2)), while the holes could react with the absorbed water to form hydroxyl radicals (Reaction (R3)), lastly, the contaminants absorbed on the active sites of ZnBiErO4 was oxidized by the active species (holes, superoxide anion radicals and hydroxyl radicals) (Reaction (R4)). The significant enhancement in photoactivity under visible light irradiation could be ascribed to the reduced recombination of photogenerated electrons and photogenerated holes by ZnBiErO4 [54].
As for the scheme II, MB adsorbed on the surface ZnBiErO4 was excited by visible light irradiation (Reaction (R5)). Subsequently, the electrons were injected from the excited MB to the conduction band of ZnBiErO4 (Reaction (R6)), then, the dissolved oxygen could scavenge the electron to produce superoxide anion radicals (Reaction (R7)), finally, the MB contaminants were degraded by the active species of superoxide anion radicals (Reaction (R8)).
Band gap energy indicated the minimum energy needed to photocatalytically degrade MB. Figure 22 shows the band structure of ZnBiErO4. The positions and width of the conduction band (CB) and the valence band (VB) were obtained after the calculation of the electronic band structure of ZnBiErO4 by the plane-wave-based density functional method. The band structure calculations of ZnBiErO4 were carried out with the program of Cambridge serial total energy package (CASTEP) and first principles simulation. According to above method, we deduced that the band gap energy of ZnBiErO4 was 1.69 eV, at the same time, the valence band of ZnBiErO4 was composed of O 2p and Bi 6s orbital components and the conduction band of ZnBiErO4 was composed of Zn 4s and Er 4d orbital components. The results indicated that electron-hole pairs could be produced when ZnBiErO4 absorbed enough photons which were more than needed photons.

3. Experimental

3.1. Synthesis of ZnBiErO4 and N-Doped TiO2

The novel ZnBiErO4 photocatalyst was synthesized by a solid-state reaction method. In the first place, ZnO, Bi2O3 and Er2O3 with purity of 99.99% (Sinopharm Group Chemical Reagent Co., Ltd., Shanghai, China) were mixed with 2:1.1:1 molar ration to serve as raw materials. All chemicals were dried at 200 °C for 4 h before the preparation of the catalyst. Then the precursors were fully mingled with each other, pressed into small columns and put into an alumina crucible (Shenyang Crucible Co., Ltd., Shenyang, China). Finally, calcination was carried out at 1000 °C for 35 h in an electric furnace (KSL 1700X, Hefei Kejing Materials Technology Co., Ltd., Hefei, China). After sintering and grounding within a quartz mortar, ultrafine ZnBiErO4 powder was fabricated. In this study, we prepared about 5 g ZnBiErO4 for characterization and experimental and its costs were about four dollars.
Nitrogen-doped titania (N-doped TiO2) catalyst was prepared by sol-gel method with tetrabutyl titanate (Shanghai Lingfeng Chemical Reagent Co. Ltd., Shanghai, China) as a precursor and ethanol (Sinopharm Group Chemical Reagent Co., Ltd., Shanghai, China) as solvent. The procedure was as follows: Firstly, 17 mL tetrabutyl titanate and 40 mL absolute ethyl alcohol were combined to serve as to solution A; 40 mL absolute ethyl alcohol, 10 mL glacial acetic acid (Sinopharm Group Chemical Reagent Co., Ltd., Shanghai, China) and 5 mL double distilled water were blended to be solution B; subsequently solution A was added dropwise into solution under vigorous stirring, then a transparent colloidal suspension was formed. Subsequently, aqua ammonia (Shanghai Lingfeng Chemical Reagent Co. Ltd., Shanghai, China), which N/Ti proportion was 8 mol %, was put into the resulting transparent colloidal suspension in the condition of stirring for 1 h. Then the xerogel was generated after being aged for 2 days. The xerogel was grounded into powder which was calcined at 500 °C for 2 h, subsequently above powder was grounded in agate mortar and screened by shaker to obtain N-doped TiO2 powders.

3.2. Characterization of ZnBiErO4

The crystal structures of ZnBiErO4 was determined by X-ray diffraction (XRD, D/MAX-RB, Rigaku Corporation, Tokyo, Japan) method, which operated at 40 kV and 40 mA with CuKα radiation (λ = 1.54056 Å). The data were collected at 295 K with a step-scan procedure, whose scan range was 2θ = 5–100°. The step interval was 0.02° and the time per step was 1 s. The chemical composition and the surface structure of the compound was determined by means of scanning electron microscope-X-ray energy dispersion spectrum (SEM-EDS, LEO 1530VP, LEO Corporation, Dresden, Germany). Transmission electron microscopy (TEM, Tecnal F20 S-Twin, FEI Corporation, Hillsboro, OR, USA) was used to observe the surface state and structure of the photocatalysts. The particle sizes of the photocatalysts were measured by Malvern’s mastersize-2000 particle size analyzer (Malvern Instruments Ltd., Malvern, UK). The content of Zn2+, Bi3+, Er3+ and O2− of ZnBiErO4 and the valence state of elements were analyzed by X-ray photoelectron spectroscopy (XPS, ESCALABMK-2, VG Scientific Ltd., London, UK). At the same time, X-ray photoelectron spectroscopy also could be utilized to examine the chemical composition within the depth profile of ZnBiErO4 using the method of argon ion denudation. The photophysical property of ZnBiErO4 was analyzed with an UV-2550 UV-visible spectrophotometer (Shimadzu, UV-2550, Shimadzu Corporation, Kyoto, Japan) in a UV-Vis diffuse reflectance experiment by the dry-pressed disk samples and BaSO4 was used as the reference material.

3.3. Photocatalytic Activity Tests

The photocatalytic activity of ZnBiErO4 was evaluated by photocatalytic degradation of methyl blue (C37H27N3O9S3) (Tianjin Bodi Chemical Co., Ltd., Tianjin, China) under visible light. The photoreaction was carried out in a photochemical reaction apparatus (Xujiang Electromechanical Plant, Nanjing, China) with high pressure xenon lamp and Cooling System. The process of our experiments was as follows: 0.3 g photocatalyst powder of N-doped TiO2 or ZnBiErO4 was added into 300 mL MB aqueous solution in every quartz tube, whose initial concentration was 0.0293 mM and initial pH value was 7.0. In order to ensure the establishment of an adsorption/desorption equilibrium among photocatalysts, the MB dye and atmospheric oxygen, every solution was magnetically stirred in the dark for 45 min. In this paper, a 500 W xenon lamp was used as a light source with a 400 nm cutoff filter to provide visible light irradiation. During visible light illumination, the MB dye pollution was stirred by a magnetic stirrer with 500 rpm and the photocatalyst powder was kept suspended in the solution. One of the quartz tubes was taken out from the photochemical reaction apparatus at various time intervals. Subsequently the suspension was filtered through 0.22 µm membrane filters and the filtrate was analyzed by a Shimadzu UV-2450 UV-Visible spectrometer with the detecting wavelength at 591 nm. Above experiments were performed under oxygen-saturation conditions ([O2] sat = 1.02 × 10−3 M), pH adjustment was not carried out and the initial pH value was 7.0 and a near constant reaction temperature (20 °C) was maintained by recycling water.
The identification of MB and the intermediate products of MB were measured by liquid chromatograph-mass spectrometer (LC-MS, Thermo Quest LCQ Duo, Thermo Fisher Scientific Corporation, Silicon Valley, CA, USA. Beta Basic-C18 HPLC column: 150 × 2.1 mm, ID of 5 μm, Thermo Fisher Scientific Corporation, Silicon Valley, CA, USA). There were post-photocatalysis solution (20 μL) in the LC-MS system, which were injected automatically. The fluent was composed of 60% methanol and 40% water and the flow rate was 0.2 mL·min−1. MS conditions consisted of an electrospray ionization interface, a capillary temperature of 27 °C with a voltage of 19.00 V, a spray voltage of 5000 V and a constant sheath gas flow rate. The negative ion scan mode and the m/z range swept from 50 to 600 in the spectrum were acquired. The concentration of total organic carbon (TOC) was determined with a TOC analyzer (TOC-5000, Shimadzu Corporation, Kyoto, Japan). Evolution of CO2 was measured with an intersmatTM IGC120-MB gas chromatograph (Thermo Separation Products Corporation, Brussels, Belgium). The device equipped with a porapack Q column (3 m in length and with an inner diameter of 0.2 in), which was connected to a catharometer detector. The obtained inorganic products from MB degradation were analyzed by ion chromatograph (DX-300, Dionex Corporation, Sunnyvale, CA, USA). The photonic efficiency was calculated according to the following equation [55,56]:
ξ = R/I0
where ξ is the photonic efficiency (%), R is the rate of MB degradation (Mol L−1 s−1) and I0 is the incident photon flux (Einstein·L−1 s−1). The incident photon flux I0 measured by a radiometer (Model FZ-A, Photoelectric Instrument Factory Beijing Normal University, Beijing, China) was determined to be 4.76 × 10−6 Einstein·L−1·s−1 under visible light irradiation.

4. Conclusions

In our study, ZnBiErO4 was prepared by a solid-state reaction method for the first time. The structural properties of ZnBiErO4 was characterized by some material characterization methods and the photocatalytic properties of ZnBiErO4 was also verified in comparison with N-doped TiO2. XRD results indicated a good agreement between the observed and calculated intensities in a tetragonal crystal structure with space group I41/A. The latticeparameters for ZnBiErO4 were proved to be a = b = 10.255738 Å and c = 9.938888 Å. XPS results confirmed that the synthesized ZnBiErO4 was highly pure under our preparation conditions. The band gap of ZnBiErO4 was estimated to be about 1.69 eV, indicating that ZnBiErO4 could absorb visible light (λ > 400 nm). Photocatalytic decomposition of aqueous MB was realized under visible light irradiation in the presence of ZnBiErO4 and N-doped TiO2. The results could obviously tell that the photodegradation rate of MB and the photonic efficiency with ZnBiErO4 as catalyst was higher than those with N-doped TiO2, which illustrated that ZnBiErO4 exhibited higher photocatalytic activities for MB degradation under visible light irradiation compared with N-doped TiO2. The photocatalytic degradation of MB with ZnBiErO4 or N-doped TiO2 as catalyst followed the first-order reaction kinetics. The apparent first-order rate constant of ZnBiErO4 or N-doped TiO2 was 0.01607 min−1 or 0.00435 min−1. The reduction of total organic carbon, formation of inorganic products such as SO42− and NO3 and the evolution of CO2 revealed the continuous mineralization of MB during the photocatalytic process. In addition, the possible photocatalytic degradation pathway of MB was deduced under visible light irradiation. Otherwise, the costs (about 0.8 dollars per gram) of ZnBiErO4 photocatalyst are low and the prepared materials are available, proving the availability and feasibility for a large scale application. In conclusion, the results indicated that ZnBiErO4/visible light photocatalysis system might be regarded as a practical method for the treatment of dye wastewater. As this system did not need high pressure of oxygen, heating or any chemical reagents, it has a promising application in purifying and reusing colored aqueous effluents.

Acknowledgments

This work was supported by a grant from China-Israel Joint Research Program in Water Technology and Renewable Energy (No. 5).

Author Contributions

Jingfei Luan was involved with all aspects of the study including conceiving, designing, data interpretation and writing the manuscript. Yan Zhuang performed the experiments and analyzed data. Jingfei Luan and Yan Zhuang wrote the paper. All authors read and approved the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Han, F.; Kambala, V.S.R.; Srinivasan, M.; Rajarathnam, D.; Naidu, R. Tailored titanium dioxide photocatalysts for the degradation of organic dyes in wastewater treatment: A review. Appl. Catal. A-Gen. 2009, 359, 25–40. [Google Scholar] [CrossRef]
  2. Gupta, V.K.; Suhas. Application of low-cost adsorbents for dye removal—A review. J. Environ. Manag. 2009, 90, 2313–2342. [Google Scholar] [CrossRef] [PubMed]
  3. Martinez-Huitle, C.A.; Brillas, E. Decontamination of wastewaters containing synthetic organic dyes by electrochemical methods: A general review. Appl. Catal. B 2009, 87, 105–145. [Google Scholar] [CrossRef]
  4. Nasuha, N.; Hameed, B.H. Adsorption of methylene blue from aqueous solution onto NaOH-modified rejected tea. Chem. Eng. J. 2011, 166, 783–786. [Google Scholar] [CrossRef]
  5. Safarik, I.; Horska, K.; Safarikova, M. Magnetically modified spent grain for dye removal. J. Cereal Sci. 2011, 53, 78–80. [Google Scholar] [CrossRef]
  6. Khan, R.; Bhawana, P.; Fulekar, M.H. Microbial decolorization and degradation of synthetic dyes: A review. Rev. Environ. Sci. Bio/Technol. 2013, 12, 75–97. [Google Scholar] [CrossRef]
  7. Popli, S.; Patel, U.D. Destruction of azo dyes by anaerobic-aerobic sequential biological treatment: A review. Int. J. Environ. Sci. Technol. 2015, 12, 405–420. [Google Scholar] [CrossRef]
  8. Khanna, A.; Shetty, V.K. Solar photocatalysis for treatment of Acid Yellow-17 (AY-17) dye contaminated water using Ag@TiO2 core-shell structured nanoparticles. Environ. Sci. Pollut. Res. 2013, 20, 5692–5707. [Google Scholar] [CrossRef] [PubMed]
  9. Qin, X.; Cheng, H.; Wang, W.; Huang, B.; Zhang, X.; Dai, Y. Three dimensional BiOX (X = Cl, Br and I) hierarchical architectures: Facile ionic liquid-assisted solvothermal synthesis and photocatalysis towards organic dye degradation. Mater. Lett. 2013, 100, 285–288. [Google Scholar] [CrossRef]
  10. Chang, Y.; Yu, K.; Zhang, C.X.; Li, R.; Zhao, P.Y.; Lou, L.L.; Liu, S.X. Three-dimensionally ordered macroporous WO3 supported Ag3PO4 with enhanced photocatalytic activity and durability. Appl. Catal. B 2015, 176, 363–373. [Google Scholar] [CrossRef]
  11. Zhu, Y.P.; Li, M.; Liu, Y.L.; Ren, T.Z.; Yuan, Z.Y. Carbon-doped ZnO hybridized homogeneously with graphitic carbon nitride nanocomposites for photocatalysis. J. Phys. Chem. C 2014, 118, 10963–10971. [Google Scholar] [CrossRef]
  12. Abdullah, H.; Kuo, D.H.; Kuo, Y.R.; Yu, F.A.; Cheng, K.B. Facile synthesis and recyclability of thin nylon film-supported n-type ZnO/p-type Ag2O nano composite for visible light photocatalytic degradation of organic dye. J. Phys. Chem. C 2016, 120, 7144–7154. [Google Scholar] [CrossRef]
  13. Teng, F.; Liu, Z.L.; Zhang, A.; Li, M. Photocatalytic performances of Ag3PO4 polypods for degradation of dye pollutant under natural indoor weak light irradiation. Environ. Sci. Technol. 2015, 49, 9489–9494. [Google Scholar] [CrossRef] [PubMed]
  14. Shao, D.L.; Gao, J.; Xin, G.Q.; Wang, Y.P.; Li, L.; Shi, J.; Lian, J.; Koratkar, N.; Sawyer, S. Cl-doped ZnO nanowire arrays on 3D graphene foam with highly efficient field emission and photocatalytic properties. Small 2015, 11, 4785–4792. [Google Scholar] [CrossRef] [PubMed]
  15. Ao, Y.H.; Wang, K.D.; Wang, P.F.; Wang, C.; Hou, J. Synthesis of novel 2D-2D p-n heterojunction BiOBr/La2Ti2O7 composite photocatalyst with enhanced photocatalytic performance under both UV and visible light irradiation. Appl. Catal. B 2016, 194, 157–168. [Google Scholar] [CrossRef]
  16. Sathyaseelan, B.; Manikandan, E.; Lakshmanan, V.; Baskaran, I.; Sivakumar, K.; Ladchumananandasivam, R.; Kennedy, J.; Maaza, M. Structural, optical and morphological properties of post-growth calcined TiO2 nanopowder for opto-electronic device application: Ex-situ studies. J. Alloys Compd. 2016, 671, 486–492. [Google Scholar] [CrossRef]
  17. Yoshinaga, M.; Yamamoto, K.; Sato, N.; Aoki, K.; Morikawa, T.; Muramatsu, A. Remarkably enhanced photocatalytic activity by nickel nanoparticle deposition on sulfur-doped titanium dioxide thin film. Appl. Catal. B 2009, 87, 239–244. [Google Scholar] [CrossRef]
  18. Erjavec, B.; Hudoklin, P.; Perc, K.; Tisler, T.; Dolenc, M.S.; Pintar, A. Glass fiber-supported TiO2 photocatalyst: Efficient mineralization and removal of toxicity/estrogenicity of bisphenol A and its analogs. Appl. Catal. B 2016, 183, 149–158. [Google Scholar] [CrossRef]
  19. Yao, B.H.; Peng, C.; Zhang, W.; Zhang, Q.K.; Niu, J.F.; Zhao, J.E. A novel Fe(III) porphyrin-conjugated TiO2 visible-light photocatalyst. Appl. Catal. B 2015, 174, 77–84. [Google Scholar] [CrossRef]
  20. Yaghoubi, H.; Li, Z.; Chen, Y.; Ngo, H.T.; Bhethanabotla, V.R.; Joseph, B.; Ma, S.Q.; Schlaf, R.; Takshi, A. Toward a visible light-driven photocatalyst: The effect of midgap-states-induced energy gap of undoped TiO2 nanoparticles. ACS Catal. 2015, 5, 327–335. [Google Scholar] [CrossRef]
  21. Lu, F.; Meng, F. Research evolution of doping modification on TiO2 photocatalyst. B. Chin. Ceram. Soc. 2011, 30, 116–119. [Google Scholar]
  22. Park, H.; Park, Y.; Kim, W.; Choi, W. Surface modification of TiO2 photocatalyst for environmental applications. J. Photochem. Photobiol. C 2013, 15, 1–20. [Google Scholar] [CrossRef]
  23. Sheng, G.; Li, J.; Wang, S.; Wang, X. Modification to promote visible-light catalytic activity of TiO2. Prog. Chem. 2009, 21, 2492–2504. [Google Scholar]
  24. Ding, D.W.; Liu, K.; He, S.N.; Gao, C.B.; Yin, Y.D. Ligand-Exchange Assisted Formation of Au/TiO2 Schottky Contact for Visible-Light Photocatalysis. Nano Lett. 2014, 14, 6731–6736. [Google Scholar] [CrossRef] [PubMed]
  25. Pastrana-Martinez, L.M.; Morales-Torres, S.; Figueiredo, J.L.; Faria, J.L.; Silva, A.M.T. Graphene oxide based ultrafiltration membranes for photocatalytic degradation of organic pollutants in salty water. Water Res. 2015, 77, 179–190. [Google Scholar] [CrossRef] [PubMed]
  26. Rengifo-Herrera, J.A.; Blanco, M.; Wist, J.; Florian, P.; Pizzio, L.R. TiO2 modified with polyoxotungstates should induce visible-light absorption and high photocatalytic activity through the formation of surface complexes. Appl. Catal. B 2016, 189, 99–109. [Google Scholar] [CrossRef]
  27. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-light photocatalysis in nitrogen-doped titanium oxides. Science 2001, 293, 269–271. [Google Scholar] [CrossRef] [PubMed]
  28. Mohapatra, P.; Parida, K.M. Photocatalytic activity of sulfate modified titania 3: Decolorization of methylene blue in aqueous solution. J. Mol. Catal. A-Chem. 2006, 258, 118–123. [Google Scholar] [CrossRef]
  29. Jang, Y.J.; Simer, C.; Ohm, T. Comparison of zinc oxide nanoparticles and its nano-crystalline particles on the photocatalytic degradation of methylene blue. Mater. Res. Bull. 2006, 41, 67–77. [Google Scholar] [CrossRef]
  30. Li, H.P.; Liu, J.Y.; Hou, W.G.; Du, N.; Zhang, R.J.; Tao, X.T. Synthesis and characterization of g-C3N4/Bi2MoO6 heterojunctions with enhanced visible light photocatalytic activity. Appl. Catal. B 2014, 160, 89–97. [Google Scholar] [CrossRef]
  31. Saison, T.; Chemin, N.; Chaneac, C.; Durupthy, O.; Mariey, L.; Mauge, F.; Brezova, V.; Jolivet, J.P. New Insights into BiVO4 Properties as Visible Light Photocatalyst. J. Phys. Chem. C 2015, 119, 12967–12977. [Google Scholar] [CrossRef]
  32. Ding, X.; Zhao, K.; Zhang, L.Z. Enhanced photocatalytic removal of sodium pentachlorophenate with self-doped Bi2WO6 under visible light by generating more superoxide ions. Environ. Sci. Technol. 2014, 48, 5823–5831. [Google Scholar] [CrossRef] [PubMed]
  33. Yang, S.B.; Xu, D.B.; Chen, B.Y.; Luo, B.F.; Yan, X.; Xiao, L.S.; Shi, W.D. Synthesis and visible-light-driven photocatalytic activity of p-n heterojunction Ag2O/NaTaO3 nanocubes. Appl. Surf. Sci. 2016, 383, 214–221. [Google Scholar] [CrossRef]
  34. Wu, L.N.; Fang, S.; Ge, L.; Han, C.C.; Qiu, P.; Xin, Y.J. Facile synthesis of Ag@CeO2 core-shell plasmonic photocatalysts with enhanced visible-light photocatalytic performance. J. Hazard. Mater. 2015, 300, 93–103. [Google Scholar] [CrossRef] [PubMed]
  35. Christoforidis, K.C.; Montini, T.; Bontempi, E.; Zafeiratos, S.; Jaen, J.J.D.; Fornasiero, P. Synthesis and photocatalytic application of visible-light active beta-Fe2O3/g-C3N4 hybrid nanocomposites. Appl. Catal. B 2016, 187, 171–180. [Google Scholar] [CrossRef]
  36. Cui, B.; Lin, H.; Li, Y.Z.; Li, J.B.; Sun, P.; Zhao, X.C. Photophysical and photocatalytic properties of core-ring struc-tured NiCo2O4 nanoplatelets. J. Phys. Chem. C 2009, 32, 14083–14087. [Google Scholar] [CrossRef]
  37. Chen, C.H.; Liang, Y.H.; Zhang, W.D. ZnFe2O4/MWCNTs composite with enhanced photocatalytic activity under visible-light irradiation. J. Alloys Compd. 2010, 1, 168–172. [Google Scholar] [CrossRef]
  38. Carta, G.; Casarin, M.; El Habra, N.; Natali, M.; Rossetto, G.; Sada, C.; Tondello, E.; Zanella, P. MOCVD deposition of CoAl2O4 films. Electrochim. Acta 2005, 50, 4592–4599. [Google Scholar] [CrossRef]
  39. Tang, J.W.; Zou, Z.G.; Ye, J.H. Efficient photocatalytic decomposition of organic contaminants over CaBi2O4 under visible-light irradiation. Angew. Chem.-Int. Ed. 2004, 43, 4463–4466. [Google Scholar] [CrossRef] [PubMed]
  40. Cui, B.; Lin, H.; Zhao, X.C.; Li, J.B.; Li, W.D. Visible light induced photocatalytic activity of ZnCo2O4 nanoparticles. Acta Phys.-Chim. Sin. 2011, 27, 2411–2415. [Google Scholar]
  41. Kale, B.B.; Baeg, J.O.; Kong, K.J.; Moon, S.J.; Lee, S.M.; So, W.W. Synthesis and structural analysis of visible light photocatalyst, ZnBiGaO4 for photocatalytic solar hydrogen production. Int. J. Energy Res. 2010, 34, 404–411. [Google Scholar] [CrossRef]
  42. Arshadnia, I.; Movahedi, M.; Rasouli, N. MgFe2O4 and MgFe2O4/ZnFe2O4 coated with polyaniline as a magnetically separable photocatalyst for removal of a two dye mixture in aqueous solution. Res. Chem. Intermed. 2017, 43, 4459–4474. [Google Scholar] [CrossRef]
  43. Tang, J.W.; Zou, Z.G.; Yin, J.; Ye, J. Photocatalytic degradation of methylene blue on Caln(2)O(4) under visible light irradiation. Chem. Phys. Lett. 2003, 382, 175–179. [Google Scholar] [CrossRef]
  44. Tauc, J.; Grigorov, R.; Vancu, A. Optical properties and electronic structure of amorphous germanium. Phys. Status Solidi 1966, 15, 627–637. [Google Scholar] [CrossRef]
  45. Xujiang Electromechanical Plant, Nanjing, China. Available online: http://www.njxu.com (accessed on 12 February 2018).
  46. Cai, J.B.; Wu, X.Q.; Li, Y.H.; Lin, Y.; Yang, H.; Li, S.X. Noble metal sandwich-like TiO2@Pt@C3N4 hollow spheres enhance photocatalytic performance. J. Colloid Interface Sci. 2018, 514, 791–800. [Google Scholar] [CrossRef] [PubMed]
  47. Vaiano, V.; Sacco, O.; Sannino, D.; Ciambelli, P. Nanostructured N-doped TiO2 coated on glass spheres for the photocatalytic removal of organic dyes under UV or visible light irradiation. Appl. Catal. B 2015, 170, 153–161. [Google Scholar] [CrossRef]
  48. Klein, M.; Nadolna, J.; Golabiewska, A.; Mazierski, P.; Klimczuk, T.; Remita, H.; Zaleska-Medynska, A. The effect of metal cluster deposition route on structure and photocatalytic activity of mono- and bimetallic nanoparticles supported on TiO2 by radiolytic method. Appl. Surf. Sci. 2016, 378, 37–48. [Google Scholar] [CrossRef]
  49. Jantawasu, P.; Sreethawong, T.; Chavadej, S. Photocatalytic activity of nanocrystalline mesoporous-assembled TiO2 photocatalyst for degradation of methyl orange monoazo dye in aqueous wastewater. Chem. Eng. J. 2009, 155, 223–233. [Google Scholar] [CrossRef]
  50. Guesh, K.; Marquez-Alvarez, C.; Chebude, Y.; Diaz, I. Enhanced photocatalytic activity of supported TiO2 by selective surface modification of zeolite Y. Appl. Surf. Sci. 2016, 378, 473–478. [Google Scholar] [CrossRef]
  51. Abdelhaleem, A.; Chu, W. Photodegradation of 4-chlorophenoxyacetic acid under visible LED activated N-doped TiO2 and the mechanism of stepwise rate increment of the reused catalyst. J. Hazard. Mater. 2017, 338, 491–501. [Google Scholar] [CrossRef] [PubMed]
  52. Lachheb, H.; Puzenat, E.; Houas, A.; Ksibi, M.; Elaloui, E.; Guillard, C.; Herrmann, J.M. Photocatalytic degradation of various types of dyes (Alizarin S, Crocein Orange G, Methyl Red, Congo Red, Methylene Blue) in water by UV-irradiated titania. Appl. Catal. B 2002, 39, 75–90. [Google Scholar] [CrossRef]
  53. Calza, P.; Rigo, L.; Sangermano, M. Investigations of photocatalytic activities of photosensitive semiconductors dispersed into epoxy matrix. Appl. Catal. B Environ. 2011, 106, 657–663. [Google Scholar] [CrossRef]
  54. Nasr, C.; Vinodgopal, K.; Fisher, L.; Hotchandani, S.; Chattopadhyay, A.K.; Kamat, P.V. Environmental photochemistry on semiconductor surfaces. Visible light induced degradation of a textile diazo dye, naphthol blue black, on TiO2 nanoparticles. J. Phys. Chem. 1996, 100, 8436–8442. [Google Scholar] [CrossRef]
  55. Marugan, J.; Hufschmidt, D.; Sagawe, G.; Selzer, V.; Bahnemann, D. Optical density and photonic efficiency of silica-supported TiO2 photocatalysts. Water Res. 2006, 40, 833–839. [Google Scholar] [CrossRef] [PubMed]
  56. Sakthivel, S.; Shankar, M.V.; Palanichamy, M.; Arabindoo, B.; Bahnemann, D.W.; Murugesan, V. Enhancement of photocatalytic activity by metal deposition: Characterisation and photonic efficiency of Pt, Au and Pd deposited on TiO2 catalyst. Water Res. 2004, 38, 3001–3008. [Google Scholar] [CrossRef] [PubMed]
Figure 1. SEM-EDS spectrum of ZnBiErO4.
Figure 1. SEM-EDS spectrum of ZnBiErO4.
Materials 11 00303 g001
Figure 2. TEM image of ZnBiErO4.
Figure 2. TEM image of ZnBiErO4.
Materials 11 00303 g002
Figure 3. The selected area electron diffraction pattern of ZnBiErO4.
Figure 3. The selected area electron diffraction pattern of ZnBiErO4.
Materials 11 00303 g003
Figure 4. The diagram of particle size of ZnBiErO4.
Figure 4. The diagram of particle size of ZnBiErO4.
Materials 11 00303 g004
Figure 5. X-ray powder diffraction patterns of ZnBiErO4, ZnO, Bi2O3 and Er2O3.
Figure 5. X-ray powder diffraction patterns of ZnBiErO4, ZnO, Bi2O3 and Er2O3.
Materials 11 00303 g005
Figure 6. Pawley refinements of XRD data for novel photocatalyst ZnBiErO4 prepared by the solid-state reaction method at 1000 °C. The solid line represents experimental X-ray diffraction pattern. The dot line represents simulation X-ray diffraction pattern. The tic marks represent reflection position. A difference between observed and calculated profile is shown beneath.
Figure 6. Pawley refinements of XRD data for novel photocatalyst ZnBiErO4 prepared by the solid-state reaction method at 1000 °C. The solid line represents experimental X-ray diffraction pattern. The dot line represents simulation X-ray diffraction pattern. The tic marks represent reflection position. A difference between observed and calculated profile is shown beneath.
Materials 11 00303 g006
Figure 7. Atomic structure of ZnBiErO4.
Figure 7. Atomic structure of ZnBiErO4.
Materials 11 00303 g007
Figure 8. (a) The XPS full spectrum of ZnBiErO4; (b) Zn2p spectrum of ZnBiErO4; (c) Bi4f spectrum of ZnBiErO4; (d) Er4d spectrum of ZnBiErO4; (e) O1s spectrum of ZnBiErO4.
Figure 8. (a) The XPS full spectrum of ZnBiErO4; (b) Zn2p spectrum of ZnBiErO4; (c) Bi4f spectrum of ZnBiErO4; (d) Er4d spectrum of ZnBiErO4; (e) O1s spectrum of ZnBiErO4.
Materials 11 00303 g008aMaterials 11 00303 g008b
Figure 9. UV-Vis absorption spectra of ZnBiErO4 and N-doped TiO2. (a) plot of ln(αhν) versus ln(hν − Eg) for ZnBiErO4. (b) plot of (αhν)1/2 versus for ZnBiErO4.
Figure 9. UV-Vis absorption spectra of ZnBiErO4 and N-doped TiO2. (a) plot of ln(αhν) versus ln(hν − Eg) for ZnBiErO4. (b) plot of (αhν)1/2 versus for ZnBiErO4.
Materials 11 00303 g009
Figure 10. The irradiance spectrum of the Xeon lamp light (cited from [45]).
Figure 10. The irradiance spectrum of the Xeon lamp light (cited from [45]).
Materials 11 00303 g010
Figure 11. Temporal spectral changes of aqueous solutions of methyl blue due to light irradiation in the presence of ZnBiErO4.
Figure 11. Temporal spectral changes of aqueous solutions of methyl blue due to light irradiation in the presence of ZnBiErO4.
Materials 11 00303 g011
Figure 12. Photocatalytic degradation of methyl blue under visible light irradiation in the presence of ZnBiErO4 and N-TiO2 as well as in the absence of a photocatalyst.
Figure 12. Photocatalytic degradation of methyl blue under visible light irradiation in the presence of ZnBiErO4 and N-TiO2 as well as in the absence of a photocatalyst.
Materials 11 00303 g012
Figure 13. Disappearance of total organic carbon (TOC) during photocatalytic degradation of methyl blue under visible light irradiation in the presence of ZnBiErO4 and N-TiO2 as well as in the absence of a photocatalyst.
Figure 13. Disappearance of total organic carbon (TOC) during photocatalytic degradation of methyl blue under visible light irradiation in the presence of ZnBiErO4 and N-TiO2 as well as in the absence of a photocatalyst.
Materials 11 00303 g013
Figure 14. First-order kinetic plots for the photocatalytic degradation of methyl blue with ZnBiErO4 or N-TiO2 as catalyst under visible light irradiation.
Figure 14. First-order kinetic plots for the photocatalytic degradation of methyl blue with ZnBiErO4 or N-TiO2 as catalyst under visible light irradiation.
Materials 11 00303 g014
Figure 15. Production of CO2 during the photocatalytic degradation of methyl blue with ZnBiErO4 or N-TiO2 as catalyst under visible light irradiation.
Figure 15. Production of CO2 during the photocatalytic degradation of methyl blue with ZnBiErO4 or N-TiO2 as catalyst under visible light irradiation.
Materials 11 00303 g015
Figure 16. The concentration variation of NO3 and SO42− during photocatalytic degradation of methyl with ZnBiErO4 as catalyst under visible light irradiation.
Figure 16. The concentration variation of NO3 and SO42− during photocatalytic degradation of methyl with ZnBiErO4 as catalyst under visible light irradiation.
Materials 11 00303 g016
Figure 17. Photocatalytic degradation of methyl blue under visible light irradiation in the presence of ZnBiErO4, ZnO, Bi2O3, Er2O3 and mixture.
Figure 17. Photocatalytic degradation of methyl blue under visible light irradiation in the presence of ZnBiErO4, ZnO, Bi2O3, Er2O3 and mixture.
Materials 11 00303 g017
Figure 18. Photocatalytic degradation of methyl blue with ZnBiErO4 or P25 as catalyst under visible light irradiation.
Figure 18. Photocatalytic degradation of methyl blue with ZnBiErO4 or P25 as catalyst under visible light irradiation.
Materials 11 00303 g018
Figure 19. Photocatalytic degradation of methyl blue with ZnBiErO4 or P25 as catalyst under sunlight irradiation.
Figure 19. Photocatalytic degradation of methyl blue with ZnBiErO4 or P25 as catalyst under sunlight irradiation.
Materials 11 00303 g019
Figure 20. The photocatalytic degradation of phenol with ZnBiErO4 as a photocatalyst under visible light irradiation.
Figure 20. The photocatalytic degradation of phenol with ZnBiErO4 as a photocatalyst under visible light irradiation.
Materials 11 00303 g020
Figure 21. Photodegradation of methyl blue over ZnBiErO4 in the presence of various scavengers: sodium oxalate, superoxide dismutase (SOD), H2O2, tert-butyl alcohol (TBA) and no scavenger.
Figure 21. Photodegradation of methyl blue over ZnBiErO4 in the presence of various scavengers: sodium oxalate, superoxide dismutase (SOD), H2O2, tert-butyl alcohol (TBA) and no scavenger.
Materials 11 00303 g021
Figure 22. Suggested band structure of ZnBiErO4.
Figure 22. Suggested band structure of ZnBiErO4.
Materials 11 00303 g022
Table 1. Atomic coordinates and structural parameters of ZnBiErO4 prepared by the solid reaction method.
Table 1. Atomic coordinates and structural parameters of ZnBiErO4 prepared by the solid reaction method.
AtomxyzOccupation Factor
Zn000.51
Bi0001
Er0001
O0.767310.140130.081881
Table 2. Binding energies (BE) for key elements from ZnBiErO4.
Table 2. Binding energies (BE) for key elements from ZnBiErO4.
CompoundZn2p1/2 BE (eV)Zn2p3/2 BE (eV)Bi4f7/2 BE (eV)Bi4f5/2 BE (eV)Er4d BE (eV)O1s BE (eV)
ZnBiErO41043.61020.3157.7163.0162.9529.8

Share and Cite

MDPI and ACS Style

Luan, J.; Zhuang, Y. Synthesis, Structural Property, Photophysical Property, Photocatalytic Property of Novel ZnBiErO4 under Visible Light Irradiation. Materials 2018, 11, 303. https://doi.org/10.3390/ma11020303

AMA Style

Luan J, Zhuang Y. Synthesis, Structural Property, Photophysical Property, Photocatalytic Property of Novel ZnBiErO4 under Visible Light Irradiation. Materials. 2018; 11(2):303. https://doi.org/10.3390/ma11020303

Chicago/Turabian Style

Luan, Jingfei, and Yan Zhuang. 2018. "Synthesis, Structural Property, Photophysical Property, Photocatalytic Property of Novel ZnBiErO4 under Visible Light Irradiation" Materials 11, no. 2: 303. https://doi.org/10.3390/ma11020303

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop