Next Article in Journal
Mechanistic Insights for Dry Reforming of Methane on Cu/Ni Bimetallic Catalysts: DFT-Assisted Microkinetic Analysis for Coke Resistance
Next Article in Special Issue
Au-TiO2 Synthesized by a Microwave- and Sonochemistry-Assisted Sol-Gel Method: Characterization and Application as Photocatalyst
Previous Article in Journal
One-Step Fabrication of PtSn/γ-Al2O3 Catalysts with La Post-Modification for Propane Dehydrogenation
Previous Article in Special Issue
The Effect of Electrolytes on the Photodegradation Kinetics of Caffeine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Characterization of Zinc Peroxide Nanoparticles for the Photodegradation of Nitrobenzene Assisted by UV-Light

by
Jesús Isaías De León Ramírez
1,
Víctor Alfredo Reyes Villegas
1,
Sergio Pérez Sicairos
2,
Esteban Hernández Guevara
1,
Mirna Del Carmen Brito Perea
1 and
Bertha Landeros Sánchez
1,*
1
Faculty of Chemical Sciences and Engineering, Autonomous University of Baja California, Calzada Tecnológico 14418, Mesa de Otay, Tijuana 22390, Mexico
2
Center for Graduates and Research in Chemistry, National Technological Institute of Mexico Technological Institute of Tijuana, Apdo, Tijuana 22000, Mexico
*
Author to whom correspondence should be addressed.
Submission received: 12 August 2020 / Revised: 2 September 2020 / Accepted: 5 September 2020 / Published: 10 September 2020
(This article belongs to the Special Issue Photocatalytic Degradation of Organic Wastes in Water)

Abstract

:
The contamination of both soil and water by nitrobenzene (NB) is a problem that has been studied, where several reactive agents have been developed for the degradation of this compound as well as different methods. Nanoparticles with semiconductive properties have been studied for organic compounds photodegradation due to their assistance in optimizing the degradation processes. Two of the most promising photocatalysts are ZnO and TiO2 because of their optimal results. In the present work the performance of the zinc peroxide (ZnO2) nanoparticles was evaluated. ZnO2 nanoparticles were synthesized from zinc acetate and hydrogen peroxide using the Sol-Gel method under ultrasound assistance. The characterization was carried out by UV–Vis spectroscopy, infrared Fourier transform total reflectance (ATR-FT-IR) spectroscopy, Raman spectroscopy, X-ray diffraction (XRD), Zeta potential, dynamic light scattering (DLS), field emission scanning electron microscopy (FE-SEM), and Energy Dispersive X-ray spectroscopy (EDX). The experiments for the degradation of NB were carried out in a photoreactor with UV lamps of 254 nm at 25 °C, using a solution of nitrobenzene with the nanoparticles. The best conditions for NB photodegradation were 30 ppm (ZnO2) and 15 ppm (NB) at pH 2, reaching up to 90% degradation in 2 h. The intermediates formed during the photodegradation of NB were identified by gas chromatography mass spectrometry.

Graphical Abstract

1. Introduction

Nitrobenzene (NB) is an aromatic molecule that causes toxicity in living organisms. Exposure can lead to absorption through different routes, such as dermal, inhalation or orally. When NB is absorbed into the systemic circulation, it causes methemoglobinemia, hepatoxicity, neurotoxicity, inhibition of liver enzymes, conjectured carcinogenicity and also has been found to cause genotoxicity in Drosophila [1,2,3]. This carcinogenic pollutant is widely used for the fabrication of different types of products, such as dyes, explosives, and pesticides [4]. NB has become a potential contaminant and of greater interest for its investigation due to the effects it can have on human beings [5,6,7,8]. The United States Environmental Protection Agency has listed NB as one of the most important pollutants. Nitrobenzene is a toxic agent that even at low concentrations can be a pollutant for the environment and living organisms [9]. Some agencies (National Institute of Environmental Health Sciences and Environmental Protection Agency) have focused their attention on NB, due to possible carcinogenic effects and water contamination, which is limited to 17 ppm in lakes and streams to prevent ingestion of fish and water that could cause health problems [10].
In this regard, several treatment processes were reported for effective nitrobenzene removal, which can be achieved by physical and chemical methods such as electrochemical reduction, catalysis, ozonation, biodegradation, adsorption, and photocatalysis [11,12,13,14,15,16,17]. This last method is one of the most promising for the following reasons: (1) complete mineralization, (2) no waste disposal problem, (3) low cost and (4) only mild temperature and pressure conditions are necessary [18].
The successful use of H2O2 as a source of active oxygen for environmental decontamination and disinfection, as well as an oxygen source for biodegradation of organic pollutants, has been amply reported [19]. In the photocatalysis field, semiconductors are of great interest, especially when they are combined with H2O2 [20,21]. When a semiconductor through the effects of electromagnetic radiation generates an electron/hole pair, their redox potential needs to be sufficiently positive and negative for the hole and electron, respectively, to generate free radicals that could subsequently degrade an organic contaminant [18]. These free radicals could be generated from environmental O2 instead of H2O2 if the needed redox potential is reached.
A technique involving a semiconductor photocatalyst and a contaminant has been used for the purification of air and water, which consists of the production of an electron/hole pair after illuminating the semiconductor with energetic enough photons [22]. It has also been seen that irradiation of a nano photocatalyst activity is enhanced through two synergistic mechanisms: the effect of plasmonic hot electrons and photothermal collective heating [23]. Titanium dioxide, TiO2, is the most common semiconductor photocatalyst and has been used for the degradation of energetic compounds like NB when exposed to ultraviolet irradiation thanks to their large bandgap (e.g., 3.0–3.2 eV) [24,25,26]. In the field of photocatalysis, several materials have been investigated. One of the most promising candidates is ZnO due to its potential as a green environmental management system. The properties that stand out in this material are direct and wide bandgap in the spectral region near the UV rays, strong oxidation capacity, good photocatalytic property and large energy of free exciton binding. There is evidence that in some cases the coupling of a semiconductor can reduce the recombination of the electron–hole pair, increasing the charge carrier’s lifetime by effects of charge separation. Lee et al. (2016) demonstrated a 25% improvement in the photodegradation efficiency of ZnO by coupling it with ZnO2 [22].
ZnO2 is also a semiconductor with a bandgap greater than that of ZnO. Nanocrystalline ZnO2 material is an indirect semiconductor with an energy gap near 4.5 eV [27], although its real value has been reported to be between 3.3–4.6 eV [28]. ZnO/ZnO2 composite has shown a superior photocatalytic activity as compared with either of its constituents (ZnO or ZnO2). A trend was found to be correlated with the amount of ZnO2 granules present on ZnO crystallites, at a higher concentration of ZnO2 on the surface a higher specific surface activity was observed, suggesting that ZnO2 alone could possess good photcalytic activity. However, an experiment using this ZnO2 powder alone presented a low photocatalytic activity regardless of its synthesis process (hydrothermally treated or untreated). This was expected since the absorption wavelength of the material (300 nm) and the light used in the irradiation had the same maximum-intensity wavelength, reducing the efficiency of electron/hole pair generation. Also, a reduction of the excitons lifetime is expected since the material exhibited a poor crystallinity [29]. Despite the fact that peroxide materials are stable at low temperatures, and that the conduction band of ZnO2 is higher than that of other catalysts, such as TiO2, it has been proved that ZnO2 continues to preserve its crystalline structure at high temperatures (below 120) [30], and its facile synthesis and non-toxic nature [31] give rise to considering it as a candidate. In addition, the conversion of ZnO2 to ZnO that takes place through the photoreaction releases O2− and ZnO [30], which is a material with a small bandgap [22]. These factors can favor degradation since the two materials (ZnO2/ZnO) will be coexisting in the system [30], increasing the photon absorption range. This phase transition does not compromise its recycling and reusability since it has been shown that ZnO2 is a material that is easy, quick and has a low cost to regenerate, conserving its efficient catalytic activity [28].
Due to the observations stated above, in this study we synthesized and tested a highly crystalline powder of ZnO2 with a light source with a wavelength of 254 nm which could generate electrons and holes more efficiently than a light source of 300 nm. The sol-gel/sonochemical route was used to obtain ZnO2 with certain characteristics (highly crystalline and in the nanoscale) to evaluate its performance as a photocatalyst. It has been observed that this method has been useful to synthesize particles with a consistent size and crystallinity, hence that is why this method was selected [31]. In addition, the material that we synthesized contains crystals of a size smaller than 10 nm, which can support to enhance the adsorption of NB molecules because of their high surface/volume ratio and in this way benefiting the degradation of NB. Another property of the manufactured nanoparticles is that having O2 available can help create more reactive species that play a role in the photodegradation reaction. Hence, we selected the ZnO2 catalyst as a model for studying the degradation profiles and its mechanism in nitrobenzene photodegradation.

2. Results and Discussion

2.1. Synthesis of ZnO2 Nanoparticles

The Following reaction (Equation (1)) takes place in the preparation of nanoparticles of zinc peroxide [31]:
Zn ( O 2 CCH 3 ) 2 +   H 2 O 2   ZnO 2 + 2 HO 2 CCH 3
The formation of ZnO2 nanoparticles was confirmed through visual assessment due to the white color and further confirmed throughout the characterization.

2.2. Characterization of ZnO2 Nanoparticles

2.2.1. UV-Vis Spectroscopy

Figure 1A shows the UV-Vis spectra of ZnO2 with a maximum excitation band at 285 nm is in agreement with the data found in the literature, perceiving a slight shift which can be due to the size of the nanoparticle [32]. The comparison of this maximum with that of Drmosh et al. (2010), where a nanoparticle with a size of 3.7 nm exhibits an absorption band at 280 nm, supports the statement that the shifts of the maximum absorption are due to the quantum confinement effects that appear in the nanoscale. When the particle size decreases enough to present properties due to quantum confinement the bandgap of the particle increases, and its absorption shifts to a lower wavelength [33]. Additionally, in Figure 1A it is seen that the spectrum of the reactant (dotted line) has no characteristic absorption, unlike the product, where it is observed that the spectrum changes after the reaction, with a noticeable change in the UV-Vis spectrum. These properties indicate that a ZnO2 phase is probably present.
As we know, the UV-Vis spectrum gives us information not only about the optical properties but also about the bandgap (Figure 1B). It is known that there is a strong hybridization between the states Zn-d and O-p. ZnO2 is an indirect semiconductor with the upper part of the valence band located in Γ and the lower part of the conductance band located between Γ and R. For ZnO2 the bandgap is 2.3 eV, however, it is generally underestimated by approximately 50–100% using density functional theory. Hence, the real value of the energy gap for ZnO2 could be in the range of 3.3–4.6 eV, in agreement with the value calculated from spectrum [27]. The energy calculation of the bandgap (Equation (2)) can be calculated from the UV-vis absorption spectrum. The band interval energy was calculated using the Tauc method as seen below:
α h ϑ = C ( h ϑ E g   ) n
where α is the absorption coefficient, h ϑ is the incident energy, n = 2, C is a constant and Eg the optical bandgap. The graph was plotted with the values of ( α h ϑ ) 1 / 2   vs   h ϑ   ( E g ) for the calculation of the indirect band energy the linear part is extrapolated to zero with the abscissa ( h ϑ   eV ) [34,35]. The bandgap calculated was 3.7 eV (Figure 1B), which is within the range consulted in the literature [27].

2.2.2. Infrared Fourier Transform (FT-IR)

The FT-IR spectra of the ZnO2 nanoparticles are shown in Figure 2A, with characteristic peaks at 667 cm−1, 1022 cm−1, 1405 cm−1 and 1565 cm−1. The peak positions centered at 667 cm−1, 1022 cm−1 and 1405 cm−1 may be due to the O–O bands corresponding to the peroxide ( O 2 2 ) ions of ZnO2 nanoparticles. Another sharp absorption band of the ZnO2 is observed located at 667 cm−1, which corresponds to Zn–O vibration. As Verma et al. (2014) observed for ZnO2 nanoparticles (average crystallite size of 151 nm) characteristic vibrations cause peaks at a wavenumber of 436 cm−1 for Zn–O bond, and 1040, 1334, and 1450 cm−1 for O–O bond [28]. The shifts of the wavenumbers could be due to the particle size, since there is evidence that suggest that the characteristic Zn–O vibration has a redshift as the particle size increases [32,36]. The difference of the crystallite size from those synthesized by Verma et al. (2014) and ours are 145 nm, with ours having an average of approximately 6 nm, supporting this evidence. In addition, the broad peak in the wavelength of 3384 cm−1 and the band 1565 cm−1 corresponds to characteristic stretches of the water molecule [37]. On the other hand, the bands shown in 1536 cm−1 and 1405 cm−1 may correspond to the vibrations of the reactant (zinc acetate) since they are characteristic of vibrations of C–O and C–C bonds [38]. This could be part of the reactant that did not react. It may sound logical that these are still present in the sample since the sample did not undergo any heat treatment, or could be arising from the absorption of atmospheric CO2 on the surface of the nanoparticles [32]. In contrast, the % proportions of the atoms from the EDX analysis (Figure 3B) correlate with ZnO2, with no extra oxygen coming from the acetate.

2.2.3. Raman Spectroscopy

The Raman spectra of ZnO2 nanoparticles are shown in Figure 2B where 5 signals are present, at the positions of 97, 395, 473, 847 and 930 cm−1 respectively. These peaks are assigned to ZnO2 according to the single phase observed in the XRD results. For instance, some peaks display low intensity and have been found in other peroxides [30] and the peaks at 847 and 930 cm−1 are in agreement with the Raman spectra of nanocrystalline ZnO2 [39]. As for the most intense peak, a stretch mode of a O–O vibration that is not well resolved in the infrared spectrum is centered at approx. 840 cm−1 [30], suggesting that this mode could be ascribed to the band at 847 cm−1.

2.2.4. X-ray Diffraction

To study the crystalline structure of the ZnO2 nanoparticles, the powder was subjected to an X-ray diffraction analysis, obtaining the diffractogram shown in Figure 3A.
Where it is observed that all the peaks fit perfectly to the data found in the JCPDS reference (01-077-2414) that belongs to a cubic structure of ZnO2. For the calculations of lattice parameter, cell volume and atomic radius the following formula (Equation (3)) was used:
d hkl =   a h 2 k 2 l 2 =   n   λ 2   sin   θ
where hkl are the miller indices of the plane, n the diffraction order (commonly given as 1), λ is the wavelength (1541 A. CuKα) and θ is the Bragg diffraction angle of the plane [40]. These calculations are shown in Table 1 where we calculated, the typical characteristics of the cubic structures. For the crystal size, the Scherrer equation was used [41,42] obtaining a crystal size of approximately 6 nm.

2.2.5. Scanning Electronic Microscope (SEM) and Dynamic Light Scattering (DLS)

In Figure 3B the typical EDX spectrum of ZnO2 is shown. This indicates the composition of the sample which is formed by Zn and O (the atomic ratio Zn/O is 1:1.54), approaching that of the consulted bibliography [43], supporting the approximated chemical composition that could be estimated for ZnO2 according to the obtained XRD results. As mentioned above the size of the crystals calculated by the Scherrer equation was about 6 nm, in Figure 3C we can observe a micrograph of a ZnO2 nanoparticle where it can be seen that the particle (~100 nm) is made of small crystals which were measured to be of about 6 nm.
Figure 4A shows a photo taken with an accelerating voltage of 5 KeV and subjected to image analysis with the help of J-image software for the measurement of the particle diameter distribution, obtaining an average diameter of approximately 100 nm with a normal distribution. This confirms the DLS analysis (Figure 4B) from which we obtained a Dh of ~100 nm, meaning that we synthesized nanoparticle clusters of about 100 nm with a normal distribution made from small crystals of ~6 nm.

2.2.6. Zeta Potential

Figure 5 presents the mean zeta potential profile at varying pH of ZnO2 suspensions. This was calculated between a pH of 3 and 9 and the initial pH of the solutions was at pH 6.3, having an average zeta potential of +15 mV. As the pH increased, the zeta potential was found to decline gradually to a point of zero charges (PZC), which was measured between pH 7.5–8.0. After this point the solution showed an agglomeration. When reaching an alkaline pH of 9 the nanoparticles presented a charge of −18 mV ± 2. At alkaline pH, the dominant species are Zn(OH)2 (aq), which could precipitate giving a new solid phase, Zn(OH)2 (s). However, if the pH keeps rising, the global charge is mediated by the prevalence of negative species such as Zn(OH)3 (aq) and Zn(OH)2−4 (aq), exhibiting a negative surface [44]. By decreasing the pH of the solution below pH 7 it showed a dispersion region by increasing the zeta potential, thus having a dissolution region and agglomeration region agreeing with the consulted reference [45]. An isoelectric point (IEP) was observed at pH 3.2 (Figure 4), however two points at zero were detected, IEP at an acid pH and a PZC at an alkaline pH matching the reference [46].

2.3. Photocatalytic Degradation of Nitrobenzene

2.3.1. Effect of Photocatalyst Concentration

Figure 6 presents the UV−vis absorption spectra of NB solution samples after photoreaction for different times. The typical spectrum of NB shows an absorption peak between 250 nm and 268 nm [47]. NB shows an absorption peak at 265, which decreased greatly in intensity as the UV exposure time increased in the presence of the ZnO2 semiconductor. When an electron/hole pair is generated at a surface, several events can occur. They can recombine, irradiating the input energy as heat, cause collective plasmonic heating, produce hot electrons [23], enter in a metastable state or react with electron donors and acceptors adsorbed or near the surface of the particle [48,49,50]. These two last reasons are the main cause of the decrease in NB concentration. The following formula (Equation (4)) was used to obtain the percentage of degradation:
Degradation   ( % ) = [ ( Co Ct / Co ) ] x 100  
where Co is the initial concentration or the initial absorbance of NB and Ct is the concentration or the absorbance of NB at a certain time ‘t’ in minutes [51].
NB absorbs at a wavelength close to the intensity of the lamps, which could suggest that the NB would degrade in the absence of the catalyst. However, the control experiment (Data not shown) with no photocatalyst was run and this event was discarded due to the low NB removal from solution, which was less than 10%.
Figure 7A shows the percentage of NB photodegradation at different ZnO2 nanoparticle concentrations, being 20, 30 and 40 ppm with a NB% degradation of 53%, 63% and 57% respectively. It can be seen that increasing the concentration from 20 ppm to 30 ppm the NB degradation increased, confirming the positive influence of the increased number of ZnO2 active sites on the process [52]. Therefore, the efficiency could be enhanced with the increase of ZnO2 concentration because the total surface area available for nitrobenzene adsorption increased [53].
On the contrary, increasing the concentration to 40 ppm had a repercussion on the degradation, the photocatalytic reaction decreases due to the decrease of light penetration and the increase of UV light scattering [54]. This decrease could also be attributed to a blocking effect taking place on the surface of the photocatalyst decreasing the efficiency of decomposition, as consequence of excessive O2 ions acting as scavengers for OH radicals [55].

2.3.2. Effect of pH

The effect of pH on the photocatalytic activity (Figure 7B) was clear. The nanoparticles of ZnO2 degraded 63%, 65%, and 98% for pH 7, 10 and 2 respectively, with pH 2 being the most effective. The reason may be that when the pH is 2.0, there is a strong Coulomb force between the surface of ZnO2 nanoparticle and nitrobenzene, which is favorable for the adsorption of nitrobenzene on the surface of the photocatalyst [53]. The ZnO2 colloid nanoparticles between a pH of 2 and 3 would, therefore, have a negative surface charge (Figure 5) and would be expected to experience a significant attraction to surfaces with a positive surface charge. The pKa of the NB is 3.98, where the anionic form dominates upon increasing the pH of the solution [56].
Therefore, at a pH between 2 and 3 ZnO2 has a negative zeta potential and NB a positive charge under these conditions, which would boost attractions between nanoparticles and NB. Additionally, below pH 2 ZnO2 nanoparticles release O2 ions [57,58]. In the holes bridging O atoms with unit formal charge O are powerful oxidizing agents, which are capable of pulling electrons from adsorbed H2O to yield H2O+, which can dissociate to yield a proton and the hydroxyl surface radical [49], as expressed in the following Equations (5)–(10):
ZnO 2 + hv e cb + h vb +
e cb + O 2 O 2 ·
h vb + +   OH OH
h vb + +   H 2 O H 2 O +
H 2 O + OH +   H +
OH +   NB   Hydroxyl   intermediates +   CO 2 +   H 2 O
Also photogenerated are electrons in the conduction band ( e cb ) and holes within the valence band ( h vb + ), which can either recombine or reside in a trapped state or within the respective band. O2 is an effective electron scavenger, forming either O2 or O2 −2 [48]. Newly formed superoxide radical anion ( O 2 ) further reacts with water to generate hydrogen peroxide anions and increase the number of hydroxyl radicals, increasing the performance of the reaction [59]. In this manner, it improves the availability of contact and the amount of surface OH generated (Equations (5) and (7)) and increasing % degradation. These equations could be supported if the hydroxyl intermediates are identified.
Frequently, in both the literature and international standards, SEC is used as an energy performance indicator to evaluate or measure the performance of energy efficiency [60]. Figure 7C shows the energy efficiency of the effect of pH on the photocatalytic activity. At the optimal parameters (pH 2 and 30 ppm ZnO2) and after 120 min of treatment, 97.7% of NB was degraded with a corresponding SEC of approx. 60,000 kWh/kg.
The photocatalytic activity of the prepared nanoparticles, compared with the most relevant reported photocatalysts for the degradation of NB (Table 2), had a high 98% removal. As shown in Table 2 our percentage of degradation (98%) is achieved in half or, in some cases, even at one quarter of the time (2 h) compared to the obtained degradation by other catalyst such as: TiO2 nanotubes (50 mg) [61], SrFeO3−δ (1000 ppm) [62], TiO (0.3% w/v) [63] and TiO2 (50 mg) [10], which degraded the NB in solution with an efficiency of 100%, 99%, 98.5% and 98% in 4, 6, 4, 8 h respectively. Moreover, G–ZnO–Au NCs, see Table 2 [64], and (0.50%) Zn doped TiO2 had a 98 and 97 percentage of degradation in 2.3 and 2 h respectively, having very similar results as our synthesized nanoparticles. In short, we have synthesized ZnO2 nanoparticles by a facile method capable of acting as a photocatalyst for the mineralization of NB as good as the catalyst reported in the literature or better. These novel findings contribute another option for the removal of this contaminant. The sol-gel/sonochemical method has proven to be useful for the synthesis of ZnO2 nanoparticles as a photocatalyst, however it is necessary to study in more detail the synthesis factors that could indirectly influence the photocatalysis reaction efficiency, such as reactant ratios, sonication time, temperature, and drying methods, which could have an effect on the size and crystallization of the particles.
To explore the mechanism of NB photodegradation by ZnO2 nanoparticles, the possible intermediates formed during the reaction were identified by GC-MS, where 2-nitrophenol (2-NPhs), 3-nitrophenol(3-NPhs) and 4-nitrophenol(4-NPhs) were detected to be the major products from our nitrobenzene degradation. The NB peak appeared in the GC chromatogram for pH 2 (Figure 8A), 7 (Figure 8B) and 10 (Figure 8C) at a retention time near 5.9 min, whereas the peaks corresponding to the intermediates appeared at retention times near 6.2 min (same retention time for all three pHs). Figure 9 shows the mass spectrum result of the mentioned peaks, which were assigned to NB (A), 2-NPhs (B), 3-NPhs (C) and 4-NPhs (D), respectively. The concentration of the intermediates was measured by calculating the area under the curve from the chromatogram peaks in proportion to the known NB concentration. The reaction intermediates detected for the pH 2 (Figure 8A), 7 (Figure 8B) and 10 (Figure 8C) were 0.51%, 6.3% and 1.2% respectively.
Rendering these products, it is proposed that nitrobenzene could be oxidized through one pathway, and hydroxyl radicals could be added onto the aromatic ring of nitrobenzene and form hydroxy cyclohexadienyl radicals [68,69].
The hydroxy cyclohexadienyl radicals by oxidation or disproportionation could produce 2-nitrophenol, 3-nitrophenol, and 4-nitrophenol, depending on the position (ortho, meta, para) of the aromatic ring of NB it was added on to. However, the nitrophenol distribution does not correlate with the deactivating characteristics of the nitro group [70].
This result confirms the role of the hydroxyl radical as an initiator of the oxidative reaction, supporting a mineralization and the mechanism stated at Equations (6)–(11). However, further experiments should be made to completely elucidate the mineralization mechanism of NB’s photodegradation by ZnO2 nanoparticles.

3. Materials and Methods

3.1. Synthesis of ZnO2 Nanoparticles

The ZnO2 nanoparticles were synthesized by a sol-gel/sonochemical route; for this purpose, a precursor solution of 50 mL was prepared using 5 mL of 30% peroxide with HPLC grade water from J.T Baker, the solution was sonicated (TPC Advance Technology UC-450 Model) in 2 min. Then, 1 g of zinc acetate (99.9%) from TSA reagents was dissolved in the precursor solution by ultrasonication for 5 min until a homogeneous solution is obtained. Subsequently it was sonicated for half an hour at a temperature of 60 ± 5 °C. Once this time elapsed a white suspension was obtained [31]. This solution was placed in a Petri dish and allowed to dry at 36 °C for 24 h. A white powder was obtained which was washed twice with excess distilled water for the synthesis of ZnO2 nanoparticles.

3.2. Characterization of ZnO2 Nanoparticles

3.2.1. UV-VIS Spectroscopy

To evaluate the optical properties of the nanoparticles and try to understand the electronic structure that the nanoparticles could have, a HACH UV-VIS DR 6000 spectrophotometer (Dusseldorf, Germany) was used and a 30 ppm solution of ZnO2 nanoparticles dissolved in water was read. The obtained spectrum was made in a sweeping range and 1 cm quartz cells were used.

3.2.2. FT-IR Analysis

To study the binding vibrations of ZnO2, FT-IR analysis was employed. The spectrum was obtained by Nicolet IS10 Thermo Scientific FT-IR equipment (Waltham, MA, USA), with the ATR method in a sweep of 500 cm−1 to 4000 cm−1 (100 scans were performed per reading and 3 readings were carried out). Usually approximately 100 mg of powder sample was used.

3.2.3. Raman Analysis

The Raman spectrum was recorded using a Raman WITec Confocal Raman Microscope System (alpha300R, Witec Inc., Ulm, German) using an Nd: YVO4 laser with an excitation wavelength of 532 nm. The typical resolution used was achieved using a 100x objective with a grid of 672 lines/mm (4 cm−1) and a grid of 1800 lines/mm (1 cm−1). The rate of acquisition of point spectra was ≤10 μs, using a lamp power of 14.4 mW (118 µm). Considering the effects of laser temperature on the sample the spectra were corrected using a linear function.

3.2.4. X-ray Diffraction (XRD)

X-ray powder diffraction experiments were performed on a Brucker D8 Advance diffractometer (Bruker Co., Billerica, MA, USA) in the range of 15–75° (2θ) in step sizes of 0.02, with a scanning speed of 0.02°/1.2004 s. The radiation tube of Cu Kα (1.5406 Å) was operated at 30 kV and 30 mA.

3.2.5. Scanning Electronic Microscope (SEM)

The sample was analyzed with a gold coating to improve the resolution of the image. For the coating the sputtering method was followed (JFC−110 ion sputter, JEOL). The samples images were obtained in a scanning electron microscope (JEOL JSM7600F Tokyo, Japan) with an accelerating voltage of 5 keV. The size distribution and average grain and crystal diameter size were determined by counting the grains and crystals with an image processing and analysis software ImageJ for counting and the software for graphics and data analysis origin pro 8 for size distribution. The average diameter of each particle and crystal was determined by the measurement of at least 180–200 particles or crystals per image from 5 images. Besides, an energy dispersion spectroscopy analysis (EDS) was carried out for the chemical composition of the material using the same electron microscope (JEOL JSM7600F) with a range of operating voltage from 0–12 keV.

3.2.6. DLS Analysis and Zeta Potential

The dynamic light scattering (DLS) measurements to obtain the hydrodynamic diameter (Dh) were carried out at room temperature (25 °C) using a Zetasizer nano-ZS from Malvern Instruments (ZEN3500, Worcestershire, UK). The same equipment was used to measure the zeta potential (ζ) at different pH values adjusted with HCl and NaOH. Both the Dh and ζ measurements were evaluated at room temperature using a 300 ppm ZnO2 nanoparticle suspension.

3.3. Photocatalytic Degradation of Nitrobenzene

The reaction conditions to measure the photocatalytic activity of the ZnO2 nanoparticles for NB degradation are specified by Reynoso-Soto et al. (2013) [65]. Briefly, for the photocatalysis reactions the solution prepared was 250 mL of NB at 15 ppm and ZnO2 nanoparticles at 20, 30 and 40 ppm at a neutral pH unless specified. The samples were centrifuged in a microcentrifuge (Thermo Espresso Personal Microcentrifuge, Thermo Fisher Scientific Inc., MA, USA) for 10 min at 15 rpm to remove the nanoparticles from the solution. For the pH modifications, HCl and NaOH were used. All the photocatalysis experiments were done in duplicate. The specific energy consumption (SEC) was estimated and the following expression (Equation (11)) was used:
SEC = L × NP × t/m
where L = number of lamps, NP = nominal power of lamps (KW), t = time for maximum percentage of degradation (h) and m = mass of NB degraded (kg), which was calculated from the initial concentration of NB and the concentration of NB at “t” [71].
To identify the degradation intermediates of the NB photocatalysis degradation, gas chromatography–mass spectrometry (GC-MS) was applied. GC-MS were recorded on a Thermo Scientific TRACE 1310 (GC) and Thermo Scientific single quadrupole ISQ LT (MS), with a column model TG-SQC (30 m × 0.25 mm inner diameter, 0.25 µm film thickness). The detector temperature was 240 °C, the injector temperature was 250 °C, and transfer line temperature was 250 °C; oven temperature started at 70 °C for 3 min, increased at a 40 °C/min rate until 280 °C, with a hold time of 7 min. Helium was employed as a carrier gas, at 1 mL/min flow. Samples of 50 mL were extracted with 5 mL dichloromethane (duplicate), at a 10:1 ratio between sample and solvent. The organic extract was dried with anhydrous Na2SO4. One milliliter of the extract was placed in a 2 mL vial and then it was manually injected to GC.
The percentage of the degradation products detected was calculated using the area under the curve of the chromatogram peaks corresponding to NB and the nitrophenols (NPhs) in proportion to the NB% detected by UV-Vis spectroscopy. The percentage of reaction intermediates corresponds to the sum of these.

4. Conclusions

In this study we synthesized ZnO2 nanoparticles by a sol-gel/sonochemical route and its photocatalytic activity was evaluated by NB degradation. The synthesized nanoparticles have a crystal size of ~6 nm which cluster to form particles of 100 nm, with good dispersion in water having a positive ζ (+15 mV) at pH = 6.3 and an IEP (pH = 3–3.5) and PZC (pH = 7.5–8). ZnO2 nanoparticles presented great results as a photocatalyst, being more effective at a pH = 2 and a concentration ratio of 2:1 (nanoparticle: NB), with a SEC of approx. 60,000 kWh/kg. The combination of the electron–hole formed in the nanoparticle and an unpaired electron is highly active, which is this the key to start a hydroxylation reaction. This guides us to propose a further investigation of these aspects that will help us have a better understanding of the mechanism of reaction due to the intermediates detected by GC-MS. The GC-MS analysis shows that 2-nitrophenol, 3-nitrophenol, and 4-nitrophenol were the principal hydroxylated intermediates detected. The observed catalyst potential of ZnO2 nanoparticles opens exciting possibilities for exploring this synthesis route and optimizes the activity of this catalyst.

Author Contributions

Conceptualization, J.I.D.L.R. and V.A.R.V.; methodology, S.P.S.; software, V.A.R.V.; validation, J.I.D.L.R., V.A.R.V. and E.H.G.; formal analysis, J.I.D.L.R.; investigation, J.I.D.L.R.; resources, J.I.D.L.R.; data curation, V.A.R.V.; writing—original draft preparation, J.I.D.L.R.; writing—review and editing, V.A.R.V., J.I.D.L.R., S.P.S.; visualization, S.P.S.; supervision, S.P.S., B.L.S.; project administration, B.L.S.; funding acquisition B.L.S., M.D.C.B.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Council of Science and Technology (CONACYT) [CB2015-253128].

Acknowledgments

We would like to acknowledge Adriana Tejeda (IIM-UNAM) for the support in carrying out X-ray measurements; Omar Novelo (IIM-UNAM) for the support in the scanning electron microscopy characterization, María Elena Villafuerte Castrejón (IIM-UNAM) and Gerardo Cesar Diaz Trujillo (UABC) for an investigation residency in the Investigation Institute of Materials (IIM) and Arturo Estolano (UABC) for the support in the gas chromatography–mass spectrum analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Beauchamp, R.O.; Irons, R.D.; Rickert, D.E.; Couch, D.B.; Hamm, T.E.; Lyon, J.P. A critical review of the literature on nitrobenzene toxicity. Crit. Rev. Toxicol. 1982, 11, 33–84. [Google Scholar] [CrossRef]
  2. Katritzky, A.R.; Oliferenko, P.; Oliferenko, A.; Lomaka, A.; Karelson, M. Nitrobenzene toxicity: QSAR correlations and mechanistic interpretations. J. Phys. Org. Chem. 2003, 16, 811–817. [Google Scholar] [CrossRef]
  3. Harrison, M.R. Toxic methaemoglobinaemia. Anaesthesia 1977, 32, 270–272. [Google Scholar] [CrossRef] [PubMed]
  4. Mu, Y.; Yu, H.Q.; Zheng, J.C.; Zhang, S.J.; Sheng, G.P. Reductive degradation of nitrobenzene in aqueous solution by zero-valent iron. Chemosphere 2004, 54, 789–794. [Google Scholar] [CrossRef] [PubMed]
  5. Wang, C.; Feng, Y.; Zhao, S.; Li, B.L. A dynamic contaminant fate model of organic compound: A case study of Nitrobenzene pollution in Songhua River, China. Chemosphere 2012, 88, 69–76. [Google Scholar] [CrossRef] [PubMed]
  6. Cai, Z.; Fu, J.; Du, P.; Zhao, X.; Hao, X.; Liu, W.; Zhao, D. Reduction of nitrobenzene in aqueous and soil phases using carboxymethyl cellulose stabilized zero-valent iron nanoparticles. Chem. Eng. J. 2018, 332, 227–236. [Google Scholar] [CrossRef]
  7. Elnashaie, S.S.E.; Danafar, F.; Abashar, M.E.E. Maximum Production Minimum Pollution (MPMP), Necessary but not Sufficient for Sustainability. Eur. J. Sustain. Dev. Res. 2018, 2, 41. [Google Scholar] [CrossRef] [Green Version]
  8. Zhao, L.; Lu, Z.; Tan, S.; Ciren, J.; Tan, C. Effects of glucose and starch on the toxicity of nitrobenzene to plants and microbes in constructed wetlands. Sci. Total Environ. 2019, 658, 809–817. [Google Scholar] [CrossRef]
  9. Liu, L.-L.; Chen, J.; Yu, C.-X.; Lv, W.-X.; Yu, H.-Y.; Cui, X.-Q.; Liu, L. A novel Ag (I)-calix[4]arene coordination polymer for the sensitive detection and efficient photodegradation of nitrobenzene in aqueous solution. Dalt. Trans. 2017, 46, 178–185. [Google Scholar] [CrossRef]
  10. Tayade, R.J.; Kulkarni, R.G.; Jasra, R.V. Photocatalytic degradation of aqueous nitrobenzene by nanocrystalline TiO2. Ind. Eng. Chem. Res. 2006, 45, 922–927. [Google Scholar] [CrossRef]
  11. Cao, L.; Zhang, C.; Zou, S.; Zhu, G.; Li, N.; Zhang, Y.; Rittmann, B.E. Simultaneous anaerobic and aerobic transformations of nitrobenzene. J. Environ. Manag. 2018, 226, 264–269. [Google Scholar] [CrossRef] [PubMed]
  12. Li, Y.P.; Cao, H.B.; Liu, C.M.; Zhang, Y. Electrochemical reduction of nitrobenzene at carbon nanotube electrode. J. Hazard. Mater. 2007, 148, 158–163. [Google Scholar] [CrossRef] [PubMed]
  13. Chong, H.; Li, P.; Xiang, J.; Fu, F.; Zhang, D.; Ran, X.; Zhu, M. Design of an ultrasmall Au nanocluster-CeO2 mesoporous nanocomposite catalyst for nitrobenzene reduction. Nanoscale 2013, 5, 7622–7628. [Google Scholar] [CrossRef] [PubMed]
  14. Caprio, V.; Insola, A.; Volpiceli, G. Ozonation of Aqueous Solutions of Nitrobenzene. Ozone Sci. Eng. 2008, 6, 115–121. [Google Scholar] [CrossRef]
  15. Nishino, S.F.; Spain, J.C. Oxidative pathway for the biodegradation of nitrobenzene by Comamonas sp. strain JS765. Appl. Environ. Microbiol. 1995, 61, 2308–2313. [Google Scholar] [CrossRef] [Green Version]
  16. Wang, D.; Shan, H.; Sun, X.; Zhang, H.; Wu, Y. Removal of nitrobenzene from aqueous solution by adsorption onto carbonized sugarcane bagasse. Adsorpt. Sci. Technol. 2018, 36, 1366–1385. [Google Scholar] [CrossRef]
  17. Priya, M.H.; Madras, G. Photocatalytic degradation of nitrobenzenes with combustion synthesized nano-TiO2. J. Photochem. Photobiol. A Chem. 2006, 178, 1–7. [Google Scholar] [CrossRef]
  18. Bhatkhande, D.S.; Pangarkar, V.G.; Beenackers, A.A.C.M. Photocatalytic degradation for environmental applications—A review. J. Chem. Technol. Biotechnol. 2002, 77, 102–116. [Google Scholar] [CrossRef]
  19. Wolanov, Y.; Prikhodchenko, P.V.; Medvedev, A.G.; Pedahzur, R.; Lev, O. Zinc Dioxide Nanoparticulates: A Hydrogen Peroxide Source at Moderate pH. Environ. Sci. Technol. 2013, 47, 8769–8774. [Google Scholar] [CrossRef]
  20. Rodríguez, M.; Kirchner, A.; Contreras, S.; Chamarro, E.; Esplugas, S. Influence of H2O2 and Fe(III) in the photodegradation of nitrobenzene. J. Photochem. Photobiol. A Chem. 2000, 133, 123–127. [Google Scholar] [CrossRef]
  21. Contreras, S.; Rodríguez, M.; Chamarro, E.; Esplugas, S.; Casado, J. Oxidation of nitrobenzene by O3/UV: The influence of H2O2 and Fe(III). Experiences in a pilot plant. Water Sci. Technol. 2001, 44, 39–46. [Google Scholar] [CrossRef] [PubMed]
  22. Lee, K.M.; Lai, C.W.; Ngai, K.S.; Juan, J.C. Recent developments of zinc oxide based photocatalyst in water treatment technology: A review. Water Res. 2016, 88, 428–448. [Google Scholar] [CrossRef] [PubMed]
  23. Rej, S.; Mascaretti, L.; Santiago, E.Y.; Tomanec, O.; Kment, Š.; Wang, Z.; Zbořil, R.; Fornasiero, P.; Govorov, A.O.; Naldoni, A. Determining Plasmonic Hot Electrons and Photothermal Effects during H2 Evolution with TiN-Pt Nanohybrids. ACS Catal. 2020, 10, 5261–5271. [Google Scholar] [CrossRef]
  24. Crişan, M.; Mardare, D.; Ianculescu, A.; Drăgan, N.; Niţoi, I.; Crişan, D.; Voicescu, M.; Todan, L.; Oancea, P.; Adomniţei, C.; et al. Iron doped TiO2 films and their photoactivity in nitrobenzene removal from water. Appl. Surf. Sci. 2018, 455, 201–215. [Google Scholar] [CrossRef]
  25. Cropek, D.; Kemme, P.A.; Makarova, O.V.; Chen, L.X.; Rajh, T. Selective photocatalytic decomposition of nitrobenzene using surface modified TiO2 nanoparticles. J. Phys. Chem. C 2008, 112, 8311–8318. [Google Scholar] [CrossRef]
  26. Ayati, A.; Tanhaei, B.; Bamoharram, F.F.; Ahmadpour, A.; Maydannik, P.; Sillanpää, M. Photocatalytic degradation of nitrobenzene by gold nanoparticles decorated polyoxometalate immobilized TiO2 nanotubes. Sep. Purif. Technol. 2016, 171, 62–68. [Google Scholar] [CrossRef]
  27. Chen, W.; Lu, Y.H.; Wang, M.; Kroner, L.; Paul, H.; Fecht, H.-J.; Bednarcik, J.; Stahl, K.; Zhang, Z.L.; Wiedwald, U.; et al. Synthesis, Thermal Stability and Properties of ZnO2 Nanoparticles. J. Phys. Chem. C. 2009, 113, 1320–1324. [Google Scholar] [CrossRef] [Green Version]
  28. Verma, S.; Jain, S.L. Nanosized zinc peroxide (ZnO2): A novel inorganic oxidant for the oxidation of aromatic alcohols to carbonyl compounds. Inorg. Chem. Front. 2014, 1, 534. [Google Scholar] [CrossRef]
  29. Hsu, C.C.; Wu, N.L. Synthesis and photocatalytic activity of ZnO/ZnO2 composite. J. Photochem. Photobiol. A Chem. 2005, 172, 269–274. [Google Scholar] [CrossRef]
  30. Giannakoudakis, D.A.; Florent, M.; Wallace, R.; Secor, J.; Karwacki, C.; Bandosz, T.J. Zinc peroxide nanoparticles: Surface, chemical and optical properties and the effect of thermal treatment on the detoxification of mustard gas. Appl. Catal. B Environ. 2018, 226, 429–440. [Google Scholar] [CrossRef]
  31. Profesional, E.; Física, D.E.I.; Carlos, R.; Surichaqui, C. Sintesis y Caracterización de nanopartículas de Zn02 y su actividad antímicrobiana. EsESCUELA Prof. Ing. FÍSICA 2013, 84. [Google Scholar] [CrossRef]
  32. Gondal, M.A.; Drmosh, Q.A.; Yamani, Z.H.; Saleh, T.A. Synthesis of ZnO2 nanoparticles by laser ablation in liquid and their annealing transformation into ZnO nanoparticles. Appl. Surf. Sci. 2009, 256, 298–304. [Google Scholar] [CrossRef]
  33. Drmosh, Q.A.; Gondal, M.A.; Yamani, Z.H.; Saleh, T.A. Spectroscopic characterization approach to study surfactants effect on ZnO2 nanoparticles synthesis by laser ablation process. Appl. Surf. Sci. 2010, 256, 4661–4666. [Google Scholar] [CrossRef]
  34. Mallika, A.N.; Ramachandrareddy, A.; Sowribabu, K.; Reddy, K.V. Synthesis and optical characterization of aluminum doped ZnO nanoparticles. Ceram. Int. 2014, 40, 12171–12177. [Google Scholar] [CrossRef]
  35. Rusdi, R.; Rahman, A.A.; Mohamed, N.S.; Kamarudin, N.; Kamarulzaman, N. Preparation and band gap energies of ZnO nanotubes, nanorods and spherical nanostructures. Powder Technol. 2011, 210, 18–22. [Google Scholar] [CrossRef]
  36. Drmosh, Q. Synthesis and Characterization of Nano-Structure Metal Oxides and Peroxides Prepared by Laser Ablation in Liquids. Ph.D. Thesis, King Fahd University of Petroleum and Minerals, Dhahran, Saudi Arabia, 2010. [Google Scholar]
  37. Malhotra, V.M.; Jasty, S.; Mu, R. FT-IR spectra of water in microporous KBr pellets and water’s desorption kinetics. Appl. Spectrosc. 1989, 43, 638–645. [Google Scholar] [CrossRef]
  38. Escobedo-Morales, A.; Esparza, R.; García-Ruiz, A.; Aguilar, A.; Rubio-Rosas, E.; Pérez, R. Structural and vibrational properties of hydrothermally grown ZnO2 nanoparticles. J. Cryst. Growth 2011, 316, 37–41. [Google Scholar] [CrossRef]
  39. Yang, L.Y.; Feng, G.P.; Wang, T.X. Green synthesis of ZnO2 nanoparticles from hydrozincite and hydrogen peroxide at room temperature. Mater. Lett. 2010, 64, 1647–1649. [Google Scholar] [CrossRef]
  40. de Mello, L.B.; Varanda, L.C.; Sigoli, F.A.; Mazali, I.O. Co-precipitation synthesis of (Zn-Mn)-co-doped magnetite nanoparticles and their application in magnetic hyperthermia. J. Alloys Compd. 2019, 698–705. [Google Scholar] [CrossRef]
  41. Gonçalves, N.S.; Carvalho, J.A.; Lima, Z.M.; Sasaki, J.M. Size-strain study of NiO nanoparticles by X-ray powder diffraction line broadening. Mater. Lett. 2012, 72, 36–38. [Google Scholar] [CrossRef]
  42. Zak, A.K.; Majid, W.H.A.; Abrishami, M.E.; Yousefi, R. X-ray analysis of ZnO nanoparticles by Williamson-Hall and size-strain plot methods. Solid State Sci. 2011, 13, 251–256. [Google Scholar] [CrossRef]
  43. Bai, H.; Liu, X. Green hydrothermal synthesis and photoluminescence property of ZnO2 nanoparticles. Mater. Lett. 2010, 64, 341–343. [Google Scholar] [CrossRef]
  44. Fatehah, M.O.; Aziz, H.A.; Stoll, S. Stability of ZnO Nanoparticles in Solution. Influence of pH, Dissolution, Aggregation and Disaggregation Effects. J. Colloid Sci. Biotechnol. 2014, 3, 75–84. [Google Scholar] [CrossRef]
  45. Omar, F.M.; Aziz, H.A.; Stoll, S. Aggregation and disaggregation of ZnO nanoparticles: Influence of pH and adsorption of Suwannee River humic acid. Sci. Total Environ. 2013, 468, 195–201. [Google Scholar] [CrossRef]
  46. Berg, J.M.; Romoser, A.; Banerjee, N.; Zebda, R.; Sayes, C.M. The relationship between pH and zeta potential of ~30 nm metal oxide nanoparticle suspensions relevant to in vitro toxicological evaluations. Nanotoxicology 2009, 3, 276–283. [Google Scholar] [CrossRef]
  47. Zhang, Q.; Xu, M.; You, B.; Zhang, Q.; Yuan, H.; Ostrikov, K. Oxygen Vacancy-Mediated ZnO Nanoparticle Photocatalyst for Degradation of Methylene Blue. Appl. Sci. 2018, 8, 353. [Google Scholar] [CrossRef] [Green Version]
  48. Moon, D.R.; Ingham, T.; Whalley, L.K.; Seakins, P.W.; Baeza-Romero, M.T.; Heard, D.E. Production of HO2 and OH radicals from near-UV irradiated airborne TiO2 nanoparticles. Phys. Chem. Chem. Phys. 2019, 21, 2325–2336. [Google Scholar] [CrossRef] [PubMed]
  49. Zent, A.P.; Ichimura, A.S.; Quinn, R.C.; Harding, H.K. The formation and stability of the superoxide radical (O2-) on rock-forming minerals: Band gaps, hydroxylation state, and implications for mars oxidant chemistry. J. Geophys. Res. 2008, 113, 1–13. [Google Scholar] [CrossRef]
  50. Wu, W.; Lin, R.; Shen, L.; Liang, R.; Yuan, R.; Wu, L. Highly efficient visible-light-induced photocatalytic hydrogenation of nitrobenzene to aniline in water. RSC Adv. 2013, 3, 10894–10899. [Google Scholar] [CrossRef]
  51. Modi, K.B.; Natarajan, K.; Kathad, C.R.; Shah, S.J.; Raval, P.Y.; Pathak, T.K.; Meshiya, U.M.; Vyas, K.G.; Bajaj, H.C.; Tayade, R.J. Synthesis and characterization of ferrite-semiconductor nano composite for photocatalytic degradation of aqueous nitrobenzene solution. AIP Conf. Proc. 2016, 1728, 1–6. [Google Scholar] [CrossRef]
  52. Shahrezaei, F.; Mansouri, Y.; Zinatizadeh, A.A.L.; Akhbari, A. Photocatalytic degradation of aniline using TiO2 nanoparticles in a vertical circulating photocatalytic reactor. Int. J. Photoenergy 2012, 2012, 430638. [Google Scholar] [CrossRef] [Green Version]
  53. Flores, S.O.; Rios-Bernij, O.; Valenzuela, M.A. Photocatalytic reduction of nitrobenzene by titanium dioxide powder. Top. Catal. 2007, 44, 507–511. [Google Scholar] [CrossRef]
  54. Chong, M.N.; Jin, B.; Chow, C.W.K.; Saint, C. Recent developments in photocatalytic water treatment technology: A review. Water Res. 2010, 44, 2997–3027. [Google Scholar] [CrossRef]
  55. Jeong, S.; Lee, H.; Park, H.; Jeon, K.J.; Park, Y.K.; Jung, S.C. Rapid photocatalytic degradation of nitrobenzene under the simultaneous illumination of UV and microwave radiation fields with a TiO2 ball catalyst. Catal. Today 2018, 307, 65–72. [Google Scholar] [CrossRef]
  56. ElMetwally, A.E.; Eshaq, G.; Al-Sabagh, A.M.; Yehia, F.Z.; Philip, C.A.; Moussa, N.A.; ElShafei, G.M.S. Insight into heterogeneous Fenton-sonophotocatalytic degradation of nitrobenzene using metal oxychlorides. Sep. Purif. Technol. 2019, 210, 452–462. [Google Scholar] [CrossRef]
  57. Bergs, C.; Simon, P.; Prots, Y.; Pich, A. Ultrasmall functional ZnO2 nanoparticles: Synthesis, characterization and oxygen release properties. RSC Adv. 2016, 6, 84777–84786. [Google Scholar] [CrossRef]
  58. Bergs, C.; Brück, L.; Rosencrantz, R.R.; Conrads, G.; Elling, L.; Pich, A. Biofunctionalized zinc peroxide (ZnO2) nanoparticles as active oxygen sources and antibacterial agents. RSC Adv. 2017, 7, 38998–39010. [Google Scholar] [CrossRef] [Green Version]
  59. Karthik, R.; Govindasamy, M.; Chen, S.M.; Cheng, Y.H.; Muthukrishnan, P.; Padmavathy, S.; Elangovan, A. Biosynthesis of silver nanoparticles by using Camellia japonica leaf extract for the electrocatalytic reduction of nitrobenzene and photocatalytic degradation of Eosin-Y. J. Photochem. Photobiol. B Biol. 2017, 170, 164–172. [Google Scholar] [CrossRef]
  60. Lawrence, A.; Thollander, P.; Andrei, M.; Karlsson, M. Specific Energy Consumption/Use (SEC) in Energy Management for Improving Energy Efficiency in Industry: Meaning, Usage and Differences. Energies 2019, 12, 247. [Google Scholar] [CrossRef] [Green Version]
  61. Tayade, R.J.; Key, D.L. Synthesis and Characterization of Titanium Dioxide Nanotubes for Photocatalytic Degradation of Aqueous Nitrobenzene in the Presence of Sunlight. Mater. Sci. Forum. 2010, 657, 62–74. [Google Scholar] [CrossRef]
  62. Srilakshmi, C.; Saraf, R.; Shivakumara, C. Effective degradation of aqueous nitrobenzene using the SrFeO3−δ photocatalyst under UV illumination and its kinetics and mechanistic studies. Ind. Eng. Chem. Res. 2015, 54, 7800–7810. [Google Scholar] [CrossRef]
  63. Bhatkhande, D.S.; Pangarkar, V.G.; Beenackers, A.A.C.M. Photocatalytic degradation of nitrobenzene using titanium dioxide and concentrated solar radiation: Chemical effects and scaleup. Water Res. 2003, 37, 1223–1230. [Google Scholar] [CrossRef]
  64. Roy, P.; Periasamy, A.P.; Liang, C.T.; Chang, H.T. Synthesis of graphene-ZnO-Au nanocomposites for efficient photocatalytic reduction of nitrobenzene. Environ. Sci. Technol. 2013, 47, 6688–6695. [Google Scholar] [CrossRef] [PubMed]
  65. Reynoso-Soto, E.A.; Pérez-Sicairos, S.; Reyes-Cruzaley, A.P.; Castro-Riquelme, C.L.; Félix-Navarro, R.M.; Paraguay-Delgado, F.; Alonso-Núñez, G.; Lin-Ho, S.W. Photocatalytic degradation of nitrobenzene using nanocrystalline TiO2 photocatalyst doped with Zn ions. J. Mex. Chem. Soc. 2013, 57, 298–305. [Google Scholar] [CrossRef]
  66. Jo, W.K.; Won, Y.; Hwang, I.; Tayade, R.J. Enhanced photocatalytic degradation of aqueous nitrobenzene using graphitic carbon-TiO2 composites. Ind. Eng. Chem. Res. 2014, 53, 3455–3461. [Google Scholar] [CrossRef]
  67. Boxi, S.S.; Paria, S. Visible light induced enhanced photocatalytic degradation of organic pollutants in aqueous media using Ag doped hollow TiO2 nanospheres. RSC Adv. 2015, 5, 37647–37668. [Google Scholar] [CrossRef]
  68. Bhatia, K. Hydroxyl radical induced oxidation of nitrobenzene. J. Phys. Chem. 1975, 79, 1032–1038. [Google Scholar] [CrossRef]
  69. Eberhardt, M.K.; Yoshida, M. Radiation-induced homolytic aromatic substitution. I. Hydroxylation of nitrobenzene, chlorobenzene, and toluene. J. Phys. Chem. 1973, 77, 589–597. [Google Scholar] [CrossRef]
  70. Zhang, Y.; Zhang, K.; Dai, C.; Zhou, X.; Si, H. An enhanced Fenton reaction catalyzed by natural heterogeneous pyrite for nitrobenzene degradation in an aqueous solution. Chem. Eng. J. 2014, 244, 438–445. [Google Scholar] [CrossRef]
  71. Pérez-Sicairos, S.; Corrales-López, K.A.; Hernández-Calderón, O.M.; Salazar-Gastélum, M.I.; Félix-Navarro, R.M. Photochemical Degradation of Nitrobenzene by S2O8–2 Ions and UV Radiation. Rev. Int. Contam. Ambient. 2016, 32, 227–236. [Google Scholar] [CrossRef]
Figure 1. (A) UV-Vis spectrum of obtained ZnO2 nanoparticles (black line) dispersed in water and comparison with reactant zinc acetate (dotted line) used for synthesis. (B) Indirect bandgap calculation for ZnO2 nanoparticles.
Figure 1. (A) UV-Vis spectrum of obtained ZnO2 nanoparticles (black line) dispersed in water and comparison with reactant zinc acetate (dotted line) used for synthesis. (B) Indirect bandgap calculation for ZnO2 nanoparticles.
Catalysts 10 01041 g001
Figure 2. (A) Infrared Fourier transform (FT-IR) and (B) Raman spectrum for ZnO2 nanoparticles.
Figure 2. (A) Infrared Fourier transform (FT-IR) and (B) Raman spectrum for ZnO2 nanoparticles.
Catalysts 10 01041 g002
Figure 3. (A) X-ray diffraction (XRD), (B) Energy Dispersive X-ray spectroscopy (EDX) and (C) Scanning Electronic Microscope (SEM) image of ZnO2 nanoparticles. The insert in the SEM image corresponds to a size distribution histogram of the ~8 nm crystals constituting a ~100 nm in diameter particle.
Figure 3. (A) X-ray diffraction (XRD), (B) Energy Dispersive X-ray spectroscopy (EDX) and (C) Scanning Electronic Microscope (SEM) image of ZnO2 nanoparticles. The insert in the SEM image corresponds to a size distribution histogram of the ~8 nm crystals constituting a ~100 nm in diameter particle.
Catalysts 10 01041 g003
Figure 4. (A) SEM image with a gold coating and (B) hydrodynamic diameter (Dh) distribution by dynamic light scattering (DLS) analysis for ZnO2 nanoparticles. The insert in the SEM image corresponds to a size distribution histogram of the ~100 nm in diameter particles.
Figure 4. (A) SEM image with a gold coating and (B) hydrodynamic diameter (Dh) distribution by dynamic light scattering (DLS) analysis for ZnO2 nanoparticles. The insert in the SEM image corresponds to a size distribution histogram of the ~100 nm in diameter particles.
Catalysts 10 01041 g004
Figure 5. Effect of pH on the zeta potential measurement for ZnO2 nanoparticles.
Figure 5. Effect of pH on the zeta potential measurement for ZnO2 nanoparticles.
Catalysts 10 01041 g005
Figure 6. UV spectrum of aqueous nitrobenzene at different intervals for the photodegradation with ZnO2 nanoparticles at optimum conditions.
Figure 6. UV spectrum of aqueous nitrobenzene at different intervals for the photodegradation with ZnO2 nanoparticles at optimum conditions.
Catalysts 10 01041 g006
Figure 7. Percentage of nitrobenzene (NB) photodegradation by ZnO2 nanoparticles at different (A) photocatalyst concentrations, (B) different initial pH value with 30 ppm of photocatalyst and (C) specific energy consumption for each reaction at different pH values.
Figure 7. Percentage of nitrobenzene (NB) photodegradation by ZnO2 nanoparticles at different (A) photocatalyst concentrations, (B) different initial pH value with 30 ppm of photocatalyst and (C) specific energy consumption for each reaction at different pH values.
Catalysts 10 01041 g007
Figure 8. Gas chromatography (GC) chromatograms of the NB photocatalysis extracts for the reactions at pH: (A) 2, (B) 7 and (C) 10. The two peaks were detected to be NB and reaction intermediaries, nitrophenols (NPhs).
Figure 8. Gas chromatography (GC) chromatograms of the NB photocatalysis extracts for the reactions at pH: (A) 2, (B) 7 and (C) 10. The two peaks were detected to be NB and reaction intermediaries, nitrophenols (NPhs).
Catalysts 10 01041 g008
Figure 9. Mass spectrum (MS) for the detected molecules of the NB photocatalysis extracts. (A) NB, (B) 2-nitrophenol (2-NPh), (C) 3-nitrophenol (3-NPh) and (D) 4-nitrophenol (4-NPh) (NPs).
Figure 9. Mass spectrum (MS) for the detected molecules of the NB photocatalysis extracts. (A) NB, (B) 2-nitrophenol (2-NPh), (C) 3-nitrophenol (3-NPh) and (D) 4-nitrophenol (4-NPh) (NPs).
Catalysts 10 01041 g009
Table 1. X-ray diffraction calculated ZnO2 crystal structure parameters.
Table 1. X-ray diffraction calculated ZnO2 crystal structure parameters.
HKL JCPDSLattice Parameter (Å)Atomic Radius (Å)Cell Volume (Å3)
372004.862.43115
311114.912.45118
532204.872.43115
633114.872.43115
662224.872.43115
412104.892.44117
452114.882.44116
Table 2. Comparison of photocatalytic activity of ZnO2 with the reported various catalysts for NB degradation.
Table 2. Comparison of photocatalytic activity of ZnO2 with the reported various catalysts for NB degradation.
Material[Material][NB]Time
(h)
Type of IrradiationNB Degradation (%)Principal IntermediatesReference
TiO2 Nanotubes50 mg/50 mL40 ppm4Sun light100.00%No report[61]
SrFeO3−δ1000 ppm50 ppm6125 W, Mercury Vapor Lamp99.00%No report[62]
TiO **0.3% w/v300 ppm4Sun light98.50%3-nitrophenol and 4-nitrophenol[63]
TiO250 mg50 ppm8125 W, Mercury Vapor Lamp98.00%No report[10]
G-ZnO-Au NCs4 ppm5 mM2.3500 W,
<420 nm
97.80%Aniline[64]
(0.50%) Zn doped TiO2 15 ppm2254 nm97.00%No report[65]
GT-1200 ppm50 ppm4200 W96.00%No report[66]
Mercury Vapor Lamp
Ag-h-TiO2 ***0.5 g/L61.5 ppm3.5435.8 nm95.50%n-butanol, di-ethyl ether, furan, orthoformic acid, propanol and acetic acid.[67]
TiFe0.5 ***4 glass plates2.51 × 10−4 M4320–500 nm88.45%No report[22]
ZnO2 *30 ppm15 ppm2254 nm97.55%2-NPhs, 3-NPhs and 4-NPhsThis work
Note: G–ZnO–Au NCs = Graphene–ZnO–Au Nanocomposites, GT-1 = Graphitic Carbon (1%)-TiO2 Composites, Ag–h–TiO2 = Ag Doped Hollow TiO2 Nanospheres, TiFe0.5 = 0.5 wt.% Fe-doped TiO2 deposited as thin film on glass plates (19.5 cm2 irradiation area), Nitrophenol = NPhs, 2-nitrophenol = 2-NPhs, 3-nitrophenol = 3-NPhs and 4-nitrophenol = 4-NPhs.2.4 Degradation Products Identification by High-Resolution Mass Spectrometry. * Acid ** Alkaline *** Neutral (the reactions where the pH is not specified, are assumed to be aqueous solution).

Share and Cite

MDPI and ACS Style

Ramírez, J.I.D.L.; Villegas, V.A.R.; Sicairos, S.P.; Guevara, E.H.; Brito Perea, M.D.C.; Sánchez, B.L. Synthesis and Characterization of Zinc Peroxide Nanoparticles for the Photodegradation of Nitrobenzene Assisted by UV-Light. Catalysts 2020, 10, 1041. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10091041

AMA Style

Ramírez JIDL, Villegas VAR, Sicairos SP, Guevara EH, Brito Perea MDC, Sánchez BL. Synthesis and Characterization of Zinc Peroxide Nanoparticles for the Photodegradation of Nitrobenzene Assisted by UV-Light. Catalysts. 2020; 10(9):1041. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10091041

Chicago/Turabian Style

Ramírez, Jesús Isaías De León, Víctor Alfredo Reyes Villegas, Sergio Pérez Sicairos, Esteban Hernández Guevara, Mirna Del Carmen Brito Perea, and Bertha Landeros Sánchez. 2020. "Synthesis and Characterization of Zinc Peroxide Nanoparticles for the Photodegradation of Nitrobenzene Assisted by UV-Light" Catalysts 10, no. 9: 1041. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10091041

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