Next Article in Journal
Antibacterial Bio-Nanocomposite Textile Material Produced from Natural Resources
Next Article in Special Issue
Enhancement of Methylene Blue Photodegradation Rate Using Laser Synthesized Ag-Doped ZnO Nanoparticles
Previous Article in Journal
All-Dielectric Transreflective Angle-Insensitive Near-Infrared (NIR) Filter
Previous Article in Special Issue
AuAg Nanoparticles Grafted on TiO2@N-Doped Porous Carbon: Improved Depletion of Ciprofloxacin under Visible Light through Plasmonic Photocatalysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Characterization of N and Fe-Doped TiO2 Nanoparticles for 2,4-Dimethylaniline Mineralization

by
Emerson Faustino
1,2,†,
Thalita Ferreira da Silva
1,†,
Rebeca Fabbro Cunha
1,
Diego Roberto Vieira Guelfi
1,
Priscila Sabioni Cavalheri
1,3,
Silvio César de Oliveira
1,
Anderson Rodrigues Lima Caires
4,
Gleison Antonio Casagrande
1,
Rodrigo Pereira Cavalcante
1,5,* and
Amilcar Machulek Junior
1,*
1
Institute of Chemistry, Federal University of Mato Grosso do Sul (UFMS), Av. Senador Filinto Muller, 1555, CP 549, Campo Grande 79074-460, MS, Brazil
2
Federal Institute of Education, Science and Technology of Rondônia, Rodovia RO-257, s/n—Zona Rural, Ariquemes 76870-000, RO, Brazil
3
Department of Sanitary and Environmental Engineering, Dom Bosco Catholic University, Campo Grande 79117-900, MS, Brazil
4
Optics and Photonics Group, Institute of Physics, Federal University of Mato Grosso do Sul (UFMS), Campo Grande 79070-900, MS, Brazil
5
School of Technology, University of Campinas—UNICAMP, Paschoal Marmo, 1888, Limeira 13484-332, SP, Brazil
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2022, 12(15), 2538; https://doi.org/10.3390/nano12152538
Submission received: 2 July 2022 / Revised: 20 July 2022 / Accepted: 22 July 2022 / Published: 24 July 2022
(This article belongs to the Special Issue Semiconductor-Based Nanomaterials for Photocatalytic Applications)

Abstract

:
The present study aimed to evaluate the feasibility of developing low-cost N- and Fe-doped TiO2 photocatalysts for investigating the mineralization of 2,4-dimethylaniline (2,4-DMA). With a single anatase phase, the photocatalysts showed high thermal stability with mass losses of less than 2%. The predominant oxidative state is Ti4+, but there is presence of Ti3+ associated with oxygen vacancies. In materials with N, doping was interstitial in the NH3/NH4+ form and for doping with Fe, there was a presence of Fe-Ti bonds (indicating substitutional occupations). With an improved band gap energy from 3.16 eV to 2.82 eV the photoactivity of the photocatalysts was validated with an 18 W UVA lamp (340–415 nm) with a flux of 8.23 × 10−6 Einstein s−1. With a size of only 14.45 nm and a surface area of 84.73 m2 g−1, the photocatalyst doped with 0.0125% Fe mineralized 92% of the 2,4-DMA in just 180 min. While the 3% N photocatalyst with 12.27 nm had similar performance at only 360 min. Factors such as high surface area, mesoporous structure and improved Ebg, and absence of Fe peak in XPS analysis indicate that doping with 0.0125% Fe caused a modification in TiO2 structure.

1. Introduction

2,4-Dimethylaniline (2,4-DMA, also called 2,4-Xylidine, C.A.S. No. 95-68-1) is an aromatic amine used as a precursor for some textile dyes and veterinary drugs [1,2]. Considered a persistent contaminant, 2,4-DMA has already been detected in industrial effluents and groundwater. For example, 2,4-DMA was detected in textile industry wastewater even after anaerobic treatment processes [3]. Toxicologically, 2,4-DMA has been classified as a group 3 carcinogen and has adverse effects on the aquatic ecosystem [4]. Furthermore, 2,4-DMA is a model compound used in degradation studies because its oxidation produces intermediates that facilitate mechanistic studies [5,6].
Persistent contaminants exhibit chemical and biological resistance to mineralization through traditional water treatment methods. Therefore, the search for innovative treatment technologies has increased. Advanced oxidation processes (AOPs) are an alternative to completely mineralizing for numerous contaminants, as it is not a selective process [7]. AOPs are able to remove various types of pollutants from attacks by reactive oxygen species (ROS) generated in the system [8,9,10,11,12,13,14,15,16]. Heterogeneous photocatalysis (PC) are AOPs that involves the activation of a photocatalyst by light irradiation. Under specific conditions, photogenerated charge carriers (eCB and h+VB) undergo interfacial transfer and promote redox reactions with adsorbed molecular oxygen, water molecules, and hydroxide ions, producing various types of ROS such as superoxide, hydrogen peroxide, and hydroxyl radicals (HO) [13,17]. Many semiconductors, such as TiO2 [18], WO3 [14], ZnO [19], and BiVO4 [20], have been used to photodegrade emerging contaminants. TiO2 is the most investigated photocatalyst due to its low toxicity, low cost, chemical stability over a wide pH range, thermal stability, minimizing eCB/h+VB pair recombination, and the band gap energy (Ebg) being able to oxidize water molecules generating HO radicals. The principle of PC involves the activation of a semiconductor by solar or artificial irradiation, Equation (1). Photons absorbed with energy greater than the Ebg promote an electron from the valence band to the conduction band with the simultaneous generation of a hole (h+VB) in the valence band. The adsorbed water (H2Oads.) is oxidized in the h+VB generating HO, which can subsequently oxidize the organic contaminant (R), Equations (2) and (3). Or the R can be oxidized directly in the h+VB, Equation (4).
TiO2 + hυ → TiO2 (eCB + h+VB)
h+VB + TiO2-H2Oads. → HO + H+
HO + R → CO2 + H2O
h+VB + R → CO2 + H2O
TiO2 has a high Ebg (3.2 eV for anatase) and is excited only by ultraviolet light irradiation (λ < 385 nm) [21]. The photocatalytic activity of TiO2 can be improved by doping or modifying its crystalline structure in order to increase the optical absorption of radiation in the visible range and decrease the rate of recombination of photogenerated charges [22]. Light absorption of TiO2 in the visible region can be improved via metal- [23,24,25] or non-metal-doping [18,26,27]. Several approaches for the synthesis of TiO2 have been reported, such as sol-gel, hydrothermal, solvothermal, and spray pyrolysis [28,29,30,31]. Among them, the sol-gel method has been considered a good solution to produce nanomaterials, presenting advantages such as homogeneity, ease and flexibility in introducing doping at high concentrations, simple synthesis process, and providing nanomaterials with high purity. During sol-gel preparation, a TiO2 sol or gel is formed by precipitation through the hydrolysis and condensation of a titanium alkoxide [18]. Jadhav et al. synthesized N-doped TiO2 doped with 25% N using urea as a precursor. The material synthesized by them degraded 70% of the methylene blue in 180 min of the experiment under visible light [32]. Synthesis using ammonium salt as precursor has also been reported and its photocatalytic activity showed a 90% removal in 40 min of rhodamine B [33]. Yalçın et al. synthesized Fe-doped photocatalysts and tested the photoactivity on 4-nitrophenol degradation degrading 80% in 120 min [34].
In this work, we synthesized and characterized N-, Fe-, and N-Fe-doped TiO2 nanoparticles via sol-gel. Furthermore, we evaluated the photocatalytic activities of TiO2 NPs through 2,4-DMA mineralization. This research provides new results on the influence of nitrogen and iron on TiO2 on 2,4-DMA mineralization. It is worth mentioning that, to the best of our knowledge, no studies were reported involving the photocatalysis process with N-, Fe- doped TiO2 for the mineralization of 2,4-DMA, with good results of mineralization rates in 6 h in an unprecedented manner.

2. Materials and Methods

2.1. Synthesis of Photocatalysts

The synthesis of photocatalysts via sol-gel followed the procedure described below [18,35]. A volume of 19.10 mL of titanium IV isopropoxide (97%, Sigma Aldrich, St. Louis, MO, USA) and 16.10 mL of glacial acetic acid (99.7%, Alphatec, Carlsbad, CA, USA) was added to a beaker and left under magnetic stirring. After homogenization, 19.10 mL of 2-propanol (99,6%, Merck, Kenilworth, NJ, USA) was added and left under stirring for 60 min. Thereupon, 30 mL of deionized water (resistivity > 18 MΩ cm obtained from Gehaka DG500 UF system, São Paulo, SP, Brazil) acidified with 1 mL of nitric acid (65.0%, Vetec, Duque de Caxias, RJ, Brazil) was added dropwise to the above solution. At the end of the addition, the solution was stirred for another 120 min. Finally, the resulting product was maintained at 40 °C for about 48 h for gel formation. The gel was dried in an oven at 100 °C for 24 h. The product has been macerated and placed in a muffle furnace at 450 °C for 4 h.
The synthesis of TiO2 doped with N, Fe, and N-Fe followed the same procedure described above, but there was the addition of ammonium hydroxide (25%, Merck, Kenilworth, NJ, USA) as a source of N and iron III nitrate (98%, Vetec) as a precursor of Fe. N-doped TiO2 were prepared in the different weight content of nitrogen (%): 1.0, 2.0, 3.0, 4.0, 5.0, 6.0, 7.0, and 9.0 (w/w). Fe-doped TiO2 were prepared in the different weight content of iron (%): 0.0125, 0.025, 0.05, 0.1, 0.25, and 1.0 (w/w). N-Fe co-doped TiO2 were prepared by fixing 0.025% Fe and varying only N by 2, 3, and 4% (w/w). These concentrations of N and Fe were chosen based on previous work [18,35]. The effects of the presence of N and Fe on the physical and chemical properties of photocatalyst particles were investigated by increasing or decreasing the photocatalytic response in 2,4-DMA mineralization. Figure 1 presents a schematic representation showing the steps of the synthesis.

2.2. Photocatalysts Characterization

The nanoparticles synthesized from pure TiO2, N, and Fe-doped TiO2 are characterized by thermogravimetric (TG) and differential thermogravimetric (DTG) curves recorded by the TA Instruments thermal analyzer-TGA Q50. Approximately 10 mg of each sample was used and the thermocurves are recorded in the range of room temperature to 850 °C in atmosphere of nitrogen (N2) gas. The crystalline structure of the synthesized pure TiO2, N, and Fe-doped TiO2 nanoparticles is carried out by powder X-ray diffraction (XRD) technique employing LabX XRD-6100, Shimadzu. The XRD analysis is done in the 2θ range of 10° and 70° using Co Kα radiation of wavelength 17,889 Å. Fourier transform infrared spectroscopy (FTIR) analysis of the samples was performed with a PerkinElmer FTIR spectrophotometer, Frontier model. The samples were prepared by the KBr pellet pressing method, at room temperature and in the region of 4000–450 cm−1. The morphological of the materials was studied by Scanning Electron Microscopy (SEM) in a JSM-6380LV JEOL microscope and Transmission Electron Microscopy (TEM) in a model JEM 2100 LaB6 JEOL microscope. Using the ImageJ software and TEM images, the diameter of the particles in the different samples was measured [18]. Diffuse UV-visible reflectance spectroscopy (DRS) analyzes were performed on the samples using a PerkinElmer Lambda 650 UV/Vis/NIR spectrophotometer operating in the range of 200–800 nm. From the DRS studies, a graph of the modified Kubelka-Munk function [F(R) × hν]1/2 versus absorbed light energy was constructed for indirect band gap energy (Ebg) calculations. The surface area and pore size distribution of the samples were calculated by the Brunauer-Emmett-Teller (BET) method using the ASAP 2010, Micromeristic surface area analyzer. X-ray photoelectron spectroscopy (XPS) measurements were performed on a EA125 Sphera spectrometer. The elemental composition of the samples was determined by X-ray photoelectron spectroscopy (XPS). The analyzes were performed on an Omicron, model EA125 Sphera equipped with a hemispherical electron analyzer. All measurements were performed in an ultra-high vacuum (UHV) chamber with pressure between 10−8 and 10−12 mbar. The C1s peak of carbon contamination at 284.60 eV was used as a binding energy reference.

2.3. Photocatalytic Mineralization

The photocatalytic activity of the synthesized materials was tested for the mineralization of 2,4-DMA. In a typical procedure, 350 mL (0.1 mmol L−1) of 2,4-DMA solution and 175 mg of photocatalyst were transferred to a glass reactor. The suspension was kept under constant stirring in the dark for 30 min, ensuring the adsorption-desorption equilibrium of the test molecule. A UVA lamp (18 W, DULUX L-OSRAM, λ = 340–415 nm and 8.23 × 10−6 Einstein s−1) was used as irradiation. After the adsorption-desorption equilibrium was established, the lamp was turned on, and 4 mL aliquots were collected at different times. The mineralization was followed by the total organic carbon (TOC) analysis using a Shimadzu TOC VCPN analyzer. The calibration accuracy values with LOQ = 0.180 mg L−1 and LOD = 0.053 mg L−1 were obtained. The reuse experiments lasted 180 min and various aliquots were collected and analyzed by TOC. The photocatalyst used was collected by vacuum filtration, washed with deionized water (resistivity > 18 MΩ cm obtained from Gehaka DG500 UF system, São Paulo, Brazil) and dried in an oven at 80 °C (Medicate, MD 1.2). The reuse experiment cycle was repeated three times.
Apparent pseudo-first order kinetic constants (k2,4-DMA) (Equation (5)) to 2,4-DMA were calculated according to the following:
k 2 , 4 - DMA = l n [ TOC ] 0 [ TOC ] t t        

3. Results and Discussion

3.1. Characterization of N-, Fe-, and N-Fe Doped TiO2 Photocatalysts

Figure S1 (see Supplementary Material) shows the results of the TG and DTG analysis for the synthesized N-, Fe-, and N-Fe doped TiO2 nanoparticles. It can be seen in Figure S1A, that the synthesized materials have a high thermal stability independent of doping. The mass losses were less than 3%, which is consistent with previous works [25,36,37]. In the DTG curves (Figure S1B), an endothermic event before 100 °C can be seen and is associated with evaporation of adsorbed water and/or decomposition of residual organic matter.
Some characteristic peaks of TiO2-based photocatalysts can be seen in Figure 2A. The bands at 700 cm−1 are typical of the Ti-O and Ti-O-Ti bonds present in TiO2. The bands at 1650 cm−1 and 3200 cm−1 are associated with the O-H vibrational modes of adsorbed water. The N-doped photocatalysts present bands close to 1500 cm−1 and 3200 cm−1, referring to the vibration of the N-H group [32]. The vibrational bands of the Fe-O bond are in the range of 400 to 700 cm−1, but they were not observed due to the low concentration of iron used.
Figure 2B shows XRD patterns for the investigated photocatalysts. The anatase phase peaks at 29.54° (101) and 44.14° (200) compared to ICDD Card no. 00–021–1272, are present in the samples. The peaks of the rutile phase are absent in the diffractograms. The calcination temperature (450 °C) does not promote the formation of this crystalline phase [38]. The other peaks at 56.36° (105), 63.50° (116), 64.96° (200), and 74.44 (2015) are also due to the anatase phase belonging to the tetragonal crystal system [13]. From the Rietveld refinement (see Figure S2 and Table S1 in the Supplementary Material) the special group I41/amd (141) and a density of 3.90 g cm−3 for TiO2, 0.125%Fe-TiO2 and 3%N-TiO2. The presence of the anatase phase in the photocatalysts is consistent with the literature [25,39].
Obtaining the anatase phase is satisfactory, as it allows a more effective absorption of photons than the rutile phase of TiO2. The anatase phase of TiO2 increases the photocatalytic performance compared to the rutile phase as it increases the redox power of the charge carriers and can enhance the adsorb of hydroxyl groups on its surface [40]. In the case of Fe doping, the absence of peaks at 31.07° corresponding to Fe3O4 and 36.18° and 41.30° corresponding to FeTiO3 in Figure 2B indicates that Fe3+ cations did not react with TiO2 [34,41]. The absence of these oxides is desirable because their presence reduces photocatalytic activity [34]. Therefore, substitutional doping of Fe3+ occurred in the photocatalysts instead of interstitial doping.
The morphology and size of N-, Fe-, and N-Fe doped TiO2 photocatalysts were investigated by SEM (Figure 3) and TEM (Figure 4). According to Figure 3A–G, the morphology of the synthesized oxides does not follow a defined pattern, presenting clusters in a slightly rounded shape on their surface. In some cases, the clusters are larger than others, as seen in the A (pure), D (3% N), E (9% N), and F (2% N + 0.025% Fe) images in Figure 3.
Figure 4 shows the TEM micrographs and the particle diameter distribution histogram. The data show that the TiO2 photocatalysts have an average particle diameter of 16.16 nm (Figure 4A), while 0.0125% Fe (Figure 4B) and 3% N (Figure 4C) have 14.45 nm and 12.27 nm, respectively. TiO2 doping resulted in particle reduction, regardless of the dopant used.
N2 adsorption-desorption isotherm at liquid nitrogen temperature was conducted to analyze the pore size and surface area of N-, Fe-, and N-Fe doped TiO2 photocatalysts, as shown in Figure 5. BET surface areas (SBET), pore volume (Vp) and pore diameters (Dp) are shown in Table 1. The samples show surface areas of 78.67, 84.73, and 95.70 m2 g−1 for TiO2, 0.0125% Fe, and 3% N, respectively. In Figure 4, we clearly see that the samples are agglomerates formed by monodisperse primary particles, with sizes of 16.16, 14.45, and 12.27 nm for TiO2, 0.0125% Fe, and 3% N, respectively. Table 1 indicates that the nanoparticles are porous with mean pore diameters of 9.7 to 9.0 nm and narrow pore size distributions (Figure 5). Farhangi et al. reported that Fe doping could decrease TiO2 crystallization and also slightly restrict TiO2 crystallite growth [42]. The presence of N- and Fe- caused a decrease in the nanoparticle and an increase in its surface area. Large surface area and mesoporous structure are favorable characteristics for obtaining high photocatalytic activity.
The chemical composition and electronic states of the photocatalyst constituents: pure TiO2, 0.0125% Fe and 3% N were investigated by XPS, as shown in Figure 6. According to the survey spectrum Ti, O, and C peaks are evident. For all samples can observed the presence of peaks related to Ti 3p, Ti 3s, C 1s, Ti 2p, O 1s, and Ti 2s at around 26.0, 39.0, 286.0, 462.0, 533.0, and 566.0 eV, respectively. The origin of carbon is common in XPS analyses, resulting from the strong adsorption of a contaminating layer or hydrocarbon from the carbon strip of the instrument [35,43]. The chemical composition of photocatalysts is shown in Table 2. As seen in Figure 7A–C, two peaks at approximately 458.5 and 464.3 eV that correspond to typical Ti4+ 2p3/2 and Ti4+ 2p1/2, are observed for all samples [34,44]. In Figure 7A–C, the Ti 2p3/2 peak at 457.1 eV and the Ti 2p1/2 peak at 462.8 eV correspond to the Ti3+ species [45]. The O 1s signals for all samples (Figure 7D–F) can be divided into two peaks at 529.8 and 531.0 eV, corresponding to Ti-O and surface −OH bonds, respectively [44]. For comparison purposes, the contribution of the peaks at 531.0 eV at 0.0125% Fe (24.9%) and 3%N (17.1%) is lower than that of TiO2 (42.0%), indicating fewer oxygen vacancies in the doped catalysts. Oxygen vacancies on the surface of the particle tend to promote more −OH bonds [46]. Metallic Fe signals were not observed at 710.7 eV for 2p3/2 and 724.3 eV for 2p1/2, which indicates the absence of Fe2O3 [34]. As the presence of metallic Fe was not detected, the iron cations probably occupied substitutional positions due to the similar radius of Fe3+ (0.64 Å) and Ti4+ (0.68 Å), forming a solid Fe-Ti solution during the sintering process (<1 at Fe%) [47]. The peak at 399.1 eV in Figure S3 refers to the N 1s that are present in the 3% N sample, this peak is characteristic of interstitial N. Values above 400 eV can be assigned to groups NO2-, N2 and NHx. As shown by the results obtained from FTIR (Figure 2A) there are vibrations in the N-H group, so it can be said that the peak of N 1s is nitrogen in the NH3/NH4+ form. It is considered that nitrogen does not replace the oxygen atom in the crystal matrix, it was incorporated to form Ti–O–N bonds on the surface of the material [25,48].
The UV-visible absorption spectra of N-, Fe-, and N-Fe doped TiO2 photocatalysts are presented in Figure 8. The absorption spectra in Figure 8A shows intense absorption in the region of 300 to 450 nm, resulting from the intrinsic band gap of TiO2. The 1% Fe photocatalyst showed improved absorption in the visible range (in the 400 to 700 nm) when compared to other doped photocatalysts. The Ebg (Figure 8B) of the samples was estimated using Tauc’s formula and their values are found in Table 3. Pure TiO2 had an Ebg of 3.16 eV, a value close to that reported in the literature [18]. The Ebg values for N-doped catalysts showed a small difference in relation to pure TiO2. The likely reason behind the decrease in the Ebg is that the nitrogen inside the TiO2 doped material replaces the Ti ions inside the material [32], as seen in Figure 8. For the Fe-doped photocatalysts, a gradual decrease is observed with increasing Fe concentration. This effect may be associated with the new energy levels introduced in the TiO2 band gap caused by Fe3+ and Fe2+ ions. The 1% Fe photocatalyst presented an Ebg of 2.82 eV. This value is 0.34 eV lower than the band gap of pure TiO2 anatase nanoparticles (3.16 eV), which can be caused by Ti3+ sites (vacancies of oxygen) [49] as a role of the intermediate trap state formed in the calcination process. Barkhade and Banerjee also detected a decrease in Ebg from 3.4 eV for pure TiO2 to 2.5 eV for a TiO2 sample doped with 1.0% Fe (mol %). They attributed the decrease to the insertion of Fe3+ ions into the TiO2 lattice in which conduction band overlap occurred due to the Ti (d-orbital) of TiO2 and the metal (d-orbital) orbital of Fe3+ ions. Furthermore, the doping of Fe3+ ions induces the formation of new electronic states (Fe4+ and Fe2+) that span across the TiO2 band gap [50].
The photocatalytic efficiency of a catalyst depends on some aspects such as light absorption, charge separation and types of catalysts. The fundamental reaction in photocatalysis occurs with the absorption of light [51], which leads to charge separation and formation of active species, as shown in Equations (1)–(4). The charge carriers generated can separate and then recombine. But they can also migrate to the surface of the catalyst and recombine in surface traps or undergo interfacial electron transfer with the reduced species. The recombination kinetics is related to the presence of impurities in the bulk and recombination centers on the catalyst surface. Therefore, the quantum yield of a doped catalyst is often lower than that of the related undoped material. However, light capture plays an important role in determining the rate of degradation of organic molecules. Consequently, here the improved light-gathering capacity of N-, Fe- and N-Fe-doped TiO2 photocatalysts would be beneficial to increase their photocatalytic activity.

3.2. Photocatalytic Activity

The photocatalytic activity of N-, Fe-, and N-Fe doped TiO2 was measured using a 2,4-DMA solution under UVA light. The effects of nitrogen content (1, 2, 3, 4, 5, 6, 7, 9%), iron (0.0125, 0.025, 0.05, 0.1, 0.25, and 1%) and N-Fe co-doping (3% N +0.025% Fe, and 4% N + 0.025% Fe) were investigated by TOC (Total Organic Carbon) with their respective pseudo-first-order constants (k2,4-DMA).
Figure 9A shows the relationship between 2,4-DMA mineralization and irradiation time for each x% N photocatalyst during PC, and in Figure 9B their respective k2,4-DMA. In 360 min of the experiment, the 3% N and 4% N photocatalysts mineralized 100% and presented a k2,4-DMA of 0.007 min−1 and 0.006 min−1, respectively. Furthermore, the values of k2,4-DMA were 1.8 and 1.5 times higher than mineralization with pure TiO2 (Table 4). Analyzing the experiment time at 240 min, we see that increasing the percentage of nitrogen up to 3% (w/w) promotes an increase in mineralization.
The results also revealed that the increase above 3% of N did not improve the mineralization performance, presenting similar results to those observed for pure TiO2. The higher mineralization rate for the 3% N photocatalyst is the result of the increase in surface area (78.66 m2 g−1) in relation to pure TiO2 (95.69 m2 g−1). It was also observed that the pore diameter was reduced from 9.72 nm (pure TiO2) to 9.03 nm (3% N), and the mean particle diameter obtained by the MET results was reduced from 16.16 nm to 12.27 nm for 3% N. With the improved contact surface, more ROS can be generated thus increasing photocatalytic activity [18,35].
Figure 9C illustrates the effect of iron content on x% Fe photocatalytic activity within 360 min reaction time. It is observed that as the doping content increases, the photocatalytic activity decreases. This decrease in photoactivity is associated with the fact that Fe3+ traps both eCB and h+VB up to a certain amount; as the doping content exceeds this ideal amount, it acts as a recombination center for the photogenerated charge carriers decreasing photocatalytic activity [47,52]. Thus, the 0.0125% Fe and 0.025% Fe photocatalysts mineralized 100% of 2,4-DMA in just 240 min. Furthermore, the values of k2,4-DMA 0.010 min−1 for 0.0125% Fe and 0.025% Fe (Figure 9D, Table 4) were 2.6 times higher than mineralization with pure TiO2. In the presence of low levels of Fe doping, the charge carriers are well separated, thus increasing the efficiency of the photocatalyst. The stability of the 0.0125%Fe photocatalyst was performed and in three successive cycles, there was a reduction of only 6%. This reduction indicates good stability of the studied photocatalyst. Fe3+ can serve as photogenerated hole chaser (Equation (1)). Fe4+ reacts with hydroxyl ions adsorbed on the surface to produce HO radicals (Equations (6) and (7)). Alternatively, Fe4+ can also react with photogenerated electrons, improving photocatalytic activity (Equation (8)). The trapped holes would oxidize hydroxyls adsorbed on the surface to produce hydroxyl radicals [47,53].
Fe3+ + h+VB → Fe4+
Fe4+ + OH(ads.) → Fe3+ + HO
Fe4+ + eCB → Fe3+
Photocatalysts co-doped with N- and Fe- were synthesized by fixing 0.025% Fe and varying by 3 and 4% N, such amounts of N- and Fe- were selected from the photocatalytic activities reported here. Interestingly, the junction of N- and Fe- did not show good results, as can be seen in Figure S4A (see Supplementary Material). With co-doping between N- Fe- only 50% of the 2,4-DMA was mineralized and the k2,4-DMA value was two times lower than that of pure TiO2 (Table 4 and Figure S4B). The increase in the surface area of the co-doped photocatalysts may have occurred, which contributed to the lack of the desired synergistic effect and to improve the photocatalytic activity. Yuan et al. reported the synthesis of N-Fe co-doped TiO2 photocatalysts and tested their photocatalytic activity on 2-methylisoborneol degradation. Using a xenon lamp the highest removal efficiency was achieved when the catalyst was co-doped with 0.001% Fe and 0.5% N, with a 2-methylisoborneol removal efficiency of 99.78% in 240 min. They observed that concentrations higher than 5% of N decrease the photoactivity due to an increase in the surface area of the catalysts [54].
Table 5 shows the mineralization results of N- Fe-doped TiO2 obtained in the present investigation and results of previously published studies [51,52,53,54,55,56]. For a better comparison, some parameters such as photocatalyst type, experimental conditions, model molecule for PC, light source conditions, and mineralization rate were also summarized in Table 5. In the literature, the mineralization rates found are achieved using a low volume of solution and a large amount of catalyst.

4. Conclusions

This study provided an understanding between N- and Fe-doped photocatalysts in the mineralization of 2,4-DMA. The main results found were that:
  • Nanoparticles from 16.16 to 12.27 nm were synthesized by the sol-gel method;
  • The crystalline phase was exclusively anatase with a density of 3.90 g cm−3;
  • The decrease in surface area from 97.70 to 78.67 m2 g−1 was observed to have mesopores;
  • Doping with Fe at the investigated concentrations promoted substitutional positions, increasing the formation of Fe-Ti bonds. For nitrogen, there was the presence of NH3/NH4+ species indicating interstitial doping;
  • The 2,4-DMA mineralization was viable with high removal rates.
Therefore, our findings offer the opportunity to reconsider studies taking into account the mineralization of emerging pollutants. This avoids the use of expensive chromatographic techniques and toxicity studies as mineralization converts the pollutant into CO2, H2O, and inorganic anions.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12152538/s1. Figure S1: Thermal analysis. (A) TG, and (B) DTG; Table S1: Rietveld refinement parameters for photocatalysts; Figure S2: XRD with Rietveld refinement for TiO2 (A), 0.125% Fe (B) and 3% N (C); Figure S3: High-resolution XPS spectra of N 1s; Figure S4: Results for 2,4-DMA mineralization using co-doped photocatalysts fixing 0.025%Fe and varying N doping by 3 and 4%.

Author Contributions

Conceptualization, E.F., T.F.d.S., R.P.C. and A.M.J.; Formal analysis, E.F., T.F.d.S., P.S.C., R.F.C., D.R.V.G., S.C.d.O. and R.P.C.; Funding acquisition, A.M.J.; Investigation, E.F. and T.F.d.S.; Methodology, E.F. and T.F.d.S.; Project administration, A.M.J.; Resources, A.R.L.C. and G.A.C.; Supervision, R.P.C. and A.M.J.; Writing—original draft, T.F.d.S.; Writing—review & editing, S.C.d.O., A.R.L.C., G.A.C., R.P.C. and A.M.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Brazilian funding agencies: Conselho Nacional de Desenvolvimento Científico e Tecnológico, CNPq; Coordenação de Aperfeiçoamento de Pessoal de Nível Superior, CAPES (Finance Code 001), Fundação de Amparo à Pesquisa do Estado de São Paulo, FAPESP (grant number #2019/26210-8) and Fundação de Apoio ao Desenvolvimento do Ensino, Ciência e Tecnologia do Estado de Mato Grosso do Sul, Fundect.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank MULTILAM/UFMS by SEMs analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, Y.; Li, H.; Tang, Y.; Luo, P.; Yang, W.; Wu, Y.; Yang, F.; Xiong, J. Aeromonas hydrophila-derived BioMnOx activates peroxymonosulfate for 2,4-dimethylaniline degradation in water: Mechanisms and catalyst reusability. Process Saf. Environ. Prot. 2022, 158, 308–319. [Google Scholar] [CrossRef]
  2. Amat, A.M.; Arques, A.; Bossmann, S.H.; Braun, A.M.; Göb, S.; Miranda, M.A.; Oliveros, E. Oxidative degradation of 2,4-xylidine by photosensitization with 2,4,6-triphenylpyrylium: Homogeneous and heterogeneous catalysis. Chemosphere 2004, 57, 1123–1130. [Google Scholar] [CrossRef] [PubMed]
  3. Jayapal, M.; Jagadeesan, H.; Krishnasamy, V.; Shanmugam, G.; Muniyappan, V.; Chidambaram, D.; Krishnamurthy, S. Demonstration of a plant-microbe integrated system for treatment of real-time textile industry wastewater. Environ. Pollut. 2022, 302, 119009. [Google Scholar] [CrossRef] [PubMed]
  4. Li, J.; He, C.; Cao, X.; Sui, H.; Li, X.; He, L. Low temperature thermal desorption-chemical oxidation hybrid process for the remediation of organic contaminated model soil: A case study. J. Contam. Hydrol. 2021, 243, 103908. [Google Scholar] [CrossRef] [PubMed]
  5. Machulek, A.; Gogritchiani, E.; Moraes, J.E.F.; Quina, F.H.; Braun, A.M.; Oliveros, E. Kinetic and mechanistic investigation of the ozonolysis of 2,4-xylidine (2,4-dimethyl-aniline) in acidic aqueous solution. Sep. Purif. Technol. 2009, 67, 141–148. [Google Scholar] [CrossRef]
  6. Bossmann, S.H.; Wörner, M.; Pokhrel, M.R.; Baumeister, B.; Göb, S.; Braun, A.M. Ruthenium(II)-tris-bipyridine/titanium dioxide codoped zeolite Y photocatalyst: Performance optimization using 2,4-xylidine (1-amino-2,4-dimethyl-benzene). Sep. Purif. Technol. 2009, 67, 201–207. [Google Scholar] [CrossRef]
  7. Zeshan, M.; Bhatti, I.A.; Mohsin, M.; Iqbal, M.; Amjed, N.; Nisar, J.; AlMasoud, N.; Alomar, T.S. Remediation of pesticides using TiO2 based photocatalytic strategies: A review. Chemosphere 2022, 300, 134525. [Google Scholar] [CrossRef]
  8. Friedrich, L.C.; Mendes, M.A.; Silva, V.O.; Zanta, C.L.P.S.; Machulek, A., Jr.; Quina, F.H. Mechanistic implications of zinc(II) ions on the degradation of phenol by the fenton reaction. J. Braz. Chem. Soc. 2012, 23, 1372–1377. [Google Scholar] [CrossRef] [Green Version]
  9. Luna, A.J.; Chiavone-Filho, O.; Machulek, A.; de Moraes, J.E.F.; Nascimento, C.A.O. Photo-Fenton oxidation of phenol and organochlorides (2,4-DCP and 2,4-D) in aqueous alkaline medium with high chloride concentration. J. Environ. Manag. 2012, 111, 10–17. [Google Scholar] [CrossRef]
  10. da Rosa, A.P.P.; Cavalcante, R.P.; da Silva, D.A.; da Silva, L.D.M.; da Silva, T.F.; Gozzi, F.; McGlynn, E.; Brady-Boyd, A.; Casagrande, G.A.; Wender, H.; et al. H2O2-assisted photoelectrocatalytic degradation of Mitoxantrone using CuO nanostructured films: Identification of by-products and toxicity. Sci. Total Environ. 2019, 651, 2845–2856. [Google Scholar] [CrossRef]
  11. de Melo da Silva, L.; Pereira Cavalcante, R.; Fabbro Cunha, R.; Gozzi, F.; Falcao Dantas, R.; de Oliveira, S.C.; Machulek, A. Tolfenamic acid degradation by direct photolysis and the UV-ABC/H2O2 process: Factorial design, kinetics, identification of intermediates, and toxicity evaluation. Sci. Total Environ. 2016, 573, 518–531. [Google Scholar] [CrossRef] [PubMed]
  12. Funai, D.H.; Didier, F.; Giménez, J.; Esplugas, S.; Marco, P.; Machulek, A. Photo-Fenton treatment of valproate under UVC, UVA and simulated solar radiation. J. Hazard. Mater. 2017, 323, 537–549. [Google Scholar] [CrossRef] [PubMed]
  13. Popescu, T.; Oktaviani Matei, C.; Culita, D.C.; Maraloiu, V.-A.; Rostas, A.M.; Diamandescu, L.; Iacob, N.; Savopol, T.; Ilas, M.C.; Feder, M.; et al. Facile synthesis of low toxicity iron oxide/TiO2 nanocomposites with hyperthermic and photo-oxidation properties. Sci. Rep. 2022, 12, 6887. [Google Scholar] [CrossRef] [PubMed]
  14. Yadav, A.A.; Hunge, Y.M.; Kang, S.W. Porous nanoplate-like tungsten trioxide/reduced graphene oxide catalyst for sonocatalytic degradation and photocatalytic hydrogen production. Surf. Interfaces 2021, 24, 101075. [Google Scholar] [CrossRef]
  15. Guelfi, D.R.V.; Brillas, E.; Gozzi, F.; Machulek, A.; de Oliveira, S.C.; Sirés, I. Influence of electrolysis conditions on the treatment of herbicide bentazon using artificial UVA radiation and sunlight. Identification of oxidation products. J. Environ. Manag. 2019, 231, 213–221. [Google Scholar] [CrossRef]
  16. Guelfi, D.R.V.; Gozzi, F.; Machulek, A., Jr.; Sirés, I.; Brillas, E.; de Oliveira, S.C. Degradation of herbicide S-metolachlor by electrochemical AOPs using a boron-doped diamond anode. Catal. Today 2018, 313, 182–188. [Google Scholar] [CrossRef]
  17. Park, S.H.; Katoch, A.; Chae, K.H.; Gautam, S.; Miedema, P.; Cho, S.W.; Kim, M.; Wang, R.-P.; Lazemi, M.; de Groot, F.; et al. Direct and real-time observation of hole transport dynamics in anatase TiO2 using X-ray free-electron laser. Nat. Commun. 2022, 13, 2531. [Google Scholar] [CrossRef]
  18. Cavalcante, R.P.; Dantas, R.F.; Bayarri, B.; González, O.; Giménez, J.; Esplugas, S.; Machulek, A. Synthesis and characterization of B-doped TiO2 and their performance for the degradation of metoprolol. Catal. Today 2015, 252, 27–34. [Google Scholar] [CrossRef]
  19. 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]
  20. Venkatesan, R.; Velumani, S.; Ordon, K.; Makowska-Janusik, M.; Corbel, G.; Kassiba, A. Nanostructured bismuth vanadate (BiVO4) thin films for efficient visible light photocatalysis. Mater. Chem. Phys. 2018, 205, 325–333. [Google Scholar] [CrossRef]
  21. Hunge, Y.M.; Yadav, A.A.; Kang, S.W.; Kim, H. Photocatalytic degradation of tetracycline antibiotics using hydrothermally synthesized two-dimensional molybdenum disulfide/titanium dioxide composites. J. Colloid Interface Sci. 2022, 606, 454–463. [Google Scholar] [CrossRef] [PubMed]
  22. Hunge, Y.M.; Yadav, A.A.; Khan, S.; Takagi, K.; Suzuki, N.; Teshima, K.; Terashima, C.; Fujishima, A. Photocatalytic degradation of bisphenol A using titanium dioxide@nanodiamond composites under UV light illumination. J. Colloid Interface Sci. 2021, 582, 1058–1066. [Google Scholar] [CrossRef] [PubMed]
  23. Ali, I.; Kim, S.-R.; Kim, S.-P.; Kim, J.-O. Anodization of bismuth doped TiO2 nanotubes composite for photocatalytic degradation of phenol in visible light. Catal. Today 2017, 282, 31–37. [Google Scholar] [CrossRef]
  24. Kader, S.; Al-Mamun, M.R.; Suhan, M.B.K.; Shuchi, S.B.; Islam, M.S. Enhanced photodegradation of methyl orange dye under UV irradiation using MoO3 and Ag doped TiO2 photocatalysts. Environ. Technol. Innov. 2022, 27, 102476. [Google Scholar] [CrossRef]
  25. Bezerra, P.; Cavalcante, R.; Garcia, A.; Wender, H.; Martines, M.; Casagrande, G.; Giménez, J.; Marco, P.; Oliveira, S.; Machulek, A., Jr. Synthesis, Characterization, and Photocatalytic Activity of Pure and N-, B-, or Ag- Doped TiO2. J. Braz. Chem. Soc. 2017, 28, 1788–1802. [Google Scholar] [CrossRef]
  26. Zeng, L.; Lu, Z.; Li, M.; Yang, J.; Song, W.; Zeng, D.; Xie, C. A modular calcination method to prepare modified N-doped TiO2 nanoparticle with high photocatalytic activity. Appl. Catal. B Environ. 2016, 183, 308–316. [Google Scholar] [CrossRef]
  27. Li, T.; Abdelhaleem, A.; Chu, W.; Pu, S.; Qi, F.; Zou, J. S-doped TiO2 photocatalyst for visible LED mediated oxone activation: Kinetics and mechanism study for the photocatalytic degradation of pyrimethanil fungicide. Chem. Eng. J. 2021, 411, 128450. [Google Scholar] [CrossRef]
  28. Lee, B.T.; Han, J.K.; Gain, A.K.; Lee, K.H.; Saito, F. TEM microstructure characterization of nano TiO2 coated on nano ZrO2 powders and their photocatalytic activity. Mater. Lett. 2006, 60, 2101–2104. [Google Scholar] [CrossRef]
  29. Wafi, M.A.E.; Ahmed, M.A.; Abdel-Samad, H.S.; Medien, H.A.A. Exceptional removal of methylene blue and p-aminophenol dye over novel TiO2/RGO nanocomposites by tandem adsorption-photocatalytic processes. Mater. Sci. Energy Technol. 2022, 5, 217–231. [Google Scholar] [CrossRef]
  30. Toe, E.D.; Kurniawan, W.; Mariquit, E.G.; Hinode, H. Synthesis of N-doped mesoporous TiO2 by facile one-step solvothermal process for visible light photocatalytic degradation of organic pollutant. J. Environ. Chem. Eng. 2018, 6, 5125–5134. [Google Scholar] [CrossRef]
  31. da Silva, A.L.; Trindade, F.J.; Dalmasso, J.L.; Ramos, B.; Teixeira, A.C.S.C.; Gouvêa, D. Synthesis of TiO2 microspheres by ultrasonic spray pyrolysis and photocatalytic activity evaluation. Ceram. Int. 2022, 48, 9739–9745. [Google Scholar] [CrossRef]
  32. Jadhav, P.S.; Jadhav, T.; Bhosale, M.; Jadhav, C.H.; Pawar, V.C. Structural and optical properties of N-doped TiO2 nanomaterials. Mater. Today Proc. 2021, 43, 2763–2767. [Google Scholar] [CrossRef]
  33. Huang, L.; Fu, W.; Fu, X.; Zong, B.; Liu, H.; Bala, H.; Wang, X.; Sun, G.; Cao, J.; Zhang, Z. Facile and large-scale preparation of N doped TiO2 photocatalyst with high visible light photocatalytic activity. Mater. Lett. 2017, 209, 585–588. [Google Scholar] [CrossRef]
  34. Yalçin, Y.; Kiliç, M.; Çinar, Z. Fe+3-doped TiO2: A combined experimental and computational approach to the evaluation of visible light activity. Appl. Catal. B Environ. 2010, 99, 469–477. [Google Scholar] [CrossRef]
  35. da Silva, T.F.; Cavalcante, R.P.; Guelfi, D.R.V.; de Oliveira, S.C.; Casagrande, G.A.; Caires, A.R.L.; de Oliveira, F.F.; Gubiani, J.R.; Cardoso, J.C.; Machulek, A. Photo-anodes based on B-doped TiO2 for photoelectrocatalytic degradation of propyphenazone: Identification of intermediates, and acute toxicity evaluation. J. Environ. Chem. Eng. 2022, 10, 107212. [Google Scholar] [CrossRef]
  36. Ramos, D.D.; Bezerra, P.C.S.; Quina, F.H.; Dantas, R.F.; Casagrande, G.A.; Oliveira, S.C.; Oliveira, M.R.S.; Oliveira, L.C.S.; Ferreira, V.S.; Oliveira, S.L.; et al. Synthesis and characterization of TiO2 and TiO2/Ag for use in photodegradation of methylviologen, with kinetic study by laser flash photolysis. Environ. Sci. Pollut. Res. 2015, 22, 774–783. [Google Scholar] [CrossRef]
  37. Li, D.; Meng, Y.; Li, J.; Song, Y.; Xu, F. TiO2/carbonaceous nanocomposite from titanium-alginate coordination compound. Carbohydr. Polym. 2022, 288, 119400. [Google Scholar] [CrossRef]
  38. Mancuso, A.; Sacco, O.; Sannino, D.; Pragliola, S.; Vaiano, V. Enhanced visible-light-driven photodegradation of Acid Orange 7 azo dye in aqueous solution using Fe-N co-doped TiO2. Arab. J. Chem. 2020, 13, 8347–8360. [Google Scholar] [CrossRef]
  39. Adán, C.; Bahamonde, A.; Fernández-García, M.; Martínez-Arias, A. Structure and activity of nanosized iron-doped anatase TiO2 catalysts for phenol photocatalytic degradation. Appl. Catal. B Environ. 2007, 72, 11–17. [Google Scholar] [CrossRef]
  40. Khatibnezhad, H.; Ambriz-Vargas, F.; Ben Ettouil, F.; Moreau, C. Role of phase content on the photocatalytic performance of TiO2 coatings deposited by suspension plasma spray. J. Eur. Ceram. Soc. 2022, 42, 2905–2920. [Google Scholar] [CrossRef]
  41. Kanjana, N.; Maiaugree, W.; Poolcharuansin, P.; Laokul, P. Synthesis and characterization of Fe-doped TiO2 hollow spheres for dye-sensitized solar cell applications. Mater. Sci. Eng. B 2021, 271, 115311. [Google Scholar] [CrossRef]
  42. Farhangi, N.; Ayissi, S.; Charpentier, P.A. Fe doped TiO2-graphene nanostructures: Synthesis, DFT modeling and photocatalysis. Nanotechnology 2014, 25, 305601. [Google Scholar] [CrossRef] [PubMed]
  43. Devi, L.G.; Kavitha, R. A review on non metal ion doped titania for the photocatalytic degradation of organic pollutants under UV/solar light: Role of photogenerated charge carrier dynamics in enhancing the activity. Appl. Catal. B Environ. 2013, 140–141, 559–587. [Google Scholar] [CrossRef]
  44. He, H.; Sun, D.; Zhang, Q.; Fu, F.; Tang, Y.; Guo, J.; Shao, M.; Wang, H. Iron-Doped Cauliflower-Like Rutile TiO2 with Superior Sodium Storage Properties. ACS Appl. Mater. Interfaces 2017, 9, 6093–6103. [Google Scholar] [CrossRef] [PubMed]
  45. Zheng, X.; Li, Y.; You, W.; Lei, G.; Cao, Y.; Zhang, Y.; Jiang, L. Construction of Fe-doped TiO2−x ultrathin nanosheets with rich oxygen vacancies for highly efficient oxidation of H2S. Chem. Eng. J. 2022, 430, 132917. [Google Scholar] [CrossRef]
  46. Chen, J.; Ding, Z.; Wang, C.; Hou, H.; Zhang, Y.; Wang, C.; Zou, G.; Ji, X. Black Anatase Titania with Ultrafast Sodium-Storage Performances Stimulated by Oxygen Vacancies. ACS Appl. Mater. Interfaces 2016, 8, 9142–9151. [Google Scholar] [CrossRef]
  47. Rauf, M.A.; Meetani, M.A.; Hisaindee, S. An overview on the photocatalytic degradation of azo dyes in the presence of TiO2 doped with selective transition metals. Desalination 2011, 276, 13–27. [Google Scholar] [CrossRef]
  48. Liu, G.; Hua, G.Y.; Wang, X.; Cheng, L.; Pan, J.; Gao, Q.L.; Cheng, H.M. Visible light responsive nitrogen doped anatase TiO2 sheets with dominant {001} facets derived from TiN. J. Am. Chem. Soc. 2009, 131, 12868–12869. [Google Scholar] [CrossRef]
  49. Zhao, Y.; Zhu, L.; Yu, Y.; Gao, F.; Wang, W.; Chen, D.; Zhao, X. Facile one-pot preparation of Ti3+, N co-doping TiO2 nanotube arrays and enhanced photodegradation activities by tuning tube lengths and diameters. Catal. Today 2020, 355, 563–572. [Google Scholar] [CrossRef]
  50. Barkhade, T.; Banerjee, I. Optical Properties of Fe doped TiO2 Nanocomposites Synthesized by Sol-Gel Technique. Mater. Today Proc. 2019, 18, 1204–1209. [Google Scholar] [CrossRef]
  51. Calza, P.; Minella, M.; Demarchis, L.; Sordello, F.; Minero, C. Photocatalytic rate dependence on light absorption properties of different TiO2 specimens. Catal. Today 2020, 340, 12–18. [Google Scholar] [CrossRef] [Green Version]
  52. Moradi, V.; Ahmed, F.; Jun, M.B.G.; Blackburn, A.; Herring, R.A. Acid-treated Fe-doped TiO2 as a high performance photocatalyst used for degradation of phenol under visible light irradiation. J. Environ. Sci. 2019, 83, 183–194. [Google Scholar] [CrossRef] [PubMed]
  53. Ramírez-Sánchez, I.M.; Bandala, E.R. Photocatalytic Degradation of Estriol Using Iron-Doped TiO2 under High and Low UV Irradiation. Catalysts 2018, 8, 625. [Google Scholar] [CrossRef] [Green Version]
  54. Yuan, R.; Wang, S.; Liu, D.; Shao, X.; Zhou, B. Effect of the wavelength on the pathways of 2-MIB and geosmin photocatalytic oxidation in the presence of Fe-N co-doped TiO2. Chem. Eng. J. 2018, 353, 319–328. [Google Scholar] [CrossRef]
  55. Sukhadeve, G.K.; Janbandhu, S.Y.; Kumar, R.; Lataye, D.H.; Ramteke, D.D.; Gedam, R.S. Visible light assisted photocatalytic degradation of Indigo Carmine dye and NO2 removal by Fe doped TiO2 nanoparticles. Ceram. Int. 2022. [Google Scholar] [CrossRef]
  56. Triantis, T.M.; Fotiou, T.; Kaloudis, T.; Kontos, A.G.; Falaras, P.; Dionysiou, D.D.; Pelaez, M.; Hiskia, A. Photocatalytic degradation and mineralization of microcystin-LR under UV-A, solar and visible light using nanostructured nitrogen doped TiO2. J. Hazard. Mater. 2012, 211–212, 196–202. [Google Scholar] [CrossRef]
  57. Ma, J.; Wang, K.; Wang, C.; Chen, X.; Zhu, W.; Zhu, G.; Yao, W.; Zhu, Y. Photocatalysis-self-Fenton system with high-fluent degradation and high mineralization ability. Appl. Catal. B Environ. 2020, 276, 119150. [Google Scholar] [CrossRef]
  58. Liu, Y.; Zhu, Y.; Xu, J.; Bai, X.; Zong, R.; Zhu, Y. Degradation and mineralization mechanism of phenol by BiPO4 photocatalysis assisted with H2O2. Appl. Catal. B Environ. 2013, 142–143, 561–567. [Google Scholar] [CrossRef]
  59. Guo, D.; Wang, Y.; Chen, C.; He, J.; Zhu, M.; Chen, J.; Zhang, C. A multi-structural carbon nitride co-modified by Co, S to dramatically enhance mineralization of Bisphenol f in the photocatalysis-PMS oxidation coupling system. Chem. Eng. J. 2021, 422, 130035. [Google Scholar] [CrossRef]
  60. Zeng, Z.; Ye, F.; Deng, S.; Fang, D.; Wang, X.; Bai, Y.; Xiao, H. Accelerated organic pollutants mineralization in interlayer confined single Pt atom photocatalyst for hydrogen recovery. Chem. Eng. J. 2022, 444, 136561. [Google Scholar] [CrossRef]
Figure 1. Representative scheme showing the steps of the catalyst synthesis process.
Figure 1. Representative scheme showing the steps of the catalyst synthesis process.
Nanomaterials 12 02538 g001
Figure 2. (A) FTIR spectra, and (B) X-ray diffraction patterns of N-, Fe-, and N-Fe doped TiO2 photocatalysts.
Figure 2. (A) FTIR spectra, and (B) X-ray diffraction patterns of N-, Fe-, and N-Fe doped TiO2 photocatalysts.
Nanomaterials 12 02538 g002
Figure 3. SEM image of the surfaces of N-, Fe-, and N-Fe doped TiO2 photocatalysts. (A) TiO2 pure, (B) 0.0125% Fe, (C) 1% Fe, (D) 3% N, (E) 9% N, (F) 2% N + 0.025% Fe, (G) 3% N + 0.025% Fe, and (H) 4% N + 0.025% Fe.
Figure 3. SEM image of the surfaces of N-, Fe-, and N-Fe doped TiO2 photocatalysts. (A) TiO2 pure, (B) 0.0125% Fe, (C) 1% Fe, (D) 3% N, (E) 9% N, (F) 2% N + 0.025% Fe, (G) 3% N + 0.025% Fe, and (H) 4% N + 0.025% Fe.
Nanomaterials 12 02538 g003
Figure 4. TEM images and their size distributions of particles (A) TiO2, (B) 0.0125% Fe, and (C) 3% N.
Figure 4. TEM images and their size distributions of particles (A) TiO2, (B) 0.0125% Fe, and (C) 3% N.
Nanomaterials 12 02538 g004
Figure 5. Nitrogen adsorption-desorption isotherms of (A) TiO2, (B) 0.0125% Fe, and (C) 3%N.
Figure 5. Nitrogen adsorption-desorption isotherms of (A) TiO2, (B) 0.0125% Fe, and (C) 3%N.
Nanomaterials 12 02538 g005
Figure 6. XPS spectra for the TiO2, 0.0125% Fe, and 3%N samples.
Figure 6. XPS spectra for the TiO2, 0.0125% Fe, and 3%N samples.
Nanomaterials 12 02538 g006
Figure 7. High-resolution XPS Ti 2p and O 1s spectra of TiO2 (A,D), 0.0125% Fe (B,E) and 3% N (C,F).
Figure 7. High-resolution XPS Ti 2p and O 1s spectra of TiO2 (A,D), 0.0125% Fe (B,E) and 3% N (C,F).
Nanomaterials 12 02538 g007
Figure 8. (A) UV-Vis absorbance spectra for the synthesized photocatalysts. (B) Tauc plots of photocatalysts with different % N and Fe.
Figure 8. (A) UV-Vis absorbance spectra for the synthesized photocatalysts. (B) Tauc plots of photocatalysts with different % N and Fe.
Nanomaterials 12 02538 g008
Figure 9. (A,C) Normalized 2,4-DMA mineralization curves obtained from the PC treatment of 350 mL of a solution of 0.1 mmol L−1 of 2,4-DMA with 175 mg of photocatalyst, irradiated by a UVA lamp of 18 W (B,D) pseudo-first-order kinetics for PC treatment.
Figure 9. (A,C) Normalized 2,4-DMA mineralization curves obtained from the PC treatment of 350 mL of a solution of 0.1 mmol L−1 of 2,4-DMA with 175 mg of photocatalyst, irradiated by a UVA lamp of 18 W (B,D) pseudo-first-order kinetics for PC treatment.
Nanomaterials 12 02538 g009
Table 1. BET surface areas (SBET), pore volume (Vp), and pore diameters (Dp) for the N-, Fe-, and N-Fe doped TiO2 photocatalysts.
Table 1. BET surface areas (SBET), pore volume (Vp), and pore diameters (Dp) for the N-, Fe-, and N-Fe doped TiO2 photocatalysts.
PhotocatalystsSBET (m2 g−1)DP (nm)VP (cm3 g−1)
TiO278.67 ± 1.509.730.19
TiO2 + 0.0125%Fe84.73 ± 1.479.680.20
TiO2 + 3%N95.70 ± 1.389.040.22
Table 2. Binding energies/eV and contribution of the high-resolution spectra of Ti 2p present in the photocatalysts. Values are obtained from high-resolution XPS spectra in Figure 6.
Table 2. Binding energies/eV and contribution of the high-resolution spectra of Ti 2p present in the photocatalysts. Values are obtained from high-resolution XPS spectra in Figure 6.
PhotocatalystsComposition TiBinding Energies/eV
Ti 2p1/2
Binding Energies/eV
Ti 2p3/2
Contribution/
%
TiO2Ti4+464.30458.5896.84
Ti3+462.80457.0803.16
0.0125% FeTi4+464.30458.5894.20
Ti3+462.80457.0805.80
3% NTi4+464.31458.5997.64
Ti3+462.81457.0907.38
Table 3. Band gap energy values for all photocatalysts.
Table 3. Band gap energy values for all photocatalysts.
PhotocatalystsEbg/eV
TiO23.16
TiO2 + 0.0125%Fe3.08
TiO2 + 1%Fe2.82
TiO2 + 3%N3.14
TiO2 + 9%N3.07
3% N + 0.025% Fe3.05
4% N + 0.025% Fe3.06
Table 4. Removal percentage mineralization and pseudo-first-order constant.
Table 4. Removal percentage mineralization and pseudo-first-order constant.
Photocatalysts% Dopant (w/w)Mineralization Rate (%)Pseudo-First-Order Constant
in 240 minin 360 mink2,4-DMAR2
Photolysis-040.00010.961
TiO2065780.0040.983
TiO2 + N157720.0030.993
264830.0040.988
3811000.0070.989
4641000.0060.989
555930.0060.983
653690.0030.991
762700.0040.990
928380.0020.999
TiO2 + Fe0.01251001000.0100.992
0.0251001000.0100.980
0.05801000.0100.994
0.155800.0040.990
0.2563720.0040.983
162700.0040.990
TiO2 + %N + 0.025 Fe331390.0020.997
448520.0020.992
Table 5. Comparison of N-TiO2 and Fe-TiO2 photocatalysts with others.
Table 5. Comparison of N-TiO2 and Fe-TiO2 photocatalysts with others.
PhotocatalystExperimental ConditionsModel MoleculeMineralizationRef
Mass/mgVolume/mLLight SourceAnalysis Time/h
0.0125% Fe TiO2175350UV-A412 ppm 2,4-DMA100%Present study
2 mol% Fe-TiO250100100 W fluorescent bulb110 ppm Indigo Carmine68%[55]
N-TiO215UV-A/solar/UV-vis8Microcystin-LR80%[56]
gC3N43030Visible-light irradiation320 ppm 2,4-dichlorophenol15.2%[57]
BiPO4 + H2O25010011 W low pressure lamp (λ = 254 nm)2phenol40%[58]
pg-C3N4/Co3O4/CoS550500 W xenon lamp230 ppm Bisphenol F51.9%[59]
PtSA/g-C3N425100Xenon light410 ppm p-chlorophenol70%[60]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Faustino, E.; da Silva, T.F.; Cunha, R.F.; Guelfi, D.R.V.; Cavalheri, P.S.; de Oliveira, S.C.; Caires, A.R.L.; Casagrande, G.A.; Cavalcante, R.P.; Junior, A.M. Synthesis and Characterization of N and Fe-Doped TiO2 Nanoparticles for 2,4-Dimethylaniline Mineralization. Nanomaterials 2022, 12, 2538. https://doi.org/10.3390/nano12152538

AMA Style

Faustino E, da Silva TF, Cunha RF, Guelfi DRV, Cavalheri PS, de Oliveira SC, Caires ARL, Casagrande GA, Cavalcante RP, Junior AM. Synthesis and Characterization of N and Fe-Doped TiO2 Nanoparticles for 2,4-Dimethylaniline Mineralization. Nanomaterials. 2022; 12(15):2538. https://doi.org/10.3390/nano12152538

Chicago/Turabian Style

Faustino, Emerson, Thalita Ferreira da Silva, Rebeca Fabbro Cunha, Diego Roberto Vieira Guelfi, Priscila Sabioni Cavalheri, Silvio César de Oliveira, Anderson Rodrigues Lima Caires, Gleison Antonio Casagrande, Rodrigo Pereira Cavalcante, and Amilcar Machulek Junior. 2022. "Synthesis and Characterization of N and Fe-Doped TiO2 Nanoparticles for 2,4-Dimethylaniline Mineralization" Nanomaterials 12, no. 15: 2538. https://doi.org/10.3390/nano12152538

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