Next Article in Journal
New Trends on Vanadium Chemistry, Biochemistry, and Medicinal Chemistry
Next Article in Special Issue
Synergistic Correlation in the Colloidal Properties of TiO2 Nanoparticles and Its Impact on the Photocatalytic Activity
Previous Article in Journal
A Tetranuclear Dysprosium Schiff Base Complex Showing Slow Relaxation of Magnetization
Previous Article in Special Issue
Effect of Temperature on the Adhesion and Bactericidal Activities of Ag+-Doped BiVO4 Ceramic Tiles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

TiO2-La2O3 as Photocatalysts in the Degradation of Naproxen

by
Adriana Marizcal-Barba
1,
Isaias Limón-Rocha
1,
Arturo Barrera
2,
José Eduardo Casillas
2,
O. A. González-Vargas
3,
José Luis Rico
4,
Claudia Martinez-Gómez
5 and
Alejandro Pérez-Larios
1,*
1
Materials, Water and Energy Research Laboratory, Department of Engineering, Los Altos University Center, University of Guadalajara, Tepatitlán de Morelos 47600, Mexico
2
Laboratory of Catalytic Nanomaterials, Department of Basic Sciences, La Ciénega University Center, University of Guadalajara, Ocotlán 47820, Mexico
3
Control and Automation Engineering Department, School of Mechanical and Electrical Engineering-Zacatenco, National Polytechnic Institute, Ciudad de México 07738, Mexico
4
Catalysis Laboratory, Faculty of Chemical Engineering, Michoacan University of San Nicolás de Hidalgo, Morelia 58060, Mexico
5
Department of Chemistry, Division of Natural and Exact Sciences, Campus Guanajuato, University of Guanajuato. Noria Alta S/N, Col. Noria Alta, Guanajuato 36050, Mexico
*
Author to whom correspondence should be addressed.
Submission received: 9 April 2022 / Revised: 18 May 2022 / Accepted: 19 May 2022 / Published: 22 May 2022
(This article belongs to the Special Issue Nanocomposites for Photocatalysis)

Abstract

:
The indiscriminate use of naproxen as an anti-inflammatory has been the leading cause of pollution in sewage effluents. Conversely, titanium dioxide is one of the most promising photocatalyst for the degradation of pollutants. Ti-La mixed oxides containing 0, 1, 3, 5, and 10 wt.% of lanthanum were synthetized by sol-gel and tested as photocatalysts in the degradation of naproxen (NPX). The materials were further characterized by X-ray diffraction (XRD), nitrogen physisorption (BET), scanning electron microscopy (SEM), UV-Vis and Fourier-transform infrared spectroscopy (FT-IR), transmission electron microscopy (TEM), and X-ray photoelectron spectroscopy (XPS). The XRD patterns resembled that of anatase titania. The Eg values, determined from the UV-Vis spectra, vary from 2.07 to 3.2 eV corresponded to pure titania. The photocatalytic activity of these materials showed a degradation of naproxen from 93.6 to 99.8 wt.% after 4 h under UV irradiation.

1. Introduction

Drugs top the list of emerging contaminants detected in wastewater effluents [1]. The constant increase and chemical stability of these drugs hinders their removal and/or disposal in conventional wastewater treatment plants [2,3], which causes a deterioration of aquatic life [4,5]. Different water treatments have been reported using advanced oxidation processes (AOPs) [6,7,8] for the elimination of these drugs, such as anti-inflammatories [4,9], antipyretics, and analgesics [10,11]. Among the promising AOPs, photocatalysis stands out, since it manages to degrade and mineralize up to 100% of the contaminants [12,13]. Photocatalytic processes are based on the use of nanomaterials that have the ability to transfer charge between the semiconductor and the aqueous solution, of transference, where the redistribution caused by the absorption of photons (hv) generates the hollow electron pair (e-h). This, when interacting in the medium, generates hydroxyl radicals (OH) and/or superoxide anions that can function as reducing or oxidizing agents to degrade organic molecules [14,15]. Among the most used photocatalysts are oxides such as TiO2 [16], ZnO [17,18], Nb2O5 [19], CdS [20], and SnO2 [21], among others [22]; some researchers have even used mixed oxides as nanocomposites to have greater efficiency in the degradation of this type of contaminant, including TiO-ZnO [2,23], TiO2-Fe2O3 [24], TiO-Al2O3 [25], Ti-Si-Fe [13], and Ti-Zr [26].
Many studies have focused on using TiO2 either as a mixed oxide or as support because its use prevents the recombination of the photo- generated hollow electron pair (e-h) and, thus, reduces the energy band [27,28,29]. Among the nanomaterials used for band gap reduction, rare earths can be found [30,31], such as Ce [32] and La [33]. The use of TiO2-CeO in the degradation of dyes, lignin [34], and metronidazole [35] has been reported. TiO2-La2O3 has been used to degrade volatile organic compounds [28], dyes [36,37,38], and drugs [39], and various catalysts have been reported to degrade mostly dyes (Table 1).
The effect of the synthesis parameters on the photocatalytic activity of Ti-La materials in the degradation of naproxen (NPX) is herein reported. Lanthanum is used as a dopant of titania for this purpose. The effect of La concentration on the photocatalytic activity is also explored. The samples are labelled as Ti-LaX where X represents the weight percentage of lanthanum.

2. Results and Discussion

2.1. Characterization

2.1.1. Scanning Electron Microscopy

SEM micrographs of the Ti-La oxides after calcination are presented in Figure 1. Independent of the La content, the morphology of all samples is similar [28]. Agglomerates comprised of particles of about 200 nm are noticed in all samples. The elemental composition determined by Energy-dispersive X-ray spectroscopy (EDS) is also included in Figure 1. The calculated atomic ratios are consistent with the XPS results, in agreement with other research works [38,40].

2.1.2. Nitrogen Physisorption

Figure 2 shows the N2 adsorption–desorption isotherms and the pore size distribution of the Ti-La samples. The isotherms are of type IV and present type H2 hysteresis associated with capillary condensation, Figure 2a. These isotherms are characteristic of mesopore materials [29]. Similar observations were previously noticed by Perez-Larios et al. for TiO2-ZnO, Mn/TiO2, Ti-Zr, and Ti-Co catalysts [20,24,30,31]. Figure 2b show unimodal pore distribution with a mean pore size which decreases as the lanthanum content augments. Table 2 presents the specific surface areas determined by the BET equation and the mean pore size of Ti-La samples. The presence of lanthanum into TiO2 increases the specific surface areas proportionally, whereas the mean pore size increases for Ti-La1 compared to that for TiO2 and then decreases as the lanthanum content augments.

2.1.3. X-ray Diffraction

The diffraction patterns of the calcined samples before the photocatalytic test are shown in Figure 3. All samples present diffractions which resemble those assigned for anatasa titania. The characteristic diffractions of anatase titania are located at 25, 37, 48, 54, 55, 62, 71, and 75 2θ degrees, which correspond to (101), (004), (200), (105), (211), (204), (116), and (311) crystallographic planes, respectively, according to JCPDS charts No. 00-001-0562 [9]. Additional peaks assigned to La2O3 were unnoticed, indicating, as a consequence, a well lanthanum distribution in the titanium oxide structure and/or the presence of small lanthanum oxides species out of the detection limit of the instrument. The most intensive diffraction at about 25.4 degrees 2θ decreases and became broader as the lanthanum content augments [45,46]. The unit cell parameters presented in Table 2 increase as the lanthanum content augments compared to that of pure titania. Furthermore, the crystallite size, D, decreases as function of lanthanum. Our results, therefore, suggest that La3+ cations are well dispersed in the titanium oxide structure, probably located in the interstitial positions since the ionic radii of La3+ is greater to that of Ti4+, 0.106 and 0.062 nm, respectively [47], and substitution of titanium by lanthanum ions in the TiO2 lattice is nonviable. However, assuming the presence of very small lanthanum oxides species (out of the detection limit of the XRD equipment) the replacement of lanthanum by titanium in the structure of lanthanum oxides is possible and Ti3+ can be formed [48,49]. The average crystallite sizes were determined using the Scherrer equation taken the half width (b) of the most intensive peak located at 25.4 degrees 2θ and are reported in Table 2. The average crystallite size of the Ti-La samples is significant reduced compared to that of pure titania (20.58 nm). Table 2 also shows that increasing the lanthanum content a further reduction in the average crystallite size is observed, as previously noticed for similar Ti-La samples [43,50,51].

2.1.4. UV-Vis Spectroscopy

Figure 4 shows the band gap energies of the Ti-La samples determined from UV-Vis spectra. The insert in this figure indicates that all samples exhibit an optical absorption at about 400 nm, which can be attributed to the electron transition of the Ti-O [52]. A slightly shift to the red region (2.92 to 3.01 eV) for the Ti-La samples compared to pure TiO2 resulted after the incorporation of La3+ cations into TiO2 lattice [41]. These observations are in agreement with other research works related to doped titania. A shift in the absorption edge to shorter wavenumber is clearly observed [53].

2.1.5. FT-IR Spectroscopy

Figure 5 shows the FTIR spectra of TiO2 and Ti-La samples from 4000 to 400 cm−1. The bands at about 448 cm−1 are associated to Ti-O vibrations [54] and related to the stretching of the Ti–OH and Ti–O bonds [38]. Although the samples were previously calcined before the FT-IR measurement, the bands at about 1600 and 1500 cm−1 correspond to C=O and C=C bonds originated from the residual organic compounds remained after synthesis [55].

2.1.6. X-ray Photoelectron Spectrometry

The binding energies for Ti 2p are located at about 459 and 465 eV and correspond to Ti 2p3/2 and Ti 2p1/2, respectively. The separation of these two peaks is about 5.5 eV, which is in good agreement with previous publications related to doped titania (Figure 6) [49,56,57].
The La 3d5/2 spectrum presents two peaks, one at about 834 eV and a satellite peak at about 839 eV, with a splitting of about 5 eV (Figure 7). Although the interpretation of the XPS spectrum of pure lanthanum oxide is complicated, since La is very sensitive to water and CO2 from air and form a hydroxide and/or carbonate, it is likely that lanthanum as a dopant in our Ti-La samples is chemically more stable, since it is mainly buried into the titanium oxide structure. For a pure La2O3 sample exposed to air at ambient temperature, on which the formation of a hydroxide and/or carbonate have already occurred, the splitting of the two peaks of the La 3d5/2 spectrum is about 3.6 eV [57,58,59,60], whereas in ours, the separation is about 5 eV. The spectrum of the La 3d3/2 also shows two peaks, one about 851 eV and a satellite at about 856 eV. Their positions randomly change as function of the lanthanum content, similarly as those for La 3d5/2.
The XPS spectra for O 1s of the Ti-La samples are shown in Figure 8. Although oxygen is bound to titanium and lanthanum, its contribution is overlapped, since only one broad peak is present in all spectra. However, accordingly to the difference in electronegativity of both elements, the signal of the oxygen attached to titanium would appear at lower binding energy, compared to that of the oxygen bound to lanthanum [61]. The O 1s spectrum could be deconvoluted in two signals, an intense peak at about 530.5 eV and the other at about 532 eV, for the oxygen bound to titanium and lanthanum, respectively. The small peak at 532 eV would be in agreement with the low lanthanum content in the Ti-La samples [62].
The presence of C in the XPS spectra, most likely resulting from the organic compounds used during synthesis, are presented in Figure 9. The C 1s spectra show a single broad peak at about 285 eV. However, their positions randomly change accordingly to the lanthanum content. It was reported that for a pure La2O3 sample exposed to air, the C 1s spectrum shows two signals, one at 284.5 and the other at 289.5 eV, assigned to adventitious carbon C 1s and to carbonate C 1s, respectively [63,64]. Since our C 1s spectra present a single broad peak centered at about 285 eV, the formation of lanthanum carbonate after exposing our Ti-La samples to air was excluded. Carbon species which remain after calcination, such as C-O-C and C=O, could well be present in our Ti-La samples.
Additional information obtained from XPS is presented in Table 3. The determined weight percentage is in agreement with the theoretical values. The shifts to lower binding energies for Ti-La5 in the Ti 2p and La 3d spectra, compared to other Ti-La samples, are unclear and more studies are required to elucidate these changes [57,60].
TEM images (Figure 10) show an interplanar spacing of 0.35 nm, corresponding to the (1 0 1) plane lattice space, typical of TiO2 [40]. It can also be seen in nano-globular form with other nanostructures present for the Ti-La samples containing 1 3, 5, and 10 wt.% of Lanthanum, respectively. The differences in contrast between the doped and pure titania confirm the incorporation of La3+ into the titania structure. The results are also consistent with those obtained from the XRD analysis and SEM.
Figure 10 presents some TEM images of the Ti and Ti-La samples before the photocatalytic test. The interplanar spacing of 0.35 nm, corresponding to the (1 0 1) plane of TiO2 [40], changes as function of the lanthanum content which is in agreement with our XRD results. Further analysis of the TEM images to determine the nanoparticles size distribution was performed by the Imajej software measuring at least 50 nanoparticles [41,42,43].

2.2. Phococatalytic Activity

Although the photocatalytic experiments were analyzed every 30 min, Figure 11 only shows the degradation of NPX from 120 min onwards. This was done to increase the visibility of the results at the end of the experiments. Among the samples, the best catalysts after 4 h under UV irradiation were Ti-La1, Degussa P25, and Ti-La3 which show more than 90% degradation of NPX. Conversely, in absence of a catalyst, the decomposition of NPX by photolysis using UV was about 30% after 4 h of reaction, results not shown.
Figure 11 shows the NPX photodegradation rate curves of the TiO2, Ti-La1, Ti-La3, Ti-La5, and Ti-La10 nanophotocatalysts. The adsorption effect can be observed within the first 30 min, where the Ti-La nanocomposite adsorbs more contaminant than that of the bare TiO2; this could be attributed to particle size and pore diameter, and these results are in agreement with previous reports where the Ti-Zr nanocomposite in the same relation at 1% degraded more than the others photocatalysts [56]. Otherwise, it is observed that the TiO2, Ti-La1, and Ti-La3 samples turned out to be slightly more efficient than the Ti-La5 and Ti-La10 photocatalysts at 150 min. It is interesting to note that the Ti-La1 and Ti-La3 doping photocatalysts could lead to the formation of titania lattice distortion, which is prone to form the oxygen vacancy, and the vacancy could act as a trapper to inhibit the recombination of photo-generated electron-hole pairs [43]. It is a possibility that an impurity level has formed under the coactions of Ti3+ and oxygen vacancies by La doping, which could inhibit the recombination of charge carriers. Then, La2O3 on the TiO2 surface could also transfer the electrons to the surface, which is advantageous for photocatalytic activity [42].
It can be seen from Figure 11, that pure TiO2, (1 × 10−7 min−1) showed a better photocatalytic activity than photolysis (9.7 × 10−6 min−1) under UV light, indicating that nano-TiO2 has an effect on accelerating the decomposition rate of NPX under UV light illumination. The nanocomposites Ti-La1 under the same experimental conditions was higher, which reveals that samples are readily excited under light irradiation, and photogenerated holes accelerate the oxidation of NPX before recombination with electrons. This result is consistent with other research [2,43,65,66]. Obviously, on this occasion, the optimal reaction time is 4 h, and the photocatalytic activity (1.5 × 10−6 min−1) was better than that of TiO2.
The reusability of TiO2 and Ti-La samples was evaluated by repeating the NPX degradation five times using the same catalysts. After a run, the catalyst was separated by filtration and dried in an oven overnight. Each experiment was performed with a fresh NPX solution, using the spent catalyst, during 4 h and under UV irradiation. As it can be seen in Figure 12, the photocatalytic activity of Ti-La1 is still attractive, since it declined about 5% after the fifth run.
The total amount of carbon (TOC) found in the solutions after the photocatalytic tests were also determined and the percentages of mineralization are presented in Table 4. As can be seen, TiO2 and Ti-La1 are not only effective in degrading but also to mineralize NPX. According to the literature, 2-acetyl-6-methoxynaphthalene, detected by HPLC, is the main intermediate during the decomposition of NPX [66].
The mechanism for the activation of TiO2 or Ti-La photocatalysts, under UV is similar to that previously reported by Chaker et al. for La/TiO2 catalysts [27]. Briefly, the electrons and holes (e and h+) generated during UV irradiation, reach the surface of the catalyst and react with water and oxygen to generate hydroxyl radicals and oxide radical anions through respective oxidative and reductive reactions. These radicals are strong oxidants and were able to decompose and mineralize NPX present in water. The differences in the photocatalytic activity of our samples are due to the physico-chemical properties of the catalysts. The following table summarizes the results presented in the literature concerning the decomposition of NPX on various catalysts. Although the information included in Table 5 is useful, comparison of the catalytic activities is not straightforward since the experimental conditions used in each research work were different.

3. Conclusions

In summary, TiO2 and Ti-La oxides were synthesized by sol-gel and tested as photocatalysts in the degradation of naproxen. Results from characterization indicate that the incorporation of lanthanum into titanium oxide increases the specific surface area and reduces the average crystalline size compared to those of pure TiO2. In addition, lanthanum atoms are well distributed into the titanium oxide structure, probably located in the interstitial voids of the titania lattice. XPS results show that it is unlikely that lanthanum forms carbonate and/or a hydroxide after exposing our catalysts to air, since lanthanum atoms are well buried into the titanium oxide structure. XPS also shows that carbon species remain on the catalysts even after calcination. The photocatalytic activities of our samples in the degradation of NPX vary as function of the catalyst composition. Our results indicate that low concentration of lanthanum in titanium oxide is very attractive, since Ti-La1 showed the highest photocatalytic activity per mass of catalyst after 4 h and under UV irradiation. Furthermore, it was able to mineralize NPX and presents good stability after five cycles. The positive photocatalytic effects of Ti-La1 are probably related to the decrement in the recombination of electron-hole pairs.

4. Materials and Methods

4.1. Materials

All reactants (reagent grade) were acquired from Sigma Aldrich (Chemical Co., St. Louis, MO, USA.), such as Titanium (IV) butoxide (C16H36O4Ti, ≥97%), lanthanum nitrate (La(NO)3.6H2O, ≥99.9%), ethanol (CH3CH2OH, ≥96%), and naproxen (CH3OC10H6CH(CH3)CO2H, ≥98%); deionized water (H2O) was used in all experiments.

4.2. Catalyst Preparation

The Ti–La samples were prepared by sol-gel and the required quantities of titanium (IV) butoxide and lanthanum nitrate to prepare the oxides containing 1.0, 3.0, 5.0 and 10.0 wt.% of La in titania were used. The metal compounds were dissolved in 44 mL of ethanol and 18 mL of deionized water under stirring and the pH of the solution was adjusted to 3 with an aqueous nitric-acid solution. The water/alkoxide molar ratio was 8. The solution was heated up to 70 °C and kept at this temperature for 24 h. The resulting solids were dried at 100 °C during 24 h. After this treatment, the samples were annealed at 500 °C during 4 h, applying a heating rate of 2 °C/min. Pure titania prepared following a similar procedure, and a commercial catalyst, Degussa P25, were used as references. The samples were then ready to be used as photocatalysts in the photocatalytic degradation of NPX.

4.3. Characterization Equipment

The surface morphology of the mixed oxides was analyzed using a scanning electron microscope (Tescan, MIRA 3LMU, Boston, MA, USA)) operated at 20 kV and provided with an Energy Dispersive Spectrometer (Bruker, QUANTAX, MA, USA), which was utilized to determine the superficial elemental composition.
The textural properties of the catalytic samples were determined from nitrogen adsorption-desorption experiments using a micromeritics TriStar II Plus equipment (Norcross, GA, USA). The samples were first degassed under vacuum at 200 °C for 3 h. Nitrogen adsorption isotherms were measured at liquid-nitrogen temperature (77 K), and N2 pressures ranging from 10–6 to 1.0 P/P0. The specific surface areas were calculated following the Brunauer-Emmett-Teller (BET) method and the pore size distribution was obtained according to the Barret-Joyner-Halenda (BJH) method, using the Equation (1):
ln (p/p0) = 2γ Vm/RT(rptc)
where rp is the pore radius and tc the thickness of the adsorbed multilayer film, which is formed prior to pore condensation [61].
The diffraction patterns of the mixed oxides were recorded using an X-ray diffractometer equipped with Cu Kα radiation (λ = 0.154 nm). The diffraction intensities as a function of 2θ were measured between 10 and 90°, using a step of 0.02° and counting time of 0.2 s per step. Furthermore, the average particle crystalline size was determined by the Scherer Equation (2) using the most intense peak for each sample:
D   = K λ β c o s   ( θ ) ,
where K is the shape factor (0.89), λ is the X-ray wavelength (0.154 nm), β is the peak broadening at half maximum, and θ is the Bragg angle. Additionally, interplanar spacing (d) can be calculated from the Bragg law Equation (3):
d   ( A ) = n λ 2 sin θ
The UV-Vis absorption spectra were obtained using a UV-Vis spectrophotometer (Shimadzu UV-2600, Tokio, Japan) coupled with an integration sphere for diffuse reflectance studies. The equipment was calibrated with barium sulfate BaSO4 as a reference. The optical absorption was measured in the wavelength range of 190 to 900 nm. From the plot, the band gap energy for each sample was calculated using the following empirical Equation (4):
E g = h C λ
where Eg is the band gap energy; h is Plank constant equal to 6.626 × 10−34 J/s; C is the velocity of light equal to 2.99 × 108 m/s; and λ is the detection wavelength.
The FTIR spectra of the materials were recorded with an FTIR (Shimadzu, IRTracer-100, Tokyo, Japan) spectrophotometer using attenuated total reflectance (ATR) with a diamond waveguide (XR model). A detector of fast recovery deuterated triglycine sulfate (DTGS) (standard) was used for the analysis. The spectra were recorded at room temperature, with 24 scans and 4 cm−1 of resolution and from 4000 cm−1 to 400 cm−1.
The XPS spectra were collected using a SPEC Phoibos 150 provided with a monochromatic Al Ka X-ray source (1487 eV). The position of the O 1s peak at 531.0 eV was monitored for each sample to ensure that no binding energy shift occurred due to charging.
High resolution images were acquired using a high-resolution transmission electron microscope, HRTEM, (Jeol microscope, JEM-ARM200F, Boston, MA, USA.) operated at 200 kV. The resulting images were analyzed using Gatan Micrograph software v. 3.7.0. (Pleasanton, CA, USA).

4.4. Photocatalytic Activity

The reaction was carried out in a 350 mL Pyrex reactor using an aqueous solution of 30 mg/L NPX. For the photocatalytic experiments, 200 mg of photocatalyst was used in each run. The NPX solution was irradiated with a 1 mW*cm−2 ultraviolet lamp of 254 nm wavelength submerged in a quartz tube at pH 7 ± 0.15. In the first 30 min, the absorption capacity of the material was measured under dark conditions. Irradiation was immediately started and samples were taken every 30 min for 4 h. The analysis of the samples was carried out in a Shimadzu model 2600 UV-Vis spectrophotometer at a wavelength of the maximum absorbance for NPX at 230 nm [2]. To determine the reusability, the spent catalyst was recovered, dried, and calcined, before reusing it.
The NPX concentration was calculated from the corresponding maximum absorption wavelength. The absorbance was proportional to the Beer–Lambert law in the range of the studied NPX concentrations. The degradation percentages of the NPX were calculated by the following equation (Equation (5)):
D e g r a d a t i o n   % = ( C C 0 ) 100
where C0 and C is the initial and the actual concentration of NPX, and corresponds to a change in the current concentration of the NPX relative to the initial value of C0 over time. Furthermore, the photocatalytic experiments were carried out three times to verify its reproducibility.
Furthermore, the total organic carbon in the samples was measured using a TOC-LCSN equipment (Shimadzu, model), and calculated applying the Equation (6).
T O C = T C I C
where TOC is the amount of total organic carbon (mg L−1), TC is the amount of total carbon (mg L−1), and IC inorganic carbon (mg L−1) in aqueous solution.
The quantitative analysis of the reaction kinetics of the degradation of organic compounds can be obtained by fitting the experimental data using the Langmuir–Hinshelwood model and expressed by the following Equation (7):
r = d c d t = k K c 1 + K c
where k is the rate constant, K the equilibrium constant, and c is the concentration of the organic pollutant. Nonetheless, Equation (1) can be simplified due to the low concentration of pollutant Kc < 1 to adapt to the form of a first-order apparent rate Equation (8):
l n   C 0 C = K A p p t
where kApp is the apparent pseudo-first-order rate constant (min−1), C and C0 are the final and initial organic compound concentrations (mg L−1), and t is the reaction time (min).

Author Contributions

Writing—original draft preparation, A.M.-B.; writing—review and editing, C.M.-G., A.B., J.E.C., O.A.G.-V., I.L.-R., J.L.R. and A.P.-L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Marizcal-Barba thanks CONACYT for the scholarship (799894) and appreciates the support of the microscopy Lab for characterization of photocatalysts to Martín Flores and Milton Vázquez, as well as the technicians of the equipment Ing. Sergio Oliva and José Rivera, for the characterization by XRD, SEM-DES, and XPS analysis (project 270660, Support for the Strengthening and Development of the Scientific and Technological Infrastructure).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

TiO2Titanium dioxide
NPXnaproxene
XRDX-ray diffraction
SEMscanning electron microscopy
EDSEnergy Dispersive X-ray
UV-VisVisible and ultra-violet ligth
FT-IRFourier Transform Infrared Spectroscopy
TEMtransmission electron microscopy
XPSX-ray photoelectron spectroscopy
AOP’sAdvanced Oxidation Processes
OHhydroxyl radicals
BETBrunauer-Emmett-Teller
BJHBarrett-Joyner-Halenda
JCPDSJoint Committee on Powder Diffraction Standards
M-OMetal-Oxygen
eVelectron Volt
TOCTotal Organic Carbon
TCTotal Carbon
ICInorganic Carbon
HPLCHigh Performance Liquid Chromatography

References

  1. Uheida, A.; Mohamed, A.; Belaqziz, M.; Nasser, W.S. Photocatalytic degradation of Ibuprofen, Naproxen, and Cetirizine using PAN-MWCNT nanofibers crosslinked TiO2-NH2 nanoparticles under visible light irradiation. Sep. Purif. Technol. 2019, 212, 110–118. [Google Scholar] [CrossRef]
  2. Strbac, D.; Aggelopoulos, C.; Štrbac, G.; Dimitropoulos, M.; Novaković, M.; Ivetić, T.; Yannopoulos, S.N. Photocatalytic degradation of Naproxen and methylene blue: Comparison between ZnO, TiO2 and their mixture. Process Saf. Environ. Prot. 2018, 113, 174–183. [Google Scholar] [CrossRef]
  3. Kaur, M.; Mehta, S.K.; Kansal, S.K. Construction of Multifunctional NH2-UiO-66 Metal Organic Framework: Sensing and Photocatalytic Degradation of Ketorolac Tromethamine and Tetracycline in Aqueous Medium. Environ. Sci. Pollut. Res. 2022, 1, 1–21. [Google Scholar] [CrossRef] [PubMed]
  4. Pylypchuk, I.V.; Daniel, G.; Kessler, V.G.; Seisenbaeva, G.A. Removal of Diclofenac, Paracetamol, and Carbamazepine from Model Aqueous Solutions by Magnetic Sol–Gel Encapsulated Horseradish Peroxidase and Lignin Peroxidase Composites. Nanomaterials 2020, 10, 282. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Wojcieszyńska, D.; Guzik, U. Naproxen in the environment: Its occurrence, toxicity to nontarget organisms and biodegradation. Appl. Microbiol. Biotechnol. 2020, 104, 1849–1857. [Google Scholar] [CrossRef] [Green Version]
  6. Martínez-Huitle, C.A.; Panizza, M. Electrochemical oxidation of organic pollutants for wastewater treatment. Curr. Opin. Electrochem. 2018, 11, 62–71. [Google Scholar] [CrossRef]
  7. Bello, M.M.; Raman, A.A.A.; Asghar, A. A review on approaches for addressing the limitations of Fenton oxidation for recalcitrant wastewater treatment. Process Saf. Environ. Prot. 2019, 126, 119–140. [Google Scholar] [CrossRef]
  8. Glaze, W.H. Reaction products of ozone: A review. Environ. Health Perspect. 1986, 69, 151–157. [Google Scholar] [CrossRef]
  9. Ahmadpour, N.; Sayadi, M.H.; Sobhani, S.; Hajiani, M. A potential natural solar light active photocatalyst using magnetic ZnFe2O4 @ TiO2/Cu nanocomposite as a high performance and recyclable platform for degradation of naproxen from aqueous solution. J. Clean. Prod. 2020, 268, 122023. [Google Scholar] [CrossRef]
  10. Mohamed, A.; Salama, A.; Nasser, W.S.; Uheida, A. Photodegradation of Ibuprofen, Cetirizine, and Naproxen by PAN-MWCNT/TiO2–NH2 nanofiber membrane under UV light irradiation. Environ. Sci. Eur. 2018, 30, 47. [Google Scholar] [CrossRef]
  11. Castro-Pastrana, L.; Cerro-López, M.; Toledo-Wall, M.; Gómez-Oliván, L.; Saldívar-Santiago, M.; Castro-Pastrana, L.; Cerro-López, M.; Toledo-Wall, M.; Gómez-Oliván, L.; Saldívar-Santiago, M. Análisis de fármacos en aguas residuales de tres hospitales de la ciudad de Puebla, México. Ing. Del. Agua 2021, 25, 59. [Google Scholar] [CrossRef]
  12. Reyes, M.H.; Solis, R.C.; Ruiz, F.; Rodriguez-Gonzalez, V.; Moctezuma, E. Promotional effect of metal doping on nanostructured TiO2 during the photocatalytic degradation of 4-chlorophenol and naproxen sodium as pollutants. Mater. Sci. Semicond. Process. 2019, 100, 130–139. [Google Scholar] [CrossRef]
  13. Amini, Z.; Givianrad, M.H.; Saber-Tehrani, M.; Azar, P.A.; Husain, S.W. Synthesis of N-doped TiO2/SiO2/Fe3O4 magnetic nanocomposites as a novel purple LED illumination-driven photocatalyst for photocatalytic and photoelectrocatalytic degradation of naproxen: Optimization and different scavenger agents study. J. Environ. Sci. Health Part A 2019, 54, 1254–1267. [Google Scholar] [CrossRef] [PubMed]
  14. Marizcal-Barba, A.; Sanchez-Burgos, J.A.; Zamora-Gasga, V.; Larios, A.P. Study of the Response Surface in the Photocatalytic Degradation of Acetaminophen Using TiO2. Photochem 2022, 2, 225–236. [Google Scholar] [CrossRef]
  15. Eidsvåg, H.; Bentouba, S.; Vajeeston, P.; Yohi, S.; Velauthapillai, D. TiO2 as a Photocatalyst for Water Splitting—An Experimental and Theoretical Review. Molecules 2021, 26, 1687. [Google Scholar] [CrossRef] [PubMed]
  16. Abel, S.; Jule, L.T.; Belay, F.; Shanmugam, R.; Dwarampudi, L.P.; Nagaprasad, N.; Krishnaraj, R. Application of Titanium Dioxide Nanoparticles Synthesized by Sol-Gel Methods in Wastewater Treatment. J. Nanomater. 2021, 2021, 1–6. [Google Scholar] [CrossRef]
  17. Primo, J.D.O.; Bittencourt, C.; Acosta, S.; Sierra-Castillo, A.; Colomer, J.-F.; Jaerger, S.; Teixeira, V.C.; Anaissi, F.J. Synthesis of Zinc Oxide Nanoparticles by Ecofriendly Routes: Adsorbent for Copper Removal from Wastewater. Front. Chem. 2020, 8, 571790. [Google Scholar] [CrossRef]
  18. Limón-Rocha, I.; Guzmán-González, C.A.; Anaya-Esparza, L.M.; Romero-Toledo, R.; Rico, J.L.; González-Vargas, O.A.; Pérez-Larios, A. Effect of the Precursor on the Synthesis of ZnO and Its Photocatalytic Activity. Inorganics 2022, 10, 16. [Google Scholar] [CrossRef]
  19. Troncoso, F.D.; Tonetto, G.M. Nb2O5 monolith as an efficient and reusable catalyst for textile wastewater treatment. Sustain. Environ. Res. 2021, 31, 35. [Google Scholar] [CrossRef]
  20. Al-Enizi, A.M.; El-Halwany, M.M.; Al-Abdrabalnabi, M.A.; Bakrey, M.; Ubaidullah, M.; Yousef, A. Novel Low Temperature Route to Produce CdS/ZnO Composite Nanofibers as Effective Photocatalysts. Catalysts 2020, 10, 417. [Google Scholar] [CrossRef] [Green Version]
  21. Sakwises, L.; Pisitsak, P.; Manuspiya, H.; Ummartyotin, S. Effect of Mn-substituted SnO2 particle toward photocatalytic degradation of methylene blue dye. Results Phys. 2017, 7, 1751–1759. [Google Scholar] [CrossRef]
  22. Qin, Y.; Fang, F.; Xie, Z.; Lin, H.; Zhang, K.; Yu, X.; Chang, K. La,Al-Codoped SrTiO3 as a Photocatalyst in Overall Water Splitting: Significant Surface Engineering Effects on Defect Engineering. ACS Catal. 2021, 11, 11429–11439. [Google Scholar] [CrossRef]
  23. Mousa, H.M.; Alenezi, J.F.; Mohamed, I.M.; Yasin, A.S.; Hashem, A.-F.M.; Abdal-Hay, A. Synthesis of TiO2@ZnO heterojunction for dye photodegradation and wastewater treatment. J. Alloy. Compd. 2021, 886, 161169. [Google Scholar] [CrossRef]
  24. Ameur, N.; Bachir, R.; Bedrane, S.; Choukchou-Braham, A. A Green Route to Produce Adipic Acid on TiO2-Fe2O3 Nanocomposites. J. Chin. Chem. Soc. 2017, 64, 1096–1103. [Google Scholar] [CrossRef]
  25. Ahmad, R.; Kim, J.K.; Kim, J.H.; Kim, J. Effect of polymer template on structure and membrane fouling of TiO2 /Al2O3 composite membranes for wastewater treatment. J. Ind. Eng. Chem. 2018, 57, 55–63. [Google Scholar] [CrossRef]
  26. Pérez-Larios, A.; Rico, J.L.; Anaya-Esparza, L.M.; Vargas, O.G.; González-Silva, N.; Gómez, R. Hydrogen Production from Aqueous Methanol Solutions Using Ti–Zr Mixed Oxides as Photocatalysts under UV Irradiation. Catalysts 2019, 9, 938. [Google Scholar] [CrossRef] [Green Version]
  27. Chaker, H.; Ameur, N.; Saidi-Bendahou, K.; Djennas, M.; Fourmentin, S. Modeling and Box-Behnken design optimization of photocatalytic parameters for efficient removal of dye by lanthanum-doped mesoporous TiO2. J. Environ. Chem. Eng. 2021, 9, 104584. [Google Scholar] [CrossRef]
  28. Rozman, N.; Tobaldi, D.M.; Cvelbar, U.; Puliyalil, H.; Labrincha, J.A.; Legat, A.; Škapin, A.S. Hydrothermal Synthesis of Rare-Earth Modified Titania: Influence on Phase Composition, Optical Properties, and Photocatalytic Activity. Materials 2019, 12, 713. [Google Scholar] [CrossRef] [Green Version]
  29. Ray, S.K.; Dhakal, D.; Lee, S.W. Rapid degradation of naproxen by AgBr-α-NiMoO4 composite photocatalyst in visible light: Mechanism and pathways. Chem. Eng. J. 2018, 347, 836–848. [Google Scholar] [CrossRef]
  30. Weber, A.S.; Grady, A.M.; Koodali, R.T. Lanthanide modified semiconductor photocatalysts. Catal. Sci. Technol. 2012, 2, 683–693. [Google Scholar] [CrossRef]
  31. Baiju, K.; Sibu, C.; Rajesh, K.; Pillai, P.K.; Mukundan, P.; Warrier, K.; Wunderlich, W. An aqueous sol–gel route to synthesize nanosized lanthana-doped titania having an increased anatase phase stability for photocatalytic application. Mater. Chem. Phys. 2005, 90, 123–127. [Google Scholar] [CrossRef]
  32. Štengl, V.; Bakardjieva, S.; Murafa, N. Preparation and photocatalytic activity of rare earth doped TiO2 nanoparticles. Mater. Chem. Phys. 2009, 114, 217–226. [Google Scholar] [CrossRef]
  33. Sibu, C.P.; Kumar, S.R.; Mukundan, P.; Warrier, K.G.K. Structural Modifications and Associated Properties of Lanthanum Oxide Doped Sol-Gel Nanosized Titanium Oxide. Chem. Mater. 2002, 14, 2876–2881. [Google Scholar] [CrossRef]
  34. Song, L.; Zhao, X.; Cao, L.; Moon, J.-W.; Gu, B.; Wang, W. Synthesis of rare earth doped TiO2 nanorods as photocatalysts for lignin degradation. Nanoscale 2015, 7, 16695–16703. [Google Scholar] [CrossRef] [PubMed]
  35. Elgendy, K.; Elmehasseb, I.; Kandil, S. Synthesis and characterization of rare earth metal and non-metal co-doped TiO2 nanostructure for photocatalytic degradation of Metronidazole. Environment 2019, 2508, 60–67. [Google Scholar]
  36. Du, J.; Li, B.; Huang, J.; Zhang, W.; Peng, H.; Zou, J. Hydrophilic and photocatalytic performances of lanthanum doped titanium dioxide thin films. J. Rare Earths 2013, 31, 992–996. [Google Scholar] [CrossRef]
  37. Zhang, X.; Chen, W.; Lin, Z.; Yao, J.; Tan, S. Preparation and Photocatalysis Properties of Bacterial Cellulose/TiO2 Composite Membrane Doped with Rare Earth Elements. Synth. React. Inorganic Met. Nano-Metal Chem. 2011, 41, 997–1004. [Google Scholar] [CrossRef]
  38. Liu, J.; Yang, R.; Li, S. Synthesis and Photocatalytic Activity of TiO2/V2O5 Composite Catalyst Doped with Rare Earth Ions. J. Rare Earths 2007, 25, 173–178. [Google Scholar] [CrossRef]
  39. Zhu, L.; Xiao, Y.F.; Wang, X. The Photocatalytic Oxidation Properties of La3+-Doped and Ce4+-Doped ZnO-TiO2 in Treating Pharmaceutical Wastewater. Adv. Mater. Res. 2013, 726–731, 2988–2992. [Google Scholar] [CrossRef]
  40. Li, S.; Yang, Y.; Su, Q.; Liu, X.; Zhao, H.; Zhao, Z.; Li, J.; Jin, C. Synthesis and photocatalytic activity of transition metal and rare earth element co-doped TiO2 nano particles. Mater. Lett. 2019, 252, 123–125. [Google Scholar] [CrossRef]
  41. Ðorđević, V.; Milićević, B.; Dramićanin, M.D. Rare Earth-Doped Anatase TiO2 Nanoparticles. In Titanium Dioxide; InTech: London, UK, 2017. [Google Scholar]
  42. Smitha, V.S.; Saju, P.; Hareesh, U.S.; Swapankumar, G.; Warrier, K.G.K. Optical Properties of Rare-Earth Doped TiO2 Nanocomposites and Coatings; A Sol-Gel Strategy towards Multi-functionality. ChemistrySelect 2016, 1, 2140–2147. [Google Scholar] [CrossRef]
  43. Lan, X.; Wang, L.; Zhang, B.; Tian, B.; Zhang, J. Preparation of lanthanum and boron co-doped TiO2 by modified sol–gel method and study their photocatalytic activity. Catal. Today 2014, 224, 163–170. [Google Scholar] [CrossRef]
  44. Smitha, V.S.; Jyothi, C.K.; Mohamed, A.P.; Pillai, S.; Warrier, K.G. Novel multifunctional titania–silica–lanthanum phosphate nanocomposite coatings through an all aqueous sol–gel process. Dalton Trans. 2013, 42, 4602–4612. [Google Scholar] [CrossRef] [PubMed]
  45. Khedr, T.M.; El-Sheikh, S.M.; Hakki, A.; Ismail, A.A.; Badawy, W.A.; Bahnemann, D.W. Highly active non-metals doped mixed-phase TiO 2 for photocatalytic oxidation of ibuprofen under visible light. J. Photochem. Photobiol. A Chem. 2017, 346, 530–540. [Google Scholar] [CrossRef]
  46. Escudero, A.; Becerro, A.I.; Carrillo-Carrion, C.; Núñez, N.O.; Zyuzin, M.V.; Laguna, M.; González, D.; Ocaña, M.; Parak, W.J. Rare earth based nanostructured materials: Synthesis, functionalization, properties and bioimaging and biosensing applications. Nanophotonics 2017, 6, 881–921. [Google Scholar] [CrossRef]
  47. Sing, K.S.W.; Everett, D.H.; Haul, R.A.W.; Moscou, L.; Pierotti, R.A.; Rouquerol, J.; Siemieniewska, T. Reporting Physisorption Data for Gas/Solid Systems with Special Reference to the Determination of Surface Area and Porosity. Pure Appl. Chem. 1985, 57, 603–619. [Google Scholar] [CrossRef]
  48. Pérez-Larios, A.; Hernández-Gordillo, A.; Morales-Mendoza, G.; Lartundo-Rojas, L.; Mantilla, Á.; Gómez, R. Enhancing the H2 evolution from water–methanol solution using Mn2+–Mn+3–Mn4+ redox species of Mn-doped TiO2 sol–gel photocatalysts. Catal. Today 2016, 266, 9–16. [Google Scholar] [CrossRef]
  49. Pérez-Larios, A.; Torres-Ramos, I.; Zanella, R.; Rico, J.L. Ti-Co mixed oxide as photocatalysts in the generation of hydrogen from water. Int. J. Chem. React. Eng. 2022, 20, 129–140. [Google Scholar] [CrossRef]
  50. Kabir, H.; Nandyala, S.H.; Rahman, M.M.; Kabir, M.A.; Stamboulis, A. Influence of calcination on the sol–gel synthesis of lanthanum oxide nanoparticles. Appl. Phys. A Mater. Sci. Process 2018, 124, 1–11. [Google Scholar] [CrossRef]
  51. Pérez-Larios, A.; Lopez, R.; Hernández-Gordillo, A.; Tzompantzi, F.; Gómez, R.; Torres-Guerra, L.M. Improved hydrogen production from water splitting using TiO2–ZnO mixed oxides photocatalysts. Fuel 2012, 100, 139–143. [Google Scholar] [CrossRef]
  52. Santoso, J.S.; Nusantor, Y.R.; Lestari, W.C.; Choir, A.A. Review: Pengaruh Variasi Suhu Kalsinasi C-TiO2, Zn-TiO2, dan La-TiO2 Terhadap Kristalinitas pada Metode Sol-Gel. J. Chem. 2021, 10, 3. [Google Scholar]
  53. Zalas, M.; Laniecki, M. Photocatalytic hydrogen generation over lanthanides-doped titania. Sol. Energy Mater. Sol. Cells 2005, 89, 287–296. [Google Scholar] [CrossRef]
  54. McDevitt, N.T.; Baun, W.L. Infrared absorption study of metal oxides in the low frequency region (700–240 cm1). Spectrochim. Acta 1964, 20, 799–808. [Google Scholar] [CrossRef]
  55. Catauro, M.; Tranquillo, E.; Poggetto, G.D.; Pasquali, M.; Dell’Era, A.; Ciprioti, S.V. Influence of the Heat Treatment on the Particles Size and on the Crystalline Phase of TiO2 Synthesized by the Sol-Gel Method. Materials 2018, 11, 2364. [Google Scholar] [CrossRef] [Green Version]
  56. Ruíz-Santoyo, V.; Marañon-Ruiz, V.F.; Romero-Toledo, R.; González Vargas, O.A.; Pérez-Larios, A. Photocatalytic Degradation of Rhodamine B and Methylene Orange Using TiO2-ZrO2 as Nanocomposite. Catalysts 2021, 11, 1035. [Google Scholar] [CrossRef]
  57. Nguyen-Phan, T.-D.; Song, M.B.; Kim, E.J.; Shin, E.W. The role of rare earth metals in lanthanide-incorporated mesoporous titania. Microporous Mesoporous Mater. 2009, 119, 290–298. [Google Scholar] [CrossRef]
  58. Kumaresan, L.; Prabhu, A.; Palanichamy, M.; Arumugam, E.; Murugesan, V. Synthesis and characterization of Zr4+, La3+ and Ce3+ doped mesoporous TiO2: Evaluation of their photocatalytic activity. J. Hazard. Mater. 2011, 186, 1183–1192. [Google Scholar] [CrossRef]
  59. Zhang, Z.; Ni, X.; Yao, K.; Liu, Y.; Sun, Y.; Wang, X.; Huang, W.; Zhang, Y. An electrochemical synthesis of a rare-earth(La3+)-doped ZIF-8 hydroxyapatite composite coating for a Ti/TiO2 implant material. New J. Chem. 2021, 45, 6543–6549. [Google Scholar] [CrossRef]
  60. Briggs, D. Handbool of X Ray Photoelectron Spectroscopy: A Reference Book of Standard Spectra for Identification and Interpretation of XPS Data. In Handbook of Adhesion; John Wiley & Sons, Ltd.: Chichester, UK, 2005; pp. 621–622. [Google Scholar]
  61. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S.W. Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef] [Green Version]
  62. Walford, E. Handbook of Inorganic Chemicals; McGraw Hill: New York, NY, USA, 1896. [Google Scholar]
  63. Li, J.P.H.; Zhou, X.; Pang, Y.; Zhu, L.; Vovk, E.I.; Cong, L.; van Bavel, A.P.; Li, S.; Yang, Y. Understanding of binding energy calibration in XPS of lanthanum oxide by in situ treatment. Phys. Chem. Chem. Phys. 2019, 21, 22351–22358. [Google Scholar] [CrossRef]
  64. Pérez-Larios, A.; Torres-Ramos, M.; González-Vargas, O.; Rico, J.L.; Zanella, R. Ti–Fe mixed oxides as photocatalysts in the generation of hydrogen under UV-light irradiation. Int. J. Hydrogen Energy 2022. [Google Scholar] [CrossRef]
  65. Scoca, D.L.S.; Cemin, F.; Bilmes, S.A.; Figueroa, C.A.; Zanatta, A.R.; Alvarez, F. Role of Rare Earth Elements and Entropy on the Anatase-To-Rutile Phase Transformation of TiO2 Thin Films Deposited by Ion Beam Sputtering. ACS Omega 2020, 5, 28027–28036. [Google Scholar] [CrossRef] [PubMed]
  66. Manjumol, K.A.; Smitha, V.S.; Shajesh, P.; Baiju, K.V.; Warrier, K.G.K. Synthesis of lanthanum oxide doped photocatalytic nano titanium oxide through aqueous sol–gel method for titania multifunctional ultrafiltration membrane. J. Sol-Gel Sci. Technol. 2010, 53, 353–358. [Google Scholar] [CrossRef]
  67. Eslami, A.; Amini, M.M.; Asadi, A.; Safari, A.A.; Daglioglu, N. Photocatalytic degradation of ibuprofen and naproxen in water over NS-TiO2 coating on polycarbonate: Process modeling and intermediates identification. Inorg. Chem. Commun. 2020, 115, 107888. [Google Scholar] [CrossRef]
  68. Fan, G.; Zhan, J.; Luo, J.; Zhang, J.; Chen, Z.; You, Y. Photocatalytic degradation of naproxen by a H2O2-modified titanate nanomaterial under visible light irradiation. Catal. Sci. Technol. 2019, 9, 4614–4628. [Google Scholar] [CrossRef]
  69. Regmi, C.; Kshetri, Y.K.; Pandey, R.P.; Lee, S.W. Visible-light-driven S and W co-doped dendritic BiVO 4 for efficient photocatalytic degradation of naproxen and its mechanistic analysis. Mol. Catal. 2018, 453, 149–160. [Google Scholar] [CrossRef]
  70. Méndez-Arriaga, F.; Gimenez, J.; Esplugas, S. Photolysis and TiO2 Photocatalytic Treatment of Naproxen: Degradation, Mineralization, Intermediates and Toxicity. J. Adv. Oxid. Technol. 2008, 11, 435–444. [Google Scholar] [CrossRef]
  71. Begum, S. Ahmaruzzaman Biogenic synthesis of SnO2/activated carbon nanocomposite and its application as photocatalyst in the degradation of naproxen. Appl. Surf. Sci. 2018, 449, 780–789. [Google Scholar] [CrossRef]
Figure 1. SEM micrographs and EDS spectra of the Ti-La nanocomposites.
Figure 1. SEM micrographs and EDS spectra of the Ti-La nanocomposites.
Inorganics 10 00067 g001aInorganics 10 00067 g001b
Figure 2. N2 isotherms and pore size distribution of Ti-La samples.
Figure 2. N2 isotherms and pore size distribution of Ti-La samples.
Inorganics 10 00067 g002
Figure 3. X-ray diffractograms of Ti-La nanocomposites.
Figure 3. X-ray diffractograms of Ti-La nanocomposites.
Inorganics 10 00067 g003
Figure 4. Band gap energies of the Ti-La samples determined from the UV-Vis spectra.
Figure 4. Band gap energies of the Ti-La samples determined from the UV-Vis spectra.
Inorganics 10 00067 g004
Figure 5. FT-IR spectra of TiO2 and Ti-La nanocomposites.
Figure 5. FT-IR spectra of TiO2 and Ti-La nanocomposites.
Inorganics 10 00067 g005
Figure 6. XPS spectra for Ti 2p.
Figure 6. XPS spectra for Ti 2p.
Inorganics 10 00067 g006
Figure 7. XPS spectra for La 3d.
Figure 7. XPS spectra for La 3d.
Inorganics 10 00067 g007
Figure 8. XPS spectra for O1s.
Figure 8. XPS spectra for O1s.
Inorganics 10 00067 g008
Figure 9. XPS spectra for C 1s.
Figure 9. XPS spectra for C 1s.
Inorganics 10 00067 g009
Figure 10. TEM images of photocatalysts, where (a) TiO2, (b) Ti-La1, (c) Ti-La3, (d) Ti-La5, and (e) Ti-La10.
Figure 10. TEM images of photocatalysts, where (a) TiO2, (b) Ti-La1, (c) Ti-La3, (d) Ti-La5, and (e) Ti-La10.
Inorganics 10 00067 g010
Figure 11. Naproxen degradation profiles in presence of TiO2 or Ti-La photocatalysts under UV irradiation. Degussa P25 is used as a reference.
Figure 11. Naproxen degradation profiles in presence of TiO2 or Ti-La photocatalysts under UV irradiation. Degussa P25 is used as a reference.
Inorganics 10 00067 g011aInorganics 10 00067 g011b
Figure 12. Recyclability of TiO2 and Ti-La nanocomposites in the photocatalytic degradation of NPX under UV irradiation. The activities were determined after 4 h of reaction.
Figure 12. Recyclability of TiO2 and Ti-La nanocomposites in the photocatalytic degradation of NPX under UV irradiation. The activities were determined after 4 h of reaction.
Inorganics 10 00067 g012
Table 1. Titania-based nanocomposites with rare earths and their photocatalytic activity.
Table 1. Titania-based nanocomposites with rare earths and their photocatalytic activity.
CatalystContaminant Type of LightReaction Time (min)EfficiencyReference
Ti-LaMethylene blueUV irradiation2020%[28]
Ti-CeMethylene blue585 nm20<10%[28]
Ti-GdMethylene blueUV irradiation2020%[28]
Ti-CeMetronidazoleVisible light illumination12050%[35]
Ti-Ce/N/CMetronidazoleVisible light illumination120100%[35]
Ti-Co/Ce/La/Eu/SmMethylene blue440–550 nm30>90%[40]
TiO2Methylene blue440–550 nm30<50%[40]
Ti-NdMethyl orangeSimulated solar radiation (SSR)240>90%[41]
Ti-EuMethyl orangeSSR240<70%[41]
Ti-TbMethyl orangeSSR240<40%[41]
TiO2Methyl orangeSSR240<50%[41]
Ti-LaMethylene blue484 nm120<70%[42]
Ti-La/TbMethylene blue484 nm12080%[42]
Ti-EuMethyl orangeSimulated sunlight radiation20>95%[34]
P25 TiO2Methyl orangeSimulated sunlight radiation20<50%[34]
Ti-SmLigninSimulated sunlight radiation90>95%[34]
P25 TiO2LigninSimulated sunlight radiation90<80%[34]
Ti-LaAzo-dye acid orange 7>420 nm300<80%[43]
Ti-La-BAzo-dye acid orange 7>420 nm300>90%[43]
Ti-Si-LaMethylene blueUV irradiation12094%[44]
TiO2Methylene blueUV irradiation120<60%[44]
Ti/bc-LaMethyl orangeUV irradiation300>85%[37]
Ti/bc-CeMethyl orangeUV irradiation300>95%[37]
Ti/bcMethyl orangeUV irradiation300<60%[37]
C,N,S-TiO2IbuprofenVisible light300>99%[45]
Table 2. Physical and textural properties of Ti-La samples.
Table 2. Physical and textural properties of Ti-La samples.
SamplesBET (m2/g)Pore
Diameter (nm)
Cell
Parameter a (Å)
Crystallite Size, D (nm)Eg (eV)
Ti-La1112.714.323.7909.222.92
Ti-La3127.812.953.7926.522.99
Ti-La5135.812.073.7966.332.98
Ti-La10144.68.993.8205.213.01
TiO264.06.53.79020.583.05
Table 3. Surface chemical states of Ti and La, and compositional analysis determined by XPS.
Table 3. Surface chemical states of Ti and La, and compositional analysis determined by XPS.
Element Binding Energies in eV, and Chemical State
TiO2Ti-La1Ti-La3Ti-La5Ti-La10
Ti 2p3/2 459.25459.31459.25457.78459.25
Chemical stateTi4+Ti4+ + Ti3+Ti4+ + Ti3+Ti4+ + Ti3+Ti4+ + Ti3+
Ti 2p1/2 464.96464.99464.73463.53464.76
Chemical stateTi4+Ti4+ + Ti3+Ti4+ + Ti3+Ti4+ + Ti3+Ti4+ + Ti3+
La 3d5/2-835.89835.18833.72835.18
La 3d3/2-854.96852.20850.76852.20
Chemical state-La3+La3+La3+La3+
Weight% La-1.443.525.438.58
Table 4. Decomposition and mineralization of NPX after 4 h in presence of catalysts and under UV.
Table 4. Decomposition and mineralization of NPX after 4 h in presence of catalysts and under UV.
Sample% DegradationStandard
Deviation
% MineralizationStandard
Deviation
Ti-La199.55 ±0.2099.65±0.08
Ti-La393.76 ±0.2493.78±0.29
Ti-La590.13 ±0.2993.22±0.47
Ti-La1093.45 ±0.2994.45±0.48
TiO297.90 ±0.1999.64±0.36
Photolysis45.12 ±0.3918.92±0.39
Table 5. Degradation of NPX by heterogeneous photocatalysts.
Table 5. Degradation of NPX by heterogeneous photocatalysts.
PhotocatalystPharmaceutical CompoundType of
Irradiation
Reaction Time (min)% DegradationReference
H2O2 modified titanate nanomaterialNaproxenVisible light 18099.9[67]
BiVO4NaproxenVisible light30080%[68]
AgBr-α-NiMoO4NaproxenVisible light2084[29]
TiO2NaproxenXe-lamp18040[69]
SnO2/ACNaproxenDirect sunlight12094[70]
P25-TiO2NaproxenVisible light60094[71]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Marizcal-Barba, A.; Limón-Rocha, I.; Barrera, A.; Casillas, J.E.; González-Vargas, O.A.; Rico, J.L.; Martinez-Gómez, C.; Pérez-Larios, A. TiO2-La2O3 as Photocatalysts in the Degradation of Naproxen. Inorganics 2022, 10, 67. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics10050067

AMA Style

Marizcal-Barba A, Limón-Rocha I, Barrera A, Casillas JE, González-Vargas OA, Rico JL, Martinez-Gómez C, Pérez-Larios A. TiO2-La2O3 as Photocatalysts in the Degradation of Naproxen. Inorganics. 2022; 10(5):67. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics10050067

Chicago/Turabian Style

Marizcal-Barba, Adriana, Isaias Limón-Rocha, Arturo Barrera, José Eduardo Casillas, O. A. González-Vargas, José Luis Rico, Claudia Martinez-Gómez, and Alejandro Pérez-Larios. 2022. "TiO2-La2O3 as Photocatalysts in the Degradation of Naproxen" Inorganics 10, no. 5: 67. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics10050067

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