Next Article in Journal
Centrifugal Partition Chromatography Is a Powerful Tool for the Isolation of Antibiofilm Quantum Carbon Dots Synthesized by Hydrothermal Treatment of Avocado Peels
Next Article in Special Issue
Visible Light Photo-Fenton with Hybrid Activated Carbon and Metal Ferrites for Efficient Treatment of Methyl Orange (Azo Dye)
Previous Article in Journal
Excess Properties, FT-IR Spectral Analysis, and CO2 Absorption Performance of Monoethanolamine with Diethylene Glycol Monoethyl Ether or Methyldiethanolamine Binary Solutions
Previous Article in Special Issue
Advances in Catalyst Design for β-Lactone Formation via Ring-Expansion Carbonylation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

TiO2/LaFeO3 Composites for the Efficient Degradation of Benzoic Acid and Hydrogen Production

by
Isabella Natali Sora
1,*,
Benedetta Bertolotti
1,
Renato Pelosato
1,
Andrea Lucotti
2,
Matteo Tommasini
2 and
Marica Muscetta
3
1
Dipartimento di Ingegneria e Scienze Applicate, Università di Bergamo, Viale Marconi 5, 24044 Dalmine, Italy
2
Dipartimento di Chimica, Materiali e Ingegneria Chimica “G. Natta”, Politecnico di Milano, Piazza Leonardo da Vinci 32, 20133 Milano, Italy
3
Dipartimento di Ingegneria Chimica, dei Materiali e della Produzione (DICMaPI), Università di Napoli Federico II, Piazzale V. Tecchio 80, 80125 Naples, Italy
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(7), 1526; https://doi.org/10.3390/molecules30071526
Submission received: 14 March 2025 / Revised: 26 March 2025 / Accepted: 27 March 2025 / Published: 29 March 2025
(This article belongs to the Special Issue Research on Heterogeneous Catalysis—2nd Edition)

Abstract

:
LaFeO3/TiO2 composites were prepared in the range 0–12.2 wt% of LaFeO3, characterized, and tested for both benzoic acid (BA) and 4-methoxycinnamic acid (MCA) degradation in aqueous solution, and hydrogen evolution. The preparation method was via ball-milling without thermal treatment. The composite materials presented agglomerates of LaFeO3 with an average size from 1 to 5 μm, and the TiO2 powder was well dispersed onto the surface of each sample. They showed varying activities for BA degradation depending on composition and light wavelength. The 6.2 wt% and 12.2 wt%-LaFeO3/TiO2 composites exhibited the highest activity under 380–800 nm light and could degrade BA in 300 min at BA concentration 13.4 mg L−1 and catalyst 0.12 g L−1. Using a 450 nm LED light source, all composites degraded less than 10% of BA, but in the presence of H2O2 (1 mM) the photocatalytic activity was as high as 96% in <120 min, 6.2 wt%-LaFeO3/TiO2 composite being the most efficient sample. It was found that in the presence of H2O2, BA degradation followed first order kinetic with a reaction rate constant of 4.8 × 10−4 s−1. The hydrogen production rate followed a classical volcano-like behavior, with the highest reactivity (1600 μmol h−1g−1 at 60 °C) in the presence of 3.86%wt- LaFeO3/TiO2. It was also found that LaFeO3/TiO2 exhibited high stability in four recycled tests without losing activity for hydrogen production. Furthermore, a discussion on photogenerated charge-carrier transfer mechanism is briefly provided, focusing on lacking significant photocatalytic activity under 450 nm light, so p-n heterojunction formation is unlikely.

Graphical Abstract

1. Introduction

Many semiconductor oxides can act as photocatalysts [1,2,3,4,5]. Their activation occurs by irradiation with photons of energy that match the band gap energy of the semiconductor, resulting in electron excitation from the valence band (VB) to the conduction band (CB), thereby generating holes in the VB. Due to band bending, the photogenerated e - h + pairs diffuse to the semiconductor surface where, in the absence of electron and hole scavengers, their recombination takes place in a few nanoseconds. Recombination is avoided if an adequate scavenger or surface defect states are available to trap electron-hole pairs. In this case, water adsorbed or hydroxyl groups present on the semiconductor surface can be oxidized by h + , generating H O   ( H 2 O + h +     H O + H + ), while adsorbed O 2   a d s can be reduced by e , generating superoxide radicals O 2 ( O 2   a d s + e     O 2 ) [6]. The photocatalytic processes exploit the produced radicals to start a chemical reaction or to increase the reaction rate.
The broad utilization of semiconductor photocatalysts for organic pollutant decomposition in wastewater treatment plants has been limited by the high recombination of the photogenerated electron–hole pairs, resulting in low quantum efficiency of photocatalyzed reactions. Titanium dioxide (TiO2) is one of the most employed materials among photocatalysts [7,8]. It shows high activity when exposed to light and excellent chemical stability. Anatase TiO2 has an indirect and relatively large bandgap of Eg = 3.2 eV; for activation it requires the use of UV light which is more expensive and dangerous than radiation in the visible range [9]. Because of these challenges, nanocomposite catalysts in which one component consists of anatase TiO2 have been extensively studied with two main objectives: to improve the separation of photogenerated electron and hole pairs, and to obtain a TiO2-based photocatalyst also active under visible light irradiation.
In recent years, different types of materials (e.g., ceramic, metallic, carbon-based, and polymeric nanomaterials) have been combined with TiO2 to form functional nanocomposites, exhibiting attractive chemical and physical properties as compared to any of the components separately. LaFeO3 is a stable and non-toxic semiconductor with a p-type low band gap of 2.4–2.5 eV [10]. The use of LaFeO3 in photocatalytic degradation of persistent organic pollutants in wastewater, i.e., pharmaceuticals, has been previously investigated by us [11,12,13].
Although there are studies showing the visible-light photocatalytic degradation of (i) dyes by LaFeO3/TiO2 nanocomposite [14] and anatase-TiO2/rutile-TiO2/LaFeO3 three-phase nanostructured heterojunctions [15]; (ii) phenol by LaFeO3/TiO2 nanocomposite [16]; (iii) antibiotics by LaFeO3/TiO2 heterostructures [17] and by carbon quantum dot modified TiO2@LaFeO3 hollow core shell [18]; and (iv) pesticides by LaFeO3/TiO2 heterojunction [19] and core shell [20], as far as we know the photocatalytic degradation of aromatic carboxylic acids such as benzoic acid using LaFeO3/TiO2 has never been reported previously. Hydrothermal or wet chemical synthesis are the most common strategies for the preparation of these composites [14,16].
This study is focused on the photocatalytic performance of LaFeO3/TiO2 nanocomposites, (with 0–12.2 wt% of LaFeO3) prepared via a fast and simple ball-milling method without thermal treatment. Benzoic acid (BA) was chosen as a model compound suitable for investigating the photocatalytic behavior of aromatic acids in an aqueous environment. The photocatalytic activity was evaluated for the degradation of BA by UV-vis spectrophotometry and high-pressure liquid chromatography (HPLC). The photocatalytic activity of the composites was evaluated with and without hydrogen peroxide addition. Presently, H2 is mainly produced by steam methane reforming, which involves the use of fossil fuels and high temperature, resulting in low-purity hydrogen products and greenhouse gas emissions. Alternative more sustainable processes to produce high-purity H2 are via electrochemical and photochemical water splitting [21,22]. For better evaluation of the nanocomposites, the photocatalytic hydrogen production rate using methanol as a sacrificial agent was studied. The photocatalytic hydrogen evolution performance of LaFeO3/TiO2 has been the subject of a few investigations. Lv et al. [23] synthesized 20 wt%-LaFeO3/TiO2 nanocomposites by a 2-step approach and reported that they exhibited a H2 production rate of 0.279 mmol h−1 g−1 under the simulated solar light irradiation of a 300 W Xe lamp (320–780 nm). Comparison of photocatalytic water splitting H2 evolution rate of 20 wt%-LaFeO3/TiO2 without using methanol as a sacrificial agent gives 0.0178 mmol h−1 g−1 and no hydrogen generation was observed in the presence of bare TiO2 or LaFeO3. Recently, a 49.64 wt%-LaFeO3/TiO2 powder was prepared by Jiang et al. [24] via solid state method, annealing the mixture at 500 °C for 12 h in Ar atmosphere. The powder showed a high photocatalytic H2 production rate of ~3.26 mmol h−1 g−1 under AM 1.5 simulated sunlight, which is approximately 2.28 times that of commercial TiO2 P25 (1.43 mmol h−1 g−1).

2. Results and Discussion

2.1. Phase Composition and Microstructure

As shown from X-ray diffraction (XRD) data, the LaFeO3 powder (sample F) is monophasic, and the diffraction peaks can be indexed with orthorhombic ABO3-type perovskite structure (JCPDS file No. 37-1493). The broad peaks suggest that obtained products have nanometric size. The XRD patterns of the composites B (1.3 wt%-LaFeO3/TiO2), C (3.9 wt%-LaFeO3/TiO2), D (6.2 wt%-LaFeO3/TiO2) and E (12.2 wt%-LaFeO3/TiO2) are presented in Figure 1a. Diffraction peaks belonging to the LaFeO3, TiO2 anatase (JCPDS card no. 21-1272), and TiO2 rutile (JCPDS card no. 21-1276) phases were detected in all composites. By increasing the LaFeO3 content (i.e., from B to E), the intensity of the orthoferrite diffraction peaks increases. No reaction products were observed in the XRD patterns of the composites. The XRD patterns confirm that the composites are physical mixtures of the two components as expected due to the short milling time. SEM micrographs of C, D and E composites are shown in Figure 1b–d. For brevity, other micrographs of the samples are shown in the Supporting Information (Figures S1 and S2). As shown in Figure S1, LaFeO3 powder is nanometric, very porous, and tends to agglomerate to larger particles, causing fast charge recombination. In the SEM micrographs of C, D and E, the same microstructure is evidenced: the agglomerates of LaFeO3 have an average size in the 1 to 5 μm range and the TiO2 powder is well dispersed onto the surface of each sample. This characteristic may have some relationship with their good performance on the photocatalytic degradation of benzoic acid.
To investigate the optical absorption characteristics of LaFeO3, nanoparticles at different wavelengths diffuse reflectance spectroscopy (DRS) analytical techniques were used. The UV–vis DRS spectrum (Figure S3) of LaFeO3 nanoparticles shows the absorption edge occurs at about 600 nm, attributed to its electron transitions from valence band (mainly mixed eg states of Fe 3d and O 2p) to conduction band (t2g states of Fe 3d) [25]. A broad absorption band ranging from 200 to 600 nm indicates its potential applications as visible light-driven photocatalyst. LaFeO3 optical band-gap energy was 2.47 eV.

2.2. Benzoic Acid Degradation Using Daylight Lamps in LaFeO3/TiO2 Water Suspensions

Figure 2 shows the variation of the concentration of BA during the reactions over different catalysts, without an oxidizing agent such as H2O2, under visible light and a fraction of UV-A (380–800 nm). According to HPLC data, the LaFeO3/TiO2 composites degrade up to 93% of BA in 300 min. Pure TiO2 (sample A) and LaFeO3 (sample F) degraded 80% and 9% in 300 min, respectively. In Table 1 these results are compared with those of other LaFeO3/TiO2 photocatalysts on organic pollutant degradation using UV or visible light. The experimental studies are performed with a catalyst concentration in the range 0.8–2.4 g L−1, larger than the one used in this study (0.12 g L−1).
TiO2 powder covers LaFeO3 agglomerates, so it is likely that the BA adsorption on composites is similar to that of TiO2. BA adsorption depends on the surface charge of TiO2 and on the extent of BA deprotonation (Ka = 6.46 × 10−5 at 25 °C). The TiO2 surface is positively charged at pH below the point of zero charge at pHzpc ≈ 6 [26]. At natural pH, pH ≈ 4.2–4.6 in A to E suspensions, BA adsorption should be favored through electrostatic interactions between the negatively charged benzoate ion C6H5COO in the water matrix and the –OH2+ groups on the TiO2 surface. In suspension F, pH is ≈6.1, and the benzoate would be attracted to the most positively charged LaFeO3 surface (pHzpc = 8.9) through electrostatic interactions. It seems that no direct relationship can be established between the pHzpc values and the photoactivity of the samples. However, given the good performance of the D and E composites in the degradation of BA, it could be assumed that BA is more easily degraded under a relatively weak acidic pHzpc value, since it is a weak acid.
Another factor that could influence photocatalytic activity is the surface area of the catalysts. Perovskite oxides usually have small surface areas and large particle sizes. In particular, sample F (LaFeO3) presents a specific surface area of 14 m2 g−1 significantly smaller than that of bare TiO2 (48 m2 g−1), which is detrimental to the number of the available surface-active sites. Since the effective separation of the photogenerated e - h + pairs are a key aspect in photocatalytic reactions, the large particle size of LaFeO3 can have some relationship with a difficult migration of electrons or holes onto the surface of LaFeO3 to react with adsorbed BA rather than to recombine with each other. According to the mechanism described above, it is not difficult to understand that TiO2-containing composites show a higher photocatalytic activity compared to pure LaFeO3.
As reported in previous studies [27] (Ajmera et al., 2002 [28]), the photocatalytic degradation of BA proceeds through HO radical attack preferentially on the ortho and para positions, producing hydroxy-benzoic acids and dihydroxy-benzoic acids, according to HPLC analysis. The intermediate species go through photocatalytic degradation in competition with BA, slowing down the BA degradation. As shown in Figure 3, a second hydroxylation and subsequent ring opening is likely to follow but the species expected to evolve from here would be small molecule organics (i.e., methanol) that would rapidly degrade to carbon dioxide.
The decrease of benzoic acid and the formation of salicylic acid after one hour of irradiation was assessed qualitatively using fluorescence spectroscopy. Figure 4 illustrates the temporal changes in the emission spectra of benzoic acid in aqueous solutions irradiated for one hour in the presence of a fixed amount of the photocatalysts A (Figure 4a), E (Figure 4b), and F (Figure 4c). Under identical conditions, benzoic acid emits significantly less than salicylic acid [29]. The small emission peak of benzoic acid at approximately 300 nm decreases after one hour of irradiation in the solutions prepared with A and E but not the one prepared with F. This is consistent with the HPLC data shown in Figure 2. For the photocatalysts A and E, after one hour of illumination, a new emission band appears at about 410 nm, which is a typical feature of salicylic acid excited at a wavelength of 230 nm [29]. Therefore, it is clear that salicylic acid is one of the intermediates produced during the degradation of benzoic acid by OH radical attack.
To assess the generality of the enhanced activity of the LaFeO3/TiO2 catalysts, the degradation of another aromatic carboxylic acid compound, namely 4-methoxycinnamic acid (MCA), was studied under previous experimental conditions. MCA undergoes to rapid and reversible trans-cis photoisomerization under UV radiation [30]. As shown in Figure S4, up to 98% of MCA is degraded in the presence of composite catalysts after 120 min of irradiation, better than pure TiO2 catalyst. To degrade 90% of the aromatic carboxylic acids the 6.2 wt% and 12.2 wt%-LaFeO3/TiO2 composites take <120 min in the case of MCA and 300 min for BA. Also, sample F shows a better performance with MCA substrate (72% in 120 min) as compared to BA (<10% in 300 min). As expected, the degradation of MCA is much faster than that of BA due to different degradation mechanisms. It is well known that the principal degradation pathway for cinnamic acid involves attack at the alkene group by a super oxide radical, generating benzaldehyde as the main intermediate. Benzaldehyde undergoes attack by hydroxyl radicals, generating OH substituted intermediates that are quickly decomposed to CO2 [31].

2.3. Benzoic Acid Degradation Under 450 nm Light in LaFeO3/TiO2 Water Suspensions With and Without H2O2

The photocatalytic degradation of BA under monochromatic 450 nm light over different catalysts, with and without H2O2, was evaluated, and the results are shown in Figure 5. The results evidence that without H2O2 the bare photocatalysts and all LaFeO3/TiO2 composites degraded less than 10% of BA (Figure 5a). As shown in Table 1, Garcia-Muñoz et al. also reported a far smaller activity of the composite photocatalysts under >420 nm when compared to the reactions performed under solar light [19]. On the contrary, tests with dyes show high photocatalytic activity of LaFeO3/TiO2 composites under visible light [14]. However, tests with dyes (or with any substrate absorbing radiation) do not distinguish between a pure photocatalytic process and a dye-sensitized one, the latter being dye-specific and not general [32]. This could be the reason why pure LaFeO3 and LaFeO3/TiO2 composites, while demonstrated to work well in previous studies using dyes, were actually poorly effective without H2O2, as shown in the present work.
In the presence of H2O2 (1 mM), photocatalytic activity was high, and the most efficient samples were D, B and A. The pseudo-first-order rate constants (k) for BA degradation were determined for the initial reaction period (<120 min) and the values are reported in Table 2. The D composite shows the highest reaction rate constant (4.8 × 10−4 s−1), which is 1.2 and 3.2 times faster than those of TiO2 (4.1 × 10−4 s−1) and LaFeO3 (1.5 × 10−4 s−1), respectively. The radical intermediate HO formed from H2O2 by reaction with the photogenerated electrons can act as an electron scavenger, thus inhibiting the e-h+ pairs recombination at the semiconductor surface [33] according to the following equation:
H2O2 + eCB → HO + OH
where eCB indicates electron excited to the conduction band.
To further the study of the scarce activity of all photocatalysts without H2O2 under 450 nm light, degradation runs of 22 h were carried out. The results indicate that depending on composition, 30–60% of BA is degraded by the LaFeO3/TiO2 nanocomposites.

2.4. Hydrogen Evolution

Preliminary tests were performed by comparing the hydrogen evolution rate (HER) under UV-visible light radiation in the presence of pure TiO2 (P25), pure LaFeO3 and the composite materials B (1.3 wt%-LaFeO3/TiO2), C (3.9 wt%-LaFeO3/TiO2), D (6.2 wt%-LaFeO3/TiO2) and E (12.2 wt%-LaFeO3/TiO2), prepared by manual mixing of the two powders. Specifically, no hydrogen generation was observed in the presence of pure LaFeO3, due to the unsuitable band position of this photocatalyst (see Section 2.5), while the hydrogen amount was increased by 25% in the presence of the composite material when 1%wt. of LaFeO3 was used. Based on these results, it was chosen to evaluate the effect of LaFeO3 on HER, in the presence of methanol as the sacrificial agent.
As indicated in Figure 6, the hydrogen production rate follows a classical volcano-like behavior, with the highest reactivity in the presence of 3.86%wt of LaFeO3 (sample C). Beyond this value, the reduction of HER can be probably be ascribed to the excessive coverage of titanium dioxide.
Furthermore, these materials were tested to assess the photocatalytic activity under visible light radiation. In this case, no hydrogen generation was observed probably due to (i) the wide band gap of titanium dioxide, as well as (ii) the unsuitable band gap position of LaFeO3 (see the next section). Some experiments were conducted to assess how the system’s temperature impacts photocatalytic hydrogen production. Indeed, temperature does not directly affect the photocatalytic process, as it primarily depends on the radiation wavelength and the material’s absorption properties. However, the system’s temperature may influence photocatalytic hydrogen generation by altering the adsorption behavior of organic species on the catalyst surface and modifying the reaction kinetics [34]. As illustrated in Figure 7, the amount of hydrogen produced increases as the system’s temperature rises with a HER 4.5-fold higher at the maximum tested temperature (i.e., 60 °C), with respect to the lowest temperature (i.e., 25 °C).
Finally, photostability and reusability tests (Figure 8A,B) demonstrated good photostability both (i) during a single experiment carried out up to 8 h of irradiation, and (ii) during 4 photocatalytic cycles.

2.5. Activation of LaFeO3/TiO2 Nanocomposites

Different types of photogenerated charge-carrier transfer mechanism have been proposed for LaFeO3/TiO2 composites: type-II double charge transfer mechanism [14], direct Z-scheme mechanism [23], S-scheme [20], and uncommon heterojunction [16]. The formation of a p–n heterojunction between TiO2 (n-type semiconductor) and LaFeO3 (p-type semiconductor) could avoid e - h + recombination and improve the activity. The valence band potential of a semiconductor at the point of zero charge can be calculated by the following empirical equation:
EVB = X − Ee + 0.5Eg
where EVB is the VB edge potential, Χ is the electronegativity of the semiconductor, which is the geometric mean of the electronegativity of the constituent atoms, Ee is the energy of free electrons on the hydrogen scale (about 4.5 eV), and Eg is the band gap energy of the semiconductor [35]. The conduction band potential ECB can be determined by the equation ECB = EVB − Eg. The X values for LaFeO3 and TiO2 (anatase) are about 5.57 and 5.78 eV, respectively. The calculated ECB and EVB edge potentials for LaFeO3 are +0.1 eV and +2.3 eV vs. SHE, as shown in Figure 9. For TiO2, ECB and EVB are −0.3 eV and +2.9 eV vs. SHE, in agreement with previous studies [17,20]. Before contact between the semiconductors, the calculated EVB of LaFeO3 is under that of the TiO2 and ECB is above that of the TiO2. The creation of a p–n heterojunction would modify the electronic band structure of both semiconductors. Since composites do not have significant photocatalytic activity under visible light in these samples, p–n heterojunction formation is quite unlikely, so a high rate of recombination of photogenerated charge is present.
In our study, the influence of the LaFeO3 content was as follows: (i) under pure visible light better results are achieved by samples with a lower amount of LaFeO3 (B and C); (ii) under visible light and a fraction of UV-A radiation, samples with higher wt% of LaFeO3 (D and E) are more active; (iii) finally, the composites analyzed were not able at all to produce hydrogen under visible light. This behavior can suggest different light-mediated reaction mechanisms depending on the light wavelengths. Also, Garcia-Muñoz et al. reported the influence of the LaFeO3 content on degradation of myclobutanil, but the optimum was observed at 5 wt% under solar light and 12.5 wt% under visible light, respectively [19]. The different behavior it is probably related to the core-shell system study by Garcia-Muñoz et al., while our composites are physical mixtures of the two components where TiO2 particles are well dispersed on LaFeO3 agglomerates.

3. Materials and Methods

3.1. Chemicals and Preparations

Reactants for synthesis and chemicals for the photocatalytic reactions were purchased from Sigma-Aldrich (Saint Louis, MO, USA): La2O3, Fe(NO3)3·9H2O, citric acid hydrate (C6H8O7·xH2O, 99.5%), H2O2 (>30%), nitric acid and NH3 (30% aqueous solution). Lanthanum ferrite LaFeO3 nanopowders were prepared using the citrate auto-combustion method. A specific amount of dried La2O3 was dissolved in a nitric acid solution to prepare La(NO3)3∙6H2O. Stoichiometric amounts of metal nitrates were dissolved in water (0.1 mol L−1) by stirring on a hotplate. The solution was poured in citric acid solution (molar ratio of metal ions to citric was 1:1). Aqueous NH3 was added until the pH was 6.8. The solution was then dehydrated until a brown/orange gel formed. Dry gel was heated in air to 250 °C to start ignition. The powder was then calcined at 600 °C (5 °C min−1) in air for 3 h to decompose ammonium nitrate residues. A fast and simple procedure was employed to prepare the LaFeO3/TiO2 photocatalysts [36]. The method involved manual mixing of fixed quantities of TiO2 (P25, nanopowder, mean particle diameter <50 nm, Sigma Aldrich) and LaFeO3 in a vial. The wt% of LaFeO3 to TiO2 was 0 (sample A), 1.3 (sample B), 3.9 (sample C), 6.2 (sample D), 12.2 (sample E), and 100 (sample F) wt%. Then, a ball-milling method with an agate mortar and agate balls in a planetary mill was applied (PM100, RETSCH, Haan, Germany). The rotational speed (rpm) and milling time (min) were fixed at 1 min and 200 rpm, based on the previous studies on similar systems.

3.2. Microstructural Characterization

The powder X-ray diffraction patterns of the catalysts were recorded using a Bruker (Milano, Italy) D8-Advance powder diffractometer equipped with a Cu kα X-ray source and a Lynxeye XE-T® solid-state detector. The patterns were recorded in an interval of 10–90° 2θ with a step of 0.01° and a counting time of 1 s per step. Brunauer–Emmett–Teller (BET) specific surface area (SSA) determination was performed with nitrogen absorption on about 500 mg of the samples using a Micrometric (Norcross, GA, USA) Tristar 3000 automated gas-adsorption analyzer.
Scanning electron microscopy (SEM) was performed using a FEI (Hillsboro, OR, USA) Quanta 200 ESEM microscope, operating at 20 kV on specimens from which a thin layer of gold had evaporated. On the other hand, an electron microprobe used in an energy dispersive mode (EDX) was employed to obtain information on the actual metal–content ratio present in the samples. The UV-Vis diffuse reflectance spectroscopy (DRS) spectrum of LaFeO3 powder was recorded in the 200–800 nm range by using a Jasco (Tokyo, Japan) V-650 spectrophotometer equipped with an integrating sphere for solid samples, with BaSO4 as the reference sample. The plot of the Kubelka–Munk function (F(R) × E)2 vs. photon energy (E) suggests a direct bandgap transition of 2.47 eV for LaFeO3

3.3. Florescence Spectroscopy

The fluorescence spectra reported in this work were recorded with a JASCO FP 6600 spectrofluorometer selecting the excitation wavelength of 230 nm, which corresponds to the experimental condition used to investigate photodegradation phenomena in benzoic/salicylic acids (Hidaka et al., 2006 [29]). The sample was placed in a Suprasil® quartz cuvette and the emission signal was collected at 90° to the excitation beam.

3.4. Photocatalytic Studies

The photocatalytic activity was evaluated by the degradation of benzoic acid (BA) and 4-methoxycinnamic acid (MCA); their main characteristics are shown in Table 3. The tests were performed under UV (LED lamp, 365 nm) and visible light irradiation (LED lamp, 450 nm Penn Photoreactor M2 and Multiray Daylight/fluorescent 380–800 nm). In a typical test, 2.4 mg of photocatalyst powder was added to 20 mL of a 1.1 × 10−4 M (13.4 mg L−1) aqueous solution of BA. Before turning on the light, the mixture was stirred in the dark for 20 min in a 40 mL vial in order to allow adsorption/desorption equilibrium on the catalyst surface. At different time intervals (0 ÷ 5 h), aliquots of the reacting suspension (2.6 mL) were collected. The samples were centrifugated for 10 min at 450 rpm and the supernatants were analyzed with a UV-Vis spectrophotometry and high-pressure liquid chromatography, HPLC (Agilent 1200 series, CA, USA, column Zorbax Eclipse XDB-C18 4.6 × 150 mm, λ 228 nm, eluent water/methanol 6:4 + 1.5% acetic acid). The catalytic activity of the composites was evaluated with and without hydrogen peroxide addition (1 mM).
A 0.3 L glass batch reactor was employed for the photocatalytic hydrogen evolution experiments, equipped with a medium-pressure mercury lamp (Nominal Power P = 125 W) and a thermostat to control the temperature of the system. At a fixed temperature (25 °C ÷ 60 °C), during a typical experiment, a selected amount of photocatalyst (500 ppm) was suspended in a methanol aqueous solution (2.5 M). To prevent the reaction between dissolved oxygen and the photogenerated electrons, a nitrogen (N2) stream was bubbled into the solution starting 40 min before the experiment. A Gas Chromatograph (Agilent 7820A) with an HP-PLOT Molesieve 5A column (Agilent) and a TCD detector, with argon as the carrier gas was utilized for the estimation of hydrogen production.

4. Conclusions

We have developed a facile and cost-effective approach to the preparation of a LaFeO3/TiO2 (P25) composites with LaFeO3 wt% up to 12.2%. SEM analysis of the composites evidenced that the agglomerates of LaFeO3 have an average size in the 1 to 5 μm range and the TiO2 powder is well dispersed onto the surface of each sample. This characteristic seems to have some relationship with their good performance on the photocatalytic degradation of benzoic acid. The composites display enhanced photocatalytic activity when 380–800 nm light is used for irradiation; for example, 6.2 wt% and 12.2 wt%-LaFeO3/TiO2 composites degrade 93% of BA in 300 min compared to 80% for TiO2. Fluorescence measurements corroborated that salicylic acid is one of the intermediates formed during the degradation of BA by OH• radical attack. Under 450 nm LED light, all LaFeO3/TiO2 composites degraded less than 10% of BA, but in the presence of H2O2 (1 mM) the photocatalytic activity was as high as 96% in <120 min, 6.2 wt%-LaFeO3/TiO2 composite being the most efficient sample. Then, the catalysts were suspended in a methanol aqueous solution (2.5 M) and irradiated under UV-visible light. A hydrogen evolution rate of 700 μmol h−1g−1 was measured for the 3.86%wt-LaFeO3/TiO2 nanocomposite and good recyclability was established. The photocatalytic activity of LaFeO3/TiO2 composites is characterized by a volcano-like profile for both the BA degradation and the hydrogen evolution, but the optimum semiconductors ratio (LaFeO3:TiO2) depends on the wavelength of activating light. The mechanism of the photocatalytic degradation is proposed to involve the photoactivation of both LaFeO3 and TiO2 phases, leading to the formation of photogenerated charges ( e - h + ) in both semiconductors. Further, the improvement of the photoactivity of the composites can be attributed to their larger surface area compared to pure LaFeO3 and good dispersion of TiO2 over the surface of LaFeO3, which might have decreased the recombination center.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules30071526/s1, Figure S1: SEM images of LaFeO3 powder. Figure S2: SEM image and EDX spectrum of C sample. Figure S3: Plot of Kubelka-Munk function (F(R∞)*E)2 vs. photon energy obtained from DRS for band gap calculation of LaFeO3. Figure S4: Photocatalytic degradation of MCA up to 120 min of irradiation from UV-vis data. Experimental conditions: 0.120 g L−1 catalyst loading, [MCA] = 2.5 × 10−5 M under 380–800 nm light. For reference, the MCA photolysis is reported (red stars).

Author Contributions

Conceptualization, I.N.S., R.P. and M.M.; methodology, I.N.S., B.B., R.P., A.L., M.T. and M.M.; validation I.N.S., B.B., R.P., A.L., M.T. and M.M.; investigation, I.N.S., B.B., R.P., A.L., M.T. and M.M.; resources, I.N.S. and M.M.; data curation, R.P.; writing—original draft preparation, I.N.S. and M.M.; writing—review and editing, I.N.S., M.T. and M.M.; visualization, R.P.; supervision, I.N.S. and M.M. 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

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kumari, H.; Ranga, R.; Chahal, S.; Devi, S.; Sharma, S.; Kumar, S.; Kumar, P.; Kumar, S.; Kumar, A.; Parmar, R. A Review on Photocatalysis Used For Wastewater Treatment: Dye Degradation. Water Air Soil Pollut. 2023, 234, 349. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, H.; Li, X.; Zhao, X.; Li, C.; Song, X.; Zhang, P.; Huo, P. A Review on Heterogeneous Photocatalysis for Environmental Remediation: From Semiconductors to Modification Strategies. Chin. J. Catal. 2022, 43, 178–214. [Google Scholar] [CrossRef]
  3. Karthikeyan, C.; Arunachalam, P.; Ramachandran, K.; Al-Mayouf, A.M.; Karuppuchamy, S. Recent Advances in Semiconductor Metal Oxides with Enhanced Methods for Solar Photocatalytic Applications. J. Alloys Compd. 2020, 828, 154281. [Google Scholar] [CrossRef]
  4. Low, J.; Yu, J.; Jaroniec, M.; Wageh, S.; Al-Ghamdi, A.A. Heterojunction Photocatalysts. Adv. Mater. 2017, 29, 1601694. [Google Scholar] [CrossRef]
  5. Gómez, L.D.; Rodríguez-Páez, J.E. Micro/Nanoscale Mesoporous Nb2O5 Particles: Effect of Synthesis Conditions and Doping with N, C, or S on Their Properties. Nano-Struct. Nano-Objects 2019, 17, 43–57. [Google Scholar] [CrossRef]
  6. Hoffmann, M.R.; Martin, S.T.; Choi, W.; Bahnemann, D.W. Environmental Applications of Semiconductor Photocatalysis. Chem. Rev. 1995, 95, 69–96. [Google Scholar] [CrossRef]
  7. Pelaez, M.; Nolan, N.T.; Pillai, S.C.; Seery, M.K.; Falaras, P.; Kontos, A.G.; Dunlop, P.S.M.; Hamilton, J.W.J.; Byrne, J.A.; O’Shea, K.; et al. A Review on the Visible Light Active Titanium Dioxide Photocatalysts for Environmental Applications. Appl. Catal. B Environ. 2012, 125, 331–349. [Google Scholar] [CrossRef]
  8. Ni, M.; Leung, M.K.H.; Leung, D.Y.C.; Sumathy, K. A Review and Recent Developments in Photocatalytic Water-Splitting Using TiO2 for Hydrogen Production. Renew. Sustain. Energy Rev. 2007, 11, 401–425. [Google Scholar] [CrossRef]
  9. Henderson, M.A. A Surface Science Perspective on TiO2 Photocatalysis. Surf. Sci. Rep. 2011, 66, 185–297. [Google Scholar] [CrossRef]
  10. Sora, I.N.; Fontana, F.; Passalacqua, R.; Ampelli, C.; Perathoner, S.; Centi, G.; Parrino, F.; Palmisano, L. Photoelectrochemical Properties of Doped Lanthanum Orthoferrites. Electrochim. Acta 2013, 109, 710–715. [Google Scholar] [CrossRef]
  11. Bolognino, I.; Pelosato, R.; Marcì, G.; Natali Sora, I. Comparison of Ten Metal-Doped LaFeO3 Samples on Photocatalytic Degradation of Antibiotics in Water under Visible Light: Role of Surface Area and Aqueous Phosphate Ions. Molecules 2023, 28, 3807. [Google Scholar] [CrossRef]
  12. Pelosato, R.; Carrara, V.; Sora, I.N. Enhanced Photocatalytic Degradation of Ibuprofen in Aqueous Solution under Visible-Light Irradiation: Effects of LaFeO3 and Cu/LaFeO3. Chem. Eng. Trans. 2019, 73, 181–186. [Google Scholar] [CrossRef]
  13. Turkten, N.; Sora, I.N.; Tomruk, A.; Bekbolet, M. Photocatalytic Degradation of Humic Acids Using LaFeO3. Catalysts 2018, 8, 630. [Google Scholar] [CrossRef]
  14. Dhinesh Kumar, R.; Thangappan, R.; Jayavel, R. Synthesis and Characterization of LaFeO3/TiO2 Nanocomposites for Visible Light Photocatalytic Activity. J. Phys. Chem. Solids 2017, 101, 25–33. [Google Scholar] [CrossRef]
  15. Jing, L.; Qu, Y.; Su, H.; Yao, C.; Fu, H. Synthesis of High-Activity TiO2-Based Photocatalysts by Compounding a Small Amount of Porous Nanosized LaFeO3 and the Activity-Enhanced Mechanisms. J. Phys. Chem. C 2011, 115, 12375–12380. [Google Scholar] [CrossRef]
  16. Humayun, M.; Li, Z.; Sun, L.; Zhang, X.; Raziq, F.; Zada, A.; Qu, Y.; Jing, L. Coupling of Nanocrystalline Anatase TiO2 to Porous Nanosized LaFeO3 for Efficient Visible-Light Photocatalytic Degradation of Pollutants. Nanomaterials 2016, 6, 22. [Google Scholar] [CrossRef]
  17. Tashkandi, N.Y.; Albukhari, S.M.; Ismail, A.A. Visible-Light Driven of Heterostructured LaFeO3/TiO2 Photocatalysts for Degradation of Antibiotics: Ciprofloxacin as Case Study. J. Photochem. Photobiol. A Chem. 2022, 432, 114078. [Google Scholar] [CrossRef]
  18. Hao, P.; Shi, R.; Wang, X.; Zhang, J.; Li, B.; Wang, J.; Liu, B.; Liu, Y.; Qiao, X.; Wang, Z. Efficient Tetracycline Degradation Using Carbon Quantum Dot Modified TiO2@LaFeO3 Hollow Core Shell Photocatalysts. Sci. Rep. 2024, 14, 27057. [Google Scholar] [CrossRef]
  19. Garcia-Muñoz, P.; Fresno, F.; Ivanez, J.; Robert, D.; Keller, N. Activity Enhancement Pathways in LaFeO3@TiO2 Heterojunction Photocatalysts for Visible and Solar Light Driven Degradation of Myclobutanil Pesticide in Water. J. Hazard. Mater. 2020, 400, 123099. [Google Scholar] [CrossRef]
  20. Pandey, V.; Bansal, A.; Toor, A.P. Synthesis and Performance Evaluation of S-Scheme Heterostructured LaFeO3/TiO2 Photocatalyst for the Efficient Degradation of Thiamethoxam. Environ. Sci. Pollut. Res. 2024, 31, 28578–28593. [Google Scholar] [CrossRef]
  21. Dai, J.; Zhu, Y.; Tahini, H.A.; Lin, Q.; Chen, Y.; Guan, D.; Zhou, C.; Hu, Z.; Lin, H.-J.; Chan, T.-S.; et al. Single-Phase Perovskite Oxide with Super-Exchange Induced Atomic-Scale Synergistic Active Centers Enables Ultrafast Hydrogen Evolution. Nat. Commun. 2020, 11, 5657. [Google Scholar] [CrossRef] [PubMed]
  22. Yu, J.; Li, Z.; Liu, T.; Zhao, S.; Guan, D.; Chen, D.; Shao, Z.; Ni, M. Morphology Control and Electronic Tailoring of CoxAy (A = P, S, Se) Electrocatalysts for Water Splitting. Chem. Eng. J. 2023, 460, 141674. [Google Scholar] [CrossRef]
  23. Lv, T.; Wang, H.; Hong, W.; Wang, P.; Jia, L. In Situ Self-Assembly Synthesis of Sandwich-like TiO2/Reduced Graphene Oxide/LaFeO3 Z-Scheme Ternary Heterostructure towards Enhanced Photocatalytic Hydrogen Production. Mol. Catal. 2019, 475, 110497. [Google Scholar] [CrossRef]
  24. Jiang, Y.; Lv, Q.; Xu, F.; Sun, X.; Ding, Y. Synthesis of TiO2/LaFeO3 Composites for the Photoelectrochemical Hydrogen Evolution. J. Mater. Sci. 2021, 56, 15188–15204. [Google Scholar] [CrossRef]
  25. Li, K.; Wang, D.; Wu, F.; Xie, T.; Li, T. Surface Electronic States and Photovoltage Gas-Sensitive Characters of Nanocrystalline LaFeO3. Mater. Chem. Phys. 2000, 64, 269–272. [Google Scholar] [CrossRef]
  26. Kosmulski, M. pH-Dependent Surface Charging and Points of Zero Charge II. Update. J. Colloid Interface Sci. 2004, 275, 214–224. [Google Scholar] [CrossRef] [PubMed]
  27. Mrowetz, M.; Selli, E. Photocatalytic Degradation of Formic and Benzoic Acids and Hydrogen Peroxide Evolution in TiO2 and ZnO Water Suspensions. J. Photochem. Photobiol. A Chem. 2006, 180, 15–22. [Google Scholar] [CrossRef]
  28. Ajmera, A.A.; Sawant, S.B.; Pangarkar, V.G.; Beenackers, A.A.C.M. Solar-Assisted Photocatalytic Degradation of Benzoic Acid Using Titanium Dioxide as a Photocatalyst. Chem. Eng. Technol. 2002, 25, 173–180. [Google Scholar] [CrossRef]
  29. Hidaka, H.; Honjo, H.; Horikoshi, S.; Serpone, N. Photocatalyzed Degradation on a TiO2-Coated Quartz Crystal Microbalance. Adsorption/Desorption Processes in Real Time in the Degradation of Benzoic Acid and Salicylic Acid. Catal. Commun. 2006, 7, 331–335. [Google Scholar] [CrossRef]
  30. Smith, G.J.; Miller, I.J. The Effect of Molecular Environment on the Photochemistry of P-Methoxycinnamic Acid and its Esters. J. Photochem. Photobiol. A Chem. 1998, 118, 93–97. [Google Scholar] [CrossRef]
  31. Bouleghlimat, E.; Bethell, D.; Davies, P.R. The Photocatalytic Destruction of Cinnamic Acid and Cinnamyl Alcohol: Mechanism and the Effect of Aqueous Ions. Chemosphere 2020, 251, 126469. [Google Scholar] [CrossRef] [PubMed]
  32. Barbero, N.; Vione, D. Why Dyes Should Not Be Used to Test the Photocatalytic Activity of Semiconductor Oxides. Environ. Sci. Technol. 2016, 50, 2130–2131. [Google Scholar] [CrossRef] [PubMed]
  33. Grätzel, C.K.; Jirousek, M.; Grätzel, M. Decomposition of Organophosphorus Compounds on Photoactivated TiO2 Surfaces. J. Mol. Catal. 1990, 60, 375–387. [Google Scholar] [CrossRef]
  34. Muscetta, M.; Clarizia, L.; Race, M.; Andreozzi, R.; Marotta, R.; Di Somma, I. Visible—Light Driven Systems: Effect of the Parameters Affecting Hydrogen Production through Photoreforming of Organics in Presence of Cu2O/TiO2 Nanocomposite Photocatalyst. Appl. Sci. 2023, 13, 2337. [Google Scholar] [CrossRef]
  35. Butler, M.A.; Ginley, D.S. Prediction of Flatband Potentials at Semiconductor-Electrolyte Interfaces from Atomic Electronegativities. J. Electrochem. Soc. 1978, 125, 228–232. [Google Scholar] [CrossRef]
  36. Muscetta, M.; Jitan, S.A.; Palmisano, G.; Andreozzi, R.; Marotta, R.; Cimino, S.; Di Somma, I. Visible LightDriven Photocatalytic Hydrogen Production Using Cu2O/TiO2 composites Prepared by Facile Mechanochemical Synthesis. J. Environ. Chem. Eng. 2022, 10, 107735. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of B-D-E nanocomposites and the reference compounds TiO2 and LaFeO3; (b) SEM image of C; (c) SEM image of D; (d) SEM image of E; (e) SEM image of LaFeO3.
Figure 1. (a) XRD patterns of B-D-E nanocomposites and the reference compounds TiO2 and LaFeO3; (b) SEM image of C; (c) SEM image of D; (d) SEM image of E; (e) SEM image of LaFeO3.
Molecules 30 01526 g001
Figure 2. Photocatalytic degradation of BA in the presence of samples A (TiO2), B (1.3 wt%-LaFeO3/TiO2), C (3.9 wt%-LaFeO3/TiO2), D (6.2 wt%-LaFeO3/TiO2), E (12.2 wt%-LaFeO3/TiO2) and LF (LaFeO3) from HPLC analysis. Experimental conditions: 0.120 g L−1 catalyst loading, [BA] = 1 × 10−4 M, λ = 380–800 nm light.
Figure 2. Photocatalytic degradation of BA in the presence of samples A (TiO2), B (1.3 wt%-LaFeO3/TiO2), C (3.9 wt%-LaFeO3/TiO2), D (6.2 wt%-LaFeO3/TiO2), E (12.2 wt%-LaFeO3/TiO2) and LF (LaFeO3) from HPLC analysis. Experimental conditions: 0.120 g L−1 catalyst loading, [BA] = 1 × 10−4 M, λ = 380–800 nm light.
Molecules 30 01526 g002
Figure 3. Pathway of degradation of benzoic acid following ref. [28].
Figure 3. Pathway of degradation of benzoic acid following ref. [28].
Molecules 30 01526 g003
Figure 4. Fluorescence spectral patterns against UV irradiation time (t1 = 1 h) in the decomposition of benzoic acid in the presence of (a) sample A (TiO2), (b) sample E, and (c) sample F (LaFeO3). The excitation wavelength was 230 nm. Experimental conditions: 0.120 g L−1 catalyst loading, [BA] = 1 × 10−4 M, λ = 380–800 nm radiation. The measured solutions were obtained from the supernatant after centrifugation of the corresponding suspensions containing the photocatalyst. The peaks at ca. 460 nm and 690 nm are instrumental background, as evidenced by the spectrum of pure water excited in the same conditions, which is reported in each panel (….. line) to guide the eye. The expected location of the signals from benzoic and salicylic acids is indicated by yellow lines.
Figure 4. Fluorescence spectral patterns against UV irradiation time (t1 = 1 h) in the decomposition of benzoic acid in the presence of (a) sample A (TiO2), (b) sample E, and (c) sample F (LaFeO3). The excitation wavelength was 230 nm. Experimental conditions: 0.120 g L−1 catalyst loading, [BA] = 1 × 10−4 M, λ = 380–800 nm radiation. The measured solutions were obtained from the supernatant after centrifugation of the corresponding suspensions containing the photocatalyst. The peaks at ca. 460 nm and 690 nm are instrumental background, as evidenced by the spectrum of pure water excited in the same conditions, which is reported in each panel (….. line) to guide the eye. The expected location of the signals from benzoic and salicylic acids is indicated by yellow lines.
Molecules 30 01526 g004
Figure 5. Photocatalytic degradation of BA in the presence of samples A (TiO2), B (1.3 wt%-LaFeO3/TiO2), C (3.9 wt%-LaFeO3/TiO2), D (6.2 wt%-LaFeO3/TiO2), E (12.2 wt%-LaFeO3/TiO2) and LF (LaFeO3), without H2O2 (a) and with H2O2 (b). Data from HPLC analysis. Experimental conditions: 0.120 g L−1 catalyst loading, [BA] = 1 × 10−4 M, [H2O2] = 1 mM, under monochromatic 450 nm light. For reference, the degradation of BA in the presence of H2O2 without catalyst is shown (blue dotted line).
Figure 5. Photocatalytic degradation of BA in the presence of samples A (TiO2), B (1.3 wt%-LaFeO3/TiO2), C (3.9 wt%-LaFeO3/TiO2), D (6.2 wt%-LaFeO3/TiO2), E (12.2 wt%-LaFeO3/TiO2) and LF (LaFeO3), without H2O2 (a) and with H2O2 (b). Data from HPLC analysis. Experimental conditions: 0.120 g L−1 catalyst loading, [BA] = 1 × 10−4 M, [H2O2] = 1 mM, under monochromatic 450 nm light. For reference, the degradation of BA in the presence of H2O2 without catalyst is shown (blue dotted line).
Molecules 30 01526 g005
Figure 6. Photocatalytic hydrogen evolution rate (HER) at varying LaFeO3 (%wt.) with respect to P25. Methanol was used as a model scavenger. Experimental conditions: [Methanol] = 2.5 M; 0.5 g L−1 catalyst loading (Sample C); V = 0.3 L; T = 35 °C; natural pH of the solution.
Figure 6. Photocatalytic hydrogen evolution rate (HER) at varying LaFeO3 (%wt.) with respect to P25. Methanol was used as a model scavenger. Experimental conditions: [Methanol] = 2.5 M; 0.5 g L−1 catalyst loading (Sample C); V = 0.3 L; T = 35 °C; natural pH of the solution.
Molecules 30 01526 g006
Figure 7. Photocatalytic hydrogen evolution rate (HER) at varying temperature of the system. Methanol was used as a model scavenger. Experimental conditions: [Methanol] = 2.5 M; 0.50 g L−1 catalyst loading (Sample C); V = 0.3 L; natural pH of the solution.
Figure 7. Photocatalytic hydrogen evolution rate (HER) at varying temperature of the system. Methanol was used as a model scavenger. Experimental conditions: [Methanol] = 2.5 M; 0.50 g L−1 catalyst loading (Sample C); V = 0.3 L; natural pH of the solution.
Molecules 30 01526 g007
Figure 8. (A) Photocatalytic hydrogen evolution rate (HER) in successive photocatalytic tests. (B) Photocatalytic hydrogen evolution rate (HER) during a single photostability test. Methanol was used as a model scavenger. Experimental conditions: [Methanol] = 2.5 M; 0.50 g L−1 catalyst loading (Sample C); V = 0.3 L; natural pH of the solution; T = 35 °C.
Figure 8. (A) Photocatalytic hydrogen evolution rate (HER) in successive photocatalytic tests. (B) Photocatalytic hydrogen evolution rate (HER) during a single photostability test. Methanol was used as a model scavenger. Experimental conditions: [Methanol] = 2.5 M; 0.50 g L−1 catalyst loading (Sample C); V = 0.3 L; natural pH of the solution; T = 35 °C.
Molecules 30 01526 g008
Figure 9. Scheme of the energy band structures and the correspondingly transfer process of photogenerated electron–hole pairs of semiconductor photocatalysts under UV light.
Figure 9. Scheme of the energy band structures and the correspondingly transfer process of photogenerated electron–hole pairs of semiconductor photocatalysts under UV light.
Molecules 30 01526 g009
Table 1. Pollutant photodegradation over LaFeO3/TiO2 composites.
Table 1. Pollutant photodegradation over LaFeO3/TiO2 composites.
PhotocatalystLaFeO3:TiO2RadiationOrganic CompoundCatalystRemoval
(%)
Time
(min)
Ref.
LaFeO3/TiO2 rutile
calcination 800 °C/6 h
1.5:1
(m/m)
visiblemethyl orangen.a.90180[14]
LaFeO3@TiO2 heterojunction5:100300–800 nmmyclobutanil
20 mg L−1
1 g L−1100180[19]
LaFeO3@TiO2 heterojunction12.5:100>420 nmmyclobutanil
20 mg L−1
1 g L−137240[19]
LaFeO3/TiO2 anatase
calcination 500 °C/3 h
3:100>420 nmciprofloxacin
10 mg L−1
2.4 g L−110090[17]
LaFeO3/TiO2 (P25)
calcination 450 °C/2 h
0.85:100254 nmthiamethozam
20 mg L−1
0.8 g L−197120[20]
LaFeO3/TiO2 (P25)
calcination 450 °C/2 h
0.85:100direct sunlightthiamethozam
20 mg L−1
0.8 g L-179120[20]
LaFeO3/TiO2 (P25)
ball-milling
6.2:100LED450 nmbenzoic acid
13.4 mg L−1
0.12 g L−1
H2O2 1 mM
96120this work
LaFeO3/TiO2 (P25)
ball-milling
6.2:100LED
450 nm
benzoic acid
13.4 mg L−1
0.12 g L−19120this work
LaFeO3/TiO2 (P25)
ball-milling
6.2:100380–800 nmbenzoic acid
13.4 mg L−1
0.12 g L−190300this work
Table 2. Values of the rate constant (k) for benzoic acid, and correlation coefficients of the linear fittings of the experimental data according to the pseudo-first-order reaction kinetic. Experimental conditions: 0.120 g L−1 catalyst loading, [BA] = 1 × 10−4 M, [H2O2] = 1 mM, monochromatic 450 nm light.
Table 2. Values of the rate constant (k) for benzoic acid, and correlation coefficients of the linear fittings of the experimental data according to the pseudo-first-order reaction kinetic. Experimental conditions: 0.120 g L−1 catalyst loading, [BA] = 1 × 10−4 M, [H2O2] = 1 mM, monochromatic 450 nm light.
Catalystk (s−1)R2
A4.1 × 10−40.934
B4.4 × 10−40.969
C3.2 × 10−40.996
D4.8 × 10−40.943
E4.0 × 10−40.995
F1.5 × 10−40.967
Table 3. Main physical–chemical properties of benzoic acid and 4-methoxycinnamic acid.
Table 3. Main physical–chemical properties of benzoic acid and 4-methoxycinnamic acid.
Benzoic Acid4-Methoxycinnamic Acid
N. CAS 65-85-0943-89-5
Chemical structureMolecules 30 01526 i001Molecules 30 01526 i002
Molecular formulaC6H5COOHC10H10O3
Molecular mass 122.12 g/mol178.18 g/mol
Physical formWhite powderWhite powder
Solubility in water3.4 g/L at 25 °C0.71 g/L at 25 °C
pKa4.194.60
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Natali Sora, I.; Bertolotti, B.; Pelosato, R.; Lucotti, A.; Tommasini, M.; Muscetta, M. TiO2/LaFeO3 Composites for the Efficient Degradation of Benzoic Acid and Hydrogen Production. Molecules 2025, 30, 1526. https://doi.org/10.3390/molecules30071526

AMA Style

Natali Sora I, Bertolotti B, Pelosato R, Lucotti A, Tommasini M, Muscetta M. TiO2/LaFeO3 Composites for the Efficient Degradation of Benzoic Acid and Hydrogen Production. Molecules. 2025; 30(7):1526. https://doi.org/10.3390/molecules30071526

Chicago/Turabian Style

Natali Sora, Isabella, Benedetta Bertolotti, Renato Pelosato, Andrea Lucotti, Matteo Tommasini, and Marica Muscetta. 2025. "TiO2/LaFeO3 Composites for the Efficient Degradation of Benzoic Acid and Hydrogen Production" Molecules 30, no. 7: 1526. https://doi.org/10.3390/molecules30071526

APA Style

Natali Sora, I., Bertolotti, B., Pelosato, R., Lucotti, A., Tommasini, M., & Muscetta, M. (2025). TiO2/LaFeO3 Composites for the Efficient Degradation of Benzoic Acid and Hydrogen Production. Molecules, 30(7), 1526. https://doi.org/10.3390/molecules30071526

Article Metrics

Back to TopTop