Next Article in Journal
Sensitivity Estimation for Differentially Private Query Processing
Previous Article in Journal
Towards Understanding Driver Acceptance of C-ITS Services—A Multi-Use Case Field Study Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Layered Perovskite La2Ti2O7 Obtained by Sol–Gel Method with Photocatalytic Activity

1
“Ilie Murgulescu” Institute of Physical Chemistry of the Romanian Academy, 202 Splaiul Independenței, 060021 Bucharest, Romania
2
National Institute for Research & Development in Chemistry and Petrochemistry ICECHIM, 202 Splaiul Independenței, 060021 Bucharest, Romania
3
National Institute of Materials Physics, Atomistilor 405A, 077125 Magurele, Romania
4
Chemical Sciences Department, Romanian Academy, 125 Victoriei Avenue, 1st District, 010071 Bucharest, Romania
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2025, 15(14), 7665; https://doi.org/10.3390/app15147665
Submission received: 28 May 2025 / Revised: 4 July 2025 / Accepted: 7 July 2025 / Published: 8 July 2025
(This article belongs to the Special Issue Application of Nanomaterials in the Field of Photocatalysis)

Abstract

This paper presents the synthesis of La2Ti2O7 nanoparticles by the sol–gel method starting from lanthanum nitrate and titanium alkoxide (noted as LTA). Subsequently, the lanthanum titanium oxide nanoparticles are modified with noble metals (platinum) using the chemical impregnation method, followed by a reduction process with NaBH4. The comparative analysis of the structure and surface characteristics of the nanopowders subjected to thermal treatment at 900 °C is conducted using Fourier transform infrared spectroscopy (FT-IR), Raman spectroscopy, X-ray photoelectron spectroscopy (XPS), scanning electron microscopy (SEM) with energy-dispersive X-ray spectroscopy (EDX), transmission electron microscopy (TEM), X-ray diffraction (XRD), X-ray fluorescence (XRF), ultraviolet-visible (UV–Vis) spectroscopy, as well as specific surface area and porosity measurements. The photocatalytic activity is evaluated in the oxidative photodegradation of ethanol (CH3CH2OH) under simulated solar irradiation. The modified sample shows higher specific surfaces areas and improved photocatalytic properties, proving the better conversion of CH3CH2OH than the pure sample. The highest conversion of ethanol (29.75%) is obtained in the case of LTA-Pt after 3 h of simulated solar light irradiation.

1. Introduction

Over decades of investigation, the production of innovative photocatalysts has been an important subject in the scientific community for their various applications, including the degradation of organic pollutants [1,2,3], mineralization of toxic organic compounds [4], degradation of polymer fibers [5], degradation of pharmaceutical waste [6], industrial wastewater treatment [7] and treatment of agricultural wastewater [8], or the splitting water to generate hydrogen [9,10].
La2Ti2O7 belongs to the family of layered perovskite-type materials with the general formula AnTinO3n+2 [11]. For n = 4, the layered perovskite phase is stable under ambient conditions when the cationic radius ratio is greater than or equal to 1.78; in this case, the La3+/Ti4+ ratio is 1.92 [12,13].
The La2Ti2O7 phase crystallizes in the monoclinic system (space group P21) and contains four formula units per unit cell. Its structure is composed of four distorted perovskite units aligned perpendicularly on infinite layers of TiO6 octahedra [14]. This structural anisotropy results in anisotropic dielectric, electrical, and mechanical properties [12].
La2Ti2O7 is a multifunctional oxide recognized for several notable attributes:
  • Piezoelectric/ferroelectric properties: exhibits a high Curie temperature [15,16].
  • Superior thermal stability: possesses good electron mobility due to ions on A sites and high conductivity [16].
  • Wide bandgap (~3.8 eV): capable of generating photoelectrons with a high reduction potential [17,18].
  • Good resistance to ionizing radiation [19].
  • Layered structure: favorable for the separation and transfer of photogenerated electron–hole pairs, as well as separating oxidation from reduction reaction sites [18].
  • Host matrix: can host other lanthanides to create new phosphors with conventional luminescence under UV excitation [15].
  • Photocatalytic capabilities: particularly for hydrogen generation via water splitting [20,21] and for the degradation of volatile organic compounds (VOCs) [22].
  • However, its large band gap limits photocatalytic efficiency under visible light [16], and its low solar light utilization and easy recombination of photogenerated charge carriers hinder achieving the commercial application target of 5–10% solar hydrogen conversion efficiency (STH) [9].
La2Ti2O7 is a material with increasing applications in a variety of domains, from catalysis and photocatalysis [23,24,25] to dielectrics [26], superconductors [27], and luminescent materials [19]. Numerous environmental issues could be resolved by creating stable, inexpensive, and highly efficient photocatalysts that utilize solar energy [28]. Applications of La2Ti2O7 oxide include water-splitting [29]; co-catalyzation with metals [30,31]; the photocatalytic degradation of organic pollutants and colorants [13,15,16,32,33]; the oxidation of methanol [34,35], ethanol [36,37], and phenol [14,38]; and the photodegradation of antibiotics [21,39].
However, the degree of crystallinity and the material’s exchange surface have a significant impact on the photocatalytic efficiency. In recent years, scientific research has seen a great deal of interest regarding the optimization of synthetic parameters to control the size of crystallites and their crystallinity, specifically through the following: (i) the use of polymer precursors for controlling porosity [40], (ii) the type of precursors used [41,42], and (iii) the diversification of synthesis processes. Various techniques have been used to synthesize La2Ti2O7, including solid-state processing [43], co-precipitation of hydroxides [44], urea precipitation [45], hydrothermal synthesis [46], high-temperature decomposition of metal–organic precursors [47], thermal decomposition of nitrates [48], liquid mix techniques [49], the hydrothermal route [50,51], microwave-assisted synthesis [52], and the sol–gel route [53,54,55]. While these methods offer the advantage of smaller particle sizes over the solid-state approach, most of them typically come with the drawbacks of extended heating times and multi-step synthetic procedures.
The sol–gel process has attracted attention as an interesting method for creating nanocomposites [56,57]. Better control of stoichiometry and homogeneity, lower reaction temperatures, reduced contamination, and being an easy way of fabricating ultrafine powders, thin films, and fibers are some of the benefits of the sol–gel method over many techniques. The hydrolysis and subsequent polycondensation reactions of the component alkoxides represent a major part of the sol–gel synthesis [58,59]. In contrast to conventional solid-state reactions, which typically require calcination temperatures above 1000 °C and often produce inhomogeneous, aggregated particles with low surface areas, the sol–gel method facilitates the mixing of precursors at the molecular level. This allows for better control over stoichiometry, particle size, and morphology. Compared to hydrothermal synthesis, which involves prolonged reaction times under high-pressure autoclave conditions that are difficult to scale up [51], the sol–gel method offers a more scalable and reproducible approach. Its ability to operate at lower synthesis temperatures and to fine-tune textural properties makes it particularly attractive for tailoring the photocatalytic activity of La2Ti2O7.
The aim of the present work is to study the synthesis of lanthanum and titanium perovskite through the sol–gel method and their photocatalytic activity for ethanol oxidative degradation under simulated solar light. The photocatalytic activity enhancement by impregnating La2Ti2O7 with platinum (Pt), the evolution of reaction products in time, and the relationship between catalytic performances, morphology, and the structure of catalysts are also presented.

2. Materials and Methods

2.1. Gel and Powder Preparation

Nanopowders of La2Ti2O7 were prepared by the sol–gel method (see Figure 1), while the La2Ti2O7 modified with platinum was obtained by post-synthesis impregnation.
In the synthesis, the reaction started from lanthanum (III) nitrate hexahydrated [La(NO3)3 × 6H2O] (Merck, Darmstadt, Germany) as the La precursor and titanium (IV) isopropoxide [Ti[OCH(CH3)2]4] (Merck, Hohenbrunn, Germany) as the precursor of Ti. The molar ratio of La/Ti was 1:1. The lanthanum precursor was dissolved in glacial acetic acid (CH3COOH) (Sigma-Aldrich, Steinheim, Germany) in a first solution, while the titanium precursor was mixed with isopropanol (i-C3H7OH) (Merck, Hohenbrunn, Germany) in a second solution, both with a mass ratio of 1:5. The two freshly prepared solutions were mixed and stirred for 30 min. The resulting translucent solution was placed in a water bath at 60 °C for 60 min until a whitish powder was formed. This powder was dried at 80 °C (~3 h), then thermally treated at 900 °C for 2 h to obtain the white La2Ti2O7 nanopowder, noted as LTA.
The choice of the La(NO3)3 × 6H2O and Ti[OCH(CH3)2]4 precursors for obtaining high-quality La2Ti2O7 by the sol–gel method allows for good control over the stoichiometry and leads to a homogeneous distribution of La and Ti ions at the molecular level. Lanthanum nitrate provides La3+ ions without introducing undesired anions (such as Cl, which could remain as impurities or create by-products). Titanium alkoxides are the preferred titanium precursors in sol–gel chemistry due to their tunable reactivity. The alkoxide groups readily undergo hydrolysis and condensation, forming the Ti–O–Ti and La–O–Ti structures. Another advantage of these precursors is that they decompose cleanly upon calcination. This clean decomposition leads to fewer impurities and better control of the crystallinity and phase purity in the final La2Ti2O7 powder [15,54,56,60,61].
The metal-modified catalysts were prepared as follows: First, the previously obtained nanopowder was dispersed in distilled water. Then, an aqueous solution containing the dissolved 1 mol% noble metal precursor, PtCl4 (Fluka, Buchs, Switzerland) was slowly added to La2Ti2O7 suspended previously in water and left under continuous stirring at room temperature for 24 h. The metal reduction process was completed by slowly adding 5 mL of aqueous solution with 0.04 g of freshly dissolved NaBH4 (TCI, Tokyo, Japan) to the La2Ti2O7 suspension. The solid material was recovered by filtration, washed with distilled water, and dried at 80 °C for 5 h [38].

2.2. Methods

The study and characterization of La2Ti2O7-based oxide nanopowders obtained by the sol–gel method was approached according to the main factors influencing their structure and morphology.
The following methods of characterization were used: infrared spectroscopy (FT-IR), Raman spectroscopy, X-ray photoelectron spectroscopy (XPS), scanning electron microscopy (SEM) coupled with Energy Dispersive X-ray spectroscopy (EDX), transmission electron microscopy (TEM), X-ray diffraction (XRD), the determination of the BET-specific surface area and the pore distribution, and UV-Vis spectroscopy.
FT-IR spectra of the thermally treated powders were recorded over the range of 4000–400 cm−1 on a Jasco FT-IR-4700 spectrophotometer (Jasco, Tokyo, Japan) using the KBr disk method.
The Renishaw inVia Qontor Spectrometer System (Renishaw PLC, Wotton-under-Edge, Gloucestershire, UK) for confocal Raman spectral imaging using the visible laser wavelength (532 nm) was used to examine the synthesized samples. The acquisition was performed in the spectral range 100 to 1000 cm−1, using an exposure time of 10 s, a power lower than 10% (5 mW) with different number of accumulations (usually 5 accumulations), and a grating of 1800 l/mm. The microscope objective was a Leica N PLAN X100 magnification (Leica Camera AG, Wetzlar, Germany) with a focal distance of 0.33 mm and numerical aperture of 0.85. The detector was a Master Renishaw Centrus (Renishaw, Wotton-under-Edge, UK).
XPS measurements were performed using an XPS spectrometer (SPECS Surface Nano Analysis GmbH, Berlin, Germany) equipped with a PHOIBOS 150 mm hemispherical electron energy analyzer, a multichanneltron detector, and a monochromated Al Kα radiation (1486.7 eV) anode as the excitation source. The operating power was 250 W (12.5 kV × 20 mA) and the spectra were recorded with 99 eV pass energy for general surveys and 30 eV pass energy for high-resolution, core-level measurements, with partial charge compensation provided by a neutralizer flood gun (1 eV × 100 μA). The base pressure in the analysis chamber before measurements was 10−10 mbar.
The particle morphology and elemental composition of the samples were examined using a Verios G4 UC scanning electron microscope (Thermo Scientific, Brno, Czech Republic) equipped with an Energy Dispersive X-ray (EDX) spectroscopy analyzer (Octane Elect Super SDD detector, Octane, New York, NY, USA). For SEM analysis, the specimens were mounted on aluminum stubs using double-sided carbon adhesive tape and subsequently coated with a 6 nm platinum layer via a Leica EM ACE200 sputter coater (Leica, Vienna, Austria) to enhance electrical conductivity and mitigate charging effects under electron beam irradiation. Morphological characterization was performed utilizing a secondary electron detector (TLD—Through the Lens Detector) to accurately delineate particle shape and size. SEM micrographs were acquired at an accelerating voltage of 10 kV and a beam current of 0.4 nA. For EDX elemental analysis, the samples were examined at an accelerating voltage of 20 kV and a beam current of 6.4 nA.
The TEM studies were performed using a transmission electron microscope Hitachi HT-7700 (Hitachi, Tokyo, Japan) at an accelerating voltage of 100 kV. For TEM analysis the nanoparticles were dispersed in ethanol, ultrasonicated, placed on carbon-coated copper grids with a 300-mesh size, and dried in an oven until the solvent was removed.
The X-ray diffraction (XRD) patterns of the powders were obtained using a Rigaku Ultima IV diffractometer (Rigaku Corp., Tokyo, Japan). The instrument was configured in a parallel beam geometry with Cu Kα radiation (λ = 1.5406 Å) and CBO optics, and operated at 40 kV and 30 mA. Data were collected with a step size of 0.02° and a scan speed of 2° min−1. Phase identification was carried out with a Search–Match algorithm, connected to the PDF database.
Elemental analysis was conducted via X-ray fluorescence (XRF) using a Rigaku ZSX Primus II spectrometer (Rigaku Corp., Tokyo, Japan) equipped with a 4.0 kW X-ray Rh tube. Data processing was performed using EZscan combined with Rigaku SQX fundamental parameters software v.5.18, operating in a standardless mode.
The textural analysis of the powders was performed through nitrogen adsorption–desorption analysis at −196 °C using a ASAP 2020 automated gas adsorption analyzer (Micromeritics, Norcross, GA, USA). Before analysis, the samples were degassed under vacuum at 200 °C for 5 h. The specific surface area was determined using the Brunauer–Emmett–Teller (BET) equation, while pore size distribution curves were generated from the adsorption branch based on the Barrett–Joyner–Halenda (BJH) model. The total pore volume was calculated from the amount adsorbed at a relative pressure of 0.99.
The UV–VIS spectra of the samples were recorded using a Perkin Elmer Lambda 35 spectrophotometer (PerkinElmer, Norwalk, CT, USA), equipped with an integrating sphere, in the wavelength range of 200–1100 nm [ultraviolet (200–400 nm), visible (400–780 nm), and near-infrared (780–2500 nm)], using a white reflective standard Spectralon.
ROS monitoring was performed according to previously reported works [62,63]:
-
·O2 photogeneration. UV–Vis spectra of aqueous mixtures containing 50 μM XTT sodium salt 9([2, 3-bis(2-methoxi-4-nitro-5sulfophenyl)-2H-tetrazolium-5-carboxanilide]) (AlfaAesar, Karlsruhe, Germany) and LTA, LTA-Pt powders (2 mg) were measured with an Specord 200 Plus Spectrophotometer (Analytik Jena, Jena, Germany). The reaction was carried out under simulated solar light of AM1.5 (1000 W m2, Peccel-L 01) (Peccell, Kawasaki, Japan).
-
·OH radical trapping experiments have also been performed by PL spectroscopy using the Cary Eclipse G9800A fluorescence spectrometer (Agilent Technologies, Penang, Malaysia) with coumarin (Merck, Darmstadt, Germany), checking for the presence of umbelliferone (a fluorescent product which peaked at 470 nm) and proving that hydroxyl radicals were missing.
The photoluminescence spectroscopy measurements were performed at room temperature using a Carry Eclipse G9800A fluorescence spectrometer (Agilent Technologies, Penang, Malaysia) with excitation/emission slit widths of 10/10 nm. The powders of interest (2 mg) were suspended in 3 mL of water containing 30 μL absolute ethanol.
The catalytic activity of the prepared materials was tested for the oxidative degradation of ethanol in the gaseous phase under simulated solar light irradiation. As part of the experimental process, 0.05 g of photocatalyst powder was deposited as a uniform layer (surface area of about 3.6 cm2 and 1 mm thicknesses) on a substrate. The surface was then exposed to simulated solar light. Ethanol (7.2 μL) (Merck, Darmstadt, Germany) was introduced into the photoreactor and evaporated into a volume of 120 cm3 air and then left in the dark for 30 min to balance the experimental system. A cryostat was used to regulate and maintain the photoreactor’s internal temperature at 25 °C. A Peccell L01 solar simulator (Peccell, Kawasaki, Japan) provided the AM1.5 (1000 W/m2) solar light. A gas chromatograph (Agilent, Model 7890A, Wilmington, DE, USA) equipped with the Flame Ionization Detector (FID) was used to analyze 200 μL of gas sample every 30 min, and a gas chromatograph (Buck Scientific, East Norwalk, CT, USA) equipped with the Thermal Conductivity Detector (TCD) was used to analyze 500 μL of gas sample every 60 min. The photocatalytic activity of the investigated catalysts was monitored for 180 min.

3. Results and Discussion

The samples obtained in the conditions presented in the experimental part were yellowish white powders. Based on the TG/DTA results of the La2Ti2O7 [15], the dried powders were thermally treated at 900 °C for 2 h, obtaining white crystallized powders. The thermally treated powders were investigated for their morphology, structure, and photocatalytic properties.

3.1. Fourier Transform Infrared Spectroscopy (FT-IR)

The FT-IR spectra of the thermally treated powders are shown in Figure 2 and the corresponding vibration bands are presented comparatively in Table 1. In Figure 2 the spectrum in the case of the LTA sample is shown. The vibration band at 806 cm−1 is attributed to the stretching vibrations of the Ti–O–Ti bridges, which are characteristic of the TiO6 octahedra in the titanate structures and 776 cm−1 is assigned to the Ti–O stretching vibrations, specifically related to the TiO6 octahedra framework within La2Ti2O7 [64]. The vibration bands at 620 cm−1 and 547 cm−1 represent La–O stretching and Ti–O vibrations within the La2Ti2O7 lattice, which contribute to the stability of the titanate framework. The vibration bands at 491 cm−1 and 464 cm−1 are assigned to La–O stretching vibrations and Ti–O bending modes, which are fundamental vibrations in La2Ti2O7. The presence of these bands confirms the formation of a well-crystallized La2Ti2O7 phase. The vibration bands between 800 and 400 cm−1 are assigned to metal–oxygen (M–O) stretching vibrations, suggesting the formation of La–O and Ti–O networks [65].
Comparing the FT-IR spectra in Figure 2a,b for the Pt-modified sample, no changes are observed due to the incorporation of the noble metal into the perovskite structure.

3.2. Raman Spectroscopy

Table 2 lists the Raman shifts in both LTA-Pt and LTA, according to [9,66] and our work. According to the literature [9], the wavenumbers of the rare-earth oxygen-stretching modes (La–O) generally lie in the region of 300–500 cm−1. Some of the bands in this range are due to Ti–O vibrations or, more probably, to complex motions involving the participation of both La and Ti cations.
There are no new chemical bonds forming in the Pt-modified La2Ti2O7 samples, suggesting that the dopant (Pt) might be found in the metallic form. A newly formed chemical bond would have given a vibration in the Raman spectra of the LTA-Pt. It could be the case that there is the intercalation of the Pt between the layers of the La2Ti2O7 perovskite.
The typical crystalline structure of monoclinic layered perovskite La2Ti2O7 is observed with Raman analysis. Except the peaks at 488 and 668 cm−1 from the LTA sample, the peaks at 342, 371, 405, 430, 448, 520, 541, 559, 609, and 793 cm−1 correspond to the typical monoclinic layered perovskite crystalline structure [13,67]. The increased Raman intensity at 446 cm−1 in the LTA without Pt is more considerable when compared to the LTA-Pt. For LTA the 446 cm−1 peak has similar intensity with the 402 cm−1 peak, as in LTA-Pt, where it is less intense. As the peak appearing at 446 cm−1 is related to stretching vibrations of Ti–O–Ti in the distorted [TiO6], the small disruption in crystallinity and further distortion of the TiO6 may suggest the introduction of defects by Pt [9]. This tendency was already observed in the literature for the doping Na-doped LTA [9], as the intercalation of Pt between the LTA layers disrupts the crystallinity of the pure LTA. The appearance of the 668 cm−1 peak in varying intensities according to the probed spot on the surface of pure LTA (LTA without Pt) is the most notable difference in the Raman spectra. This additional Raman mode at 668 cm−1 marked in Figure 3 could indicate local lattice imperfections. This additional mode at 668 cm−1 is close to the assigned Ti–O–Ti vibration (661 cm−1) from two-dimensional lepidocrocite-type TiO6 layers [64]. It could be related to the vibrations of the [TiO6] distorted octahedron, as the other peaks are higher than 500 nm, and it could be assigned to the slightly increased asymmetry of the layered structure when compared to the other structural symmetries of perovskites. The Raman bands of the layered perovskite structure are complex due to this asymmetry. The crystalline facet which is exposed to the Raman laser plays an important role. If present at low concentrations (<1%), then the effect on the host lattice vibrational modes may not be perceptible.

3.3. X-Ray Photoelectron Spectroscopy (XPS)

XPS analysis was employed to determine the chemical composition and valence states of the ions in the samples. C 1s was used as the reference for the calibrations of the spectra of the other elements. The two survey spectra are shown in Figure 4. The main difference between the samples is the presence of the Pt core level and Auger electron peaks in the spectral shapes of La, Ti, and O.
In the Ti 2p spectrum of the LTA sample (Figure 5a), two components are identified. The peak at approximately 458 eV is characteristic of Ti4+ [9,26,27,40]. A second peak at a higher binding energy, around 460 eV, may be attributed either to a different TiO2 phase than that present in the La2Ti2O7 structure or to a non-stoichiometric TiOx phase [68]; however, it is unlikely to correspond to Ti3+, since Ti2O3 is typically detected at lower binding energies. Moreover, the peak at the lower binding energy could be assigned to a La–O–Ti component, while the higher energy peak would correspond to TiO2; alternatively, as attributed by Lv et al. [21] for La2Ti2O7 powders forming rectangular nanosheets, the additional peaks observed at higher binding energies around 460 eV and 465 eV could be from the unsaturated Ti of TiO5 motifs on the Ti-terminated facet of the La2Ti2O7 nanostructures. The spin–orbit splitting (SOS) is 5.7 eV, consistent with the Ti4+ oxidation state. Furthermore, the presence of this additional component may reflect differences between Ti4+ species on the surface and in the bulk of the sample.
The O 1s region of the LTA sample (Figure 5b) can be deconvoluted into four peaks, corresponding to La–O/Ti–O bonds at 529.34 eV [9,27,40], TiO2 (or TiOx) at 530.72 eV, C=O and/or OH groups at 531.94 eV [16,39], and C–O bonds at 533.48 eV, respectively [16,21].
The La 3d spectrum (Figure 5c) of the LTA sample exhibits a spin–orbit splitting of 16.9 eV between the La 3d5/2 and La 3d3/2 peaks, confirming the presence of La3+ [21,40]. Component c2 (green line in Figure 5c), observed at approximately 835–836 eV, corresponds to the “antibonding” orbitals. Component c4 (orange line in Figure 5c), centered around 848 eV, exhibits a broad profile and is attributed to plasmonic loss [69]. Furthermore, the “multiplet splitting” of the La chemical state in this sample (6.6 eV) is larger than the typical values reported for La2O3 (4.6 eV) and La2Ti2O7 (4.3 eV) [26].
The C 1s region is deconvoluted into three distinct peaks (Figure 5d). The peak at 284.6 eV is assigned to C–C bonds, typically arising from carbon contamination on the sample surface. The peak at 286.27 eV corresponds to C=O functional groups, while the peak at 289.07 eV is attributed to O–C=O bonds, indicative of carbonate species. This is consistent with the tendency of lanthanum compounds to form surface carbonates under ambient conditions.
The Ti 2p XPS spectrum for the LTA-Pt sample is presented in Figure 6a. The spectrum of the doped sample exhibits a single chemical state of Ti4+, characterized by a Ti 2p3/2 peak at 458.32 eV. The spin–orbit splitting (SOS) of 5.7 eV between the Ti 2p3/2 and Ti 2p1/2 peaks further confirms the exclusive presence of Ti4+. It is worth mentioning that the additional peak observed around 460 eV is no longer detectable. One reason can be the presence of platinum on the surface of La2Ti2O7 crystals that fulfill the unsaturated Ti in TiO5 motifs on the Ti-terminated facets.
The O 1s region of the LTA-Pt sample (Figure 6b) can be deconvoluted into three peaks, corresponding to La–O/Ti–O bonds at 529.34 eV, C=O and/or OH groups at 531.94 eV, and C–O bonds at 533.48 eV, respectively. The peak at around 531 eV is no longer detectable in the doped sample, which would suggest that this peak is actually given by the presence of a further Ti–O bond in the unmodified case (not an unsaturated Ti termination).
The XPS spectrum of the La 3d region exhibits a typical profile for lanthanum-based compounds, with a similar number of components appearing at comparable binding energies, but in different relative proportions compared to the unmodified sample. This indicates that platinum doping also affects the lanthanum bonding environment. Components 1 and 3 (blue and gray lines in Figure 6c), which correspond to multiplet splitting, are separated by approximately 4.5 eV—significantly less than the splitting observed in the unmodified LTA sample, but closer to the typical values reported for La2O3 and La2Ti2O7, as noted above.
The binding energies of 70.36 eV and 71.55 eV are attributed to Pt 4f7/2. The binding energy at 70.36 eV is attributed to the platinum metal, while the binding energy at 71.55 eV can be attributed to platinum oxide.
The C 1s region has the same behavior as the unmodified sample, with small shifting of the peak values. The spectrum is deconvoluted into three distinct peaks (Figure 6e). The peak at 284.6 eV is assigned to C–C bonds. The peak at 285.84 eV corresponds to C=O functional groups, while the peak at 288.34 eV is attributed to O–C=O bonds, indicative of carbonate species.
Table 3 lists binding energy (BE) and atomic percentage obtained from the analysis of these spectra. The “deconvolutions” were performed using Voigt profiles as detailed by Teodorescu et al. [70].

3.4. Scanning Electron Microscopy and Energy Dispersive X-Ray Spectroscopy (SEM and EDX)

Figure 7 shows the results of the SEM morphology analysis for La2Ti2O7 and La2Ti2O7-Pt nanoparticles. The SEM micrographs highlight the influence of the impregnation of the noble metal on the morphology of the obtained pure La2Ti2O7 nanoparticles. The LTA sample shows irregular particles with a size ranging between 50 and 100 nm. The particles are between 50 and 100 nm in size and appear as clusters of nanoparticles (Figure 7a). In the case of the LTA-Pt sample (Figure 7b), a better individualization of the particles can be observed, as they are less aggregated, with uniformly distributed Pt.
To confirm the elemental composition, EDX measurements were performed. The elemental composition obtained by EDX is presented in Table 4. Qualitative EDX analysis confirms the presence of elements La, Ti, O, and Pt.

3.5. Transmission Electron Microscopy (TEM)

The TEM micrographs of the samples obtained by the sol–gel route are shown in Figure 8 and indicate agglomerated particles that form clusters. Crystals in unmodified samples aggregate into large particle sizes, where crystals can merge. Apparently, the LTA-Pt sample does not change its morphology after impregnation with platinum chloride.
There is no apparent porous structure in the large particles. In Figure 9, a uniform distribution of platinum particles across the entire surface of the La2Ti2O7 particle agglomerates (sample LTA-Pt) can be observed.
Particle size distributions were determined from TEM micrographs for both samples. The mean particle sizes were comparable, measuring approximately 70 nm (69.54 ± 19 nm for the LTA sample and 68.64 ± 22 nm for the LTA-Pt sample). The corresponding particle size distribution histograms are presented in Figure 10.

3.6. X-Ray Diffraction (XRD)

X-ray diffraction (XRD) analysis was performed to determine the crystalline structure of the thermally treated samples. Figure 11a,b display the XRD patterns of titanate nanopowders after heating at 900 °C for 2 h. As shown in Figure 11a, a single phase of La2Ti2O7 was obtained, according to PDF #00-070-0903.
La2Ti2O7 is a layered perovskite [9,15,27,34,39] with a monoclinic structure and space group P21(4) [14,15,16,21,64].
Table 5 summarizes the lattice parameters and crystallite sizes. The lattice constants of the sample LTA closely matched those in the PDF card #00-070-0903.
According to Raman spectroscopy and X-ray photoelectron spectroscopy (XPS) data, the LTA sample contains a titanium oxide-based phase. However, no secondary phases were detected in the X-ray diffraction (XRD) pattern of the LTA sample. This absence may be attributed to the secondary phase being present in an amorphous state, to the possible overlap of its reflections with those of the La2Ti2O7 phase, or to its concentration being below the detection limit of the instrument.
The crystalline structure of the LTA-Pt sample was also analyzed. All the reflections in the XRD diffractograms of the doped sample became sharper and higher after the impregnation with PtCl4 aqueous solution for 24 h and the subsequent drying at 80 °C for 5 h. The modified sample (LTA-Pt) also contained single-phase La2Ti2O7. No platinum-based compounds were detected. X-ray photoelectron spectroscopy (XPS) data indicated that platinum was present in the sample either in its metallic form or as an oxide. If platinum compounds were to crystallize, their diffraction line would completely overlap with those of the La2Ti2O7 phase in the X-ray diffraction (XRD) pattern. The crystallite size slightly increased to 15 nm after platinum impregnation.
The unit cell parameters of the LTA-Pt sample (listed in Table 5) are larger than those of the LTA sample. This increase can be attributed to the incorporation of platinum, which likely partially entered the interlayer spaces between the perovskite slabs.

3.7. X-Ray Fluorescence (XRF)

To analyze the elemental composition of the modified sample, X-ray fluorescence analysis was conducted to determine the presence of the dopant element. The results, presented in Table 6, indicate that platinum (Pt) was detected in quantities closely matching the initially calculated composition.

3.8. Textural Characterization (BET)

To assess the texture of the heat-treated powders, N2 adsorption–desorption isotherms were obtained and are displayed in Figure 12. The textural properties (BET surface area, total pore volume, and average pore size) are provided in Table 7.
Both samples exhibit type IV isotherms with H3 hysteresis loops, according to the IUPAC classification [71]. The small hysteresis loop area and its appearance at relatively high pressures (P/P0 > 0.8) in all samples suggest the presence of relatively large, well-defined mesopores, with some contribution from macropores. This observation is further supported by the pore size distribution (PSD) graphs (shown as inserts in the figures) which confirm broad distributions spanning almost the entire mesopore range and extending into the macropore domain. As expected for materials of this type, the specific surface areas and total pore volumes are relatively low (see Table 7).

3.9. Ultraviolet-Visible Spectroscopy (UV-Vis)

UV–Vis absorption spectra of the thermally treated samples are illustrated in Figure 13. For pure LTA samples, a strong absorption is observed in the UV region that can be assigned to the direct bandgap transition in La2Ti2O7 [17].
For LTA-Pt powder, the characteristic UV absorption of the support is preserved. However, the platinum addition increases this UV absorption band’s intensity and shifts the absorption edge to higher wavelengths. At the same time, this sample exhibits a strong light absorption in the visible region due to the d-d transitions in Pt-doped species of platinum nanoparticles [72,73].
The direct optical bandgap of the samples was calculated using the Tauc method. The obtained experimental values were 3.78 eV for LTA and 3.83 eV for LTA-Pt, being close to the literature data reported for La2Ti2O7 samples [74].

3.10. ROS Monitoring

Figure 14a,b reveal the time-course of the XTT-Formazan presence characterized by a broad absorbance band centered at 470 nm (UV–Vis spectra). This compound results from the XTT sodium salt interaction with the photogenerated superoxide anion radical (·O2) in the presence of the investigated samples (LTA, LTA-Pt) under simulated solar light irradiation. The first 30 min of light exposure is insignificant relative to the ·O2 generation, which clearly increases after 45 min of irradiation time for both samples.
Figure 14c illustrates the relative amount of ·O2 (Imax = 470 nm) produced by the two powders during the irradiation time (120 min), with a higher amount of ·O2 being photogenerated by LTA-Pt.

3.11. Photoluminescence Measurements

The photoluminescence measurements are meant to describe the radiative phenomena induced by the recombination of the photogenerated charges. The almost quenched photoluminescence of the LTA-Pt sample relative to the LTA one in the ethanol/water mixture indicates the better use of the photogenerated charges for a chemical reaction on the catalyst surface than for the recombination process of electrons and holes (Figure 15).

3.12. Photocatalytic Tests

Simulated solar light photo-oxidation processes of organic matter are important for several practical reasons: (i) they require low material and operational costs, (ii) they have the potential to clean water and air by mineralizing organic pollutants to CO2 [35,37,75,76], and (iii) they are attractive alternative routes for the selective synthesis of high-added-value oxygenated products [38,77].
Ethanol is a common compound obtained from biomass, with its photodegradation being important for hydrogen production and depollution. But it can also be present as a contaminant in air and industrial wastewater [37,78].
The procedure can be depicted from Figure 16. The representative samples, namely LTA and LTA-Pt, were selected for photocatalytic tests. The photoactive surface was exposed to simulated solar light (AM1.5). The photocatalytic test lasted three hours.
Photocatalytic tests are conducted at a 20% oxygen concentration. The main intermediate in the ethanol photogeneration process is acetaldehyde, which can be further oxidized in CO2 and H2O.
Table 8 shows the experimental data for reactant and product distribution, and ethanol conversion over bare and platinum-modified powder in the gaseous phase after 3 h of reaction time.
The conversion of ethanol is 14.29% (for LTA) and 29.75% (for LTA-Pt), meaning that platinum modification influences catalytic activity. At the end of the reaction time, both catalysts show the formation of mild oxidation products of ethanol (acetaldehyde and formic acid), as well as carbon dioxide, with the highest value obtained for LTA-Pt.
Figure 17 shows the comparative results for the oxidative degradation of ethanol on LTA and LTA-Pt catalysts exposed to simulated solar light. CH3CHO and HCOOH are formed successively in the complex process of the mineralization of CH3CH2OH to CO2. The blank test has been used to correct the results.
The photodegradation of ethanol is presented in Figure 17, showing reaction sequences to create intermediates.
Figure 17g,h show the photo-oxidative degradation of ethanol to CO2 (and water) under simulated solar light irradiation. The materials investigated show activity in terms of ethanol mineralization in the gaseous phase. The higher production is recorded for LTA-Pt, which produces around 14 μmoles CO2 after 3 h. Both catalysts of interest prove their activity in ethanol photodegradation. Following the first hour of irradiation, there is an increase in the rate of CO2 generation, which is most likely due to the mineralization of previously produced intermediates. In terms of ethanol mineralization, the La2Ti2O7 catalyst is less active than the platinum-modified catalyst by post-synthesis impregnation.
As can be seen in Figure 18, the amount of H2 generated was 0.48 μmoles for the bare LTA sample and 1 μmole for the Pt-modified sample after a 3 h reaction.
The photocatalytic reaction follows first-order kinetics according to the equation “ln(C/C0) = −kt”, where C and C0 designate the ethanol concentration at time 0 and time t, and “k” is the apparent rate constant. Figure 19 shows the graph ln(C0/C) versus the catalyst tested for the photocatalytic degradation of ethanol in the gas phase under simulated solar light irradiation. After three hours of reaction time, the value of the apparent speed k follows the following sequence: LTA (0.0008 min−1) > LTA-Pt (0.0016 min−1). These results show the increasing rate of reaction after the addition of the noble metal, which is in line with Figure 15 showing the decrease in the PL signal for the LTA-Pt sample (relative to LTA) in the water/ethanol solution, meaning a slower recombination of the photogenerated charges and their use for the catalytic process (a higher photoactivity of the LTA-Pt sample).

3.13. Proposed Mechanism of Oxidative Degradation of Ethanol Under Simulated Solar Light

Light-activated photocatalysts generate electrons and holes that, after reaching the catalyst surface, can display a variety of photoredox reactions [35,37,38,79,80,81].
In our case, the photocatalytic oxidative degradation of ethanol produces intermediates (acetaldehyde and formic acid) and end-products (carbon dioxide). The ability of the investigated materials to produce reactive oxygen species (specifically hydroxyl radical ·OH and superoxide anion radical ·O2) was checked, with only the presence of superoxide anion radical being confirmed (Figure 14).
A significant number of photocatalysts, including those based on TiO2, are active, especially under UV or near-UV light irradiation [37,38,82], with their efficiency also being decreased by charge recombination [83]. In order to increase the photocatalyst activity and extend the absorption wavelength range, noble metal doping was performed.
In the present experiment, the oxidative degradation of ethanol (CH3CH2OH) under AM1.5 irradiation on pure and Pt-modified La2Ti2O7 follows a photocatalytic mechanism involving the generation of electron–hole pairs. The reaction pathway includes the oxidation of ethanol to acetaldehyde (CH3CHO) followed by further oxidation to carbon dioxide (CO2), with formic acid (HCOOH) detected as a minor intermediate. Below is the detailed mechanism:
(a)
The process of photon absorption and charge separation (Equation (1)), according to the literature [18,84], can be described as follows:
L a 2 T i 2 O 7 + h ν e + h +
(b)
Pt-modifying role—Platinum (Pt) enhances charge separation by trapping electrons and reducing recombination (Equation (2)), leading to more efficient ethanol oxidation and CO2 formation [35,85,86] and promoting the adsorption and reduction of the oxygen (Equation (3)).
P t + e P t
O 2 + e O 2 ( a d s ) O ( a c t i v e o x y g e n )
(c)
Oxidative degradation of ethanol by photogenerated holes—ethanol is oxidized by holes (h+) to form the ethoxy radical (CH3CH2O·) (Equation (4)) and this undergoes further oxidation, yielding acetaldehyde (CH3CHO) (Equations (4)–(6)). Acetaldehyde is also oxidized by holes to generate formic acid (HCOOH) in trace amounts, finally leading to CO2 (Equations (7) and (8)) [37,87,88,89,90,91].
C H 3 C H 2 O H + h + C H 3 C H 2 O · + H +
C H 3 C H 2 O · + [ O ] C H 3 C H 2 O O ·
C H 3 C H 2 O O · + h + + e C H 3 C H O + H 2 O
C H 3 C H O + [ O ] H C O O H + H 2 O
H C O O H + O C O 2 + H 2 O
C H 3 C H 2 O H s l o w l y C H 3 C H O H C O O H r a p i d l y C O 2
(d)
Photogenerated electrons (e) can reduce protons (H+) to form molecular hydrogen (H2) (Equation (9)) [92].
2 H + + 2 e H 2
This mechanism demonstrates how ethanol undergoes oxidative degradation in the gaseous phase when exposed to simulated solar light.
The literature data present the use of La2Ti2O7 mostly for a liquid phase reaction targeting the hydrogen evolution and water depollution by organic compound degradation, especially for dyes [93,94].
Li et al. [46] obtained, by hydrothermal synthesis, La2Ti2O7 two-dimensional (2D) nanosheets and used them as photocatalysts for the decolorization of methyl orange solution and the evolution of H2 from water–ethanol solution, registering rates of H2 evolution of about 750 μmol g−1 h−1 at 160 min.
In conclusion, this study presents the photocatalytic degradation of ethanol vapor under simulated solar light with 20% oxygen in the presence of pure or platinum-modified La2Ti2O7 nanoparticles. The main products are acetaldehyde and CO2. Two samples were tested in ethanol photodegradation experiments, and showed photocatalytic activity for ethanol mineralization. The photodegradation of ethanol in simulated solar light brings new data concerning the use of double lanthanum titanium oxide obtained by the sol–gel method as an efficient depollution photocatalyst.

4. Conclusions

Layered perovskite La2Ti2O7 was prepared by the sol–gel method. The effect of noble metal addition on the properties of the resulting Pt-modified La2Ti2O7 was evaluated.
The morphology of the samples showed aggregates of nanoparticles with well-defined mesopores.
XRD analysis indicated the formation of single-phase La2Ti2O7 and the TEM micrographs indicated a uniform distribution of platinum across the surface of the La2Ti2O7.
Oxidative degradation of the ethanol in the gaseous phase under simulated solar light irradiation followed a photocatalytic mechanism involving the generation of electron–hole pairs and superoxide anion radical leading to active oxygen.
The photocatalytic activity in terms of ethanol conversion and selectivity to CO2 was increased by the addition of platinum to the La2Ti2O7 catalyst, with platinum enhancing the charge separation but also the adsorption and reduction of the oxygen. Accordingly, these powders showed the potential for application in air depollution technologies.

Author Contributions

Conceptualization, A.I., L.P., C.A., I.B. and M.Z.; methodology, A.I., L.P., C.A., I.B. and M.Z.; investigation, A.I., S.P., I.S.H., R.M.C., D.C.C. and V.B.; resources, I.B. and M.Z.; writing—original draft preparation, A.I., L.P., C.A., S.P., D.C.C., V.B., I.B. and M.Z.; writing—review and editing, A.I., L.P., C.A., S.P., I.S.H., R.M.C., I.B. and M.Z.; visualization, A.I., L.P. and M.Z. 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 original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

We thank Florica Doroftei from “Petru Poni” Institute of Macromolecular Chemistry of the Romanian Academy for her help in the microscopic analysis of the samples presented in this paper.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Chen, D.; Cheng, Y.; Zhou, N.; Chen, P.; Wang, Y.; Li, K.; Huo, S.; Cheng, P.; Peng, P.; Zhang, R.; et al. Photocatalytic Degradation of Organic Pollutants Using TiO2-Based Photocatalysts: A Review. J. Clean. Prod. 2020, 268, 121725. [Google Scholar] [CrossRef]
  2. Li, X.; Wei, H.; Song, T.; Lu, H.; Wang, X. A Review of the Photocatalytic Degradation of Organic Pollutants in Water by Modified TiO2. Water Sci. Technol. 2023, 88, 1495–1507. [Google Scholar] [CrossRef] [PubMed]
  3. Bhagwat, U.O.; Maridevaru, M.C.; Al Souwaileh, A.; Wu, J.J.; Anandan, S. Layered Perovskite La2Ti2O7 Nanostructures for Photocatalytic Degradation of Congo Red Dye Pollutant. ChemistrySelect 2024, 9, e202403511. [Google Scholar] [CrossRef]
  4. Wang, Q.; Zhang, G.; Zhang, C.; Xu, F.; Zhang, Y.; Fu, W.; Liu, J.; Li, J. Enhanced Mineralization of Organic Pollutants through Atomic Hydrogen-Mediated Alternative Transformation Pathways. Environ. Sci. Technol. 2024, 58, 11185–11192. [Google Scholar] [CrossRef] [PubMed]
  5. Zhou, H.; Wang, H.; Yue, C.; He, L.; Li, H.; Zhang, H.; Yang, S.; Ma, T. Photocatalytic Degradation by TiO2-Conjugated/Coordination Polymer Heterojunction: Preparation, Mechanisms, and Prospects. Appl. Catal. B 2024, 344, 123605. [Google Scholar] [CrossRef]
  6. Ruziwa, D.T.; Oluwalana, A.E.; Mupa, M.; Meili, L.; Selvasembian, R.; Nindi, M.M.; Sillanpaa, M.; Gwenzi, W.; Chaukura, N. Pharmaceuticals in Wastewater and Their Photocatalytic Degradation Using Nano-Enabled Photocatalysts. J. Water Process Eng. 2023, 54, 103880. [Google Scholar] [CrossRef]
  7. Ashouri, F.; Khoobi, M.; Ganjali, M.R.; Karimi, M.S. Construction, Characterization, and Photocatalytic Study of La2Ti2O7/C3N4+xHy and La2Ti2O7/GO Nanocomposites as Efficient Catalysts toward Photodegradation of Harmful Organic Dyes. J. Photochem. Photobiol. A Chem. 2023, 435, 114279. [Google Scholar] [CrossRef]
  8. Tsoukleris, D.S.; Gatou, M.-A.; Lagopati, N.; Sygellou, L.; Christodouleas, D.C.; Falaras, P.; Pavlatou, E.A. Chemically Modified TiO2 Photocatalysts as an Alternative Disinfection Approach for Municipal Wastewater Treatment Plant Effluents. Water 2023, 15, 2052. [Google Scholar] [CrossRef]
  9. Zhong, L.; Li, X.; Zhan, Y.; Liu, Y.; Yang, K.; Liu, J.; Liu, S.; Jiang, J.; Li, L.; Sakata, Y. Sodium-Doped La2Ti2O7 Synthesized by Molten Salt Synthesis Method for Photocatalytic Overall Water Splitting. Appl. Catal. A Gen. 2025, 692, 120079. [Google Scholar] [CrossRef]
  10. Maitlo, H.A.; Anand, B.; Kim, K.-H. TiO2-Based Photocatalytic Generation of Hydrogen from Water and Wastewater. Appl. Energy 2024, 361, 122932. [Google Scholar] [CrossRef]
  11. Núñez Valdez, M.; Spaldin, N.A. Origin and Evolution of Ferroelectricity in the Layered Rare-Earth-Titanate, R2Ti2O7, Carpy-Galy Phases. Polyhedron 2019, 171, 181–192. [Google Scholar] [CrossRef]
  12. Herrera, G.; Jiménez-Mier, J.; Chavira, E. Layered-Structural Monoclinic–Orthorhombic Perovskite La2Ti2O7 to Orthorhombic LaTiO3 Phase Transition and Their Microstructure Characterization. Mater. Charact. 2014, 89, 13–22. [Google Scholar] [CrossRef]
  13. Xu, J.; Zhang, Y.; Xu, X.; Fang, X.; Xi, R.; Liu, Y.; Zheng, R.; Wang, X. Constructing La2B2O7 (B = Ti, Zr, Ce) Compounds with Three Typical Crystalline Phases for the Oxidative Coupling of Methane: The Effect of Phase Structures, Superoxide Anions, and Alkalinity on the Reactivity. ACS Catal. 2019, 9, 4030–4045. [Google Scholar] [CrossRef]
  14. Leroy, S.; Blach, J.-F.; Kopia, A.; Lech, S.; Cieniek, Ł.; Kania, N.; Saitzek, S. Enhancement of Photocatalytic Properties of Nanosized La2Ti2O7 Synthesized by Glycine-Assisted Sol-Gel Route. J. Photochem. Photobiol. A Chem. 2022, 426, 113739. [Google Scholar] [CrossRef]
  15. Leroy, S.; Blach, J.-F.; Huvé, M.; Léger, B.; Kania, N.; Henninot, J.-F.; Ponchel, A.; Saitzek, S. Photocatalytic and Sonophotocatalytic Degradation of Rhodamine B by Nano-Sized La2Ti2O7 Oxides Synthesized with Sol-Gel Method. J. Photochem. Photobiol. A Chem. 2020, 401, 112767. [Google Scholar] [CrossRef]
  16. Lv, L.; Lei, L.; Chen, Q.-W.; Yin, C.-L.; Fan, H.; Zhou, J.-P. Oxygen Vacancies-Modified S-Scheme Heterojunction of Bi-Doped La2Ti2O7 and La-Doped Bi4Ti3O12 to Improve the NO Gas Removal Avoiding NO2 Product. Appl. Catal. B 2024, 343, 123464. [Google Scholar] [CrossRef]
  17. Wang, Z.; Teramura, K.; Hosokawa, S.; Tanaka, T. Photocatalytic Conversion of CO2 in Water over Ag-Modified La2Ti2O7. Appl. Catal. B 2015, 163, 241–247. [Google Scholar] [CrossRef]
  18. Tan, Y.; Li, Z.; Tian, Y.; Hu, S.; Jia, L.; Chi, B.; Pu, J.; Li, J. Charge-Compensated (Nb, Fe)-Codoped La2Ti2O7 Photocatalyst for Photocatalytic H2 Production and Optical Absorption. J. Alloys Compd. 2017, 709, 277–284. [Google Scholar] [CrossRef]
  19. Szczepanski, F.; Bayart, A.; Katelnikovas, A.; Blach, J.-F.; Rousseau, J.; Saitzek, S. Luminescence and Up-Conversion Properties in La2Ti2O7:Eu3+,Er3+ Oxides under UV and NIR Radiations towards a Two-Color Sensor. J. Alloys Compd. 2020, 826, 154157. [Google Scholar] [CrossRef]
  20. Onozuka, K.; Kawakami, Y.; Imai, H.; Yokoi, T.; Tatsumi, T.; Kondo, J.N. Perovskite-Type La2Ti2O7 Mesoporous Photocatalyst. J. Solid State Chem. 2012, 192, 87–92. [Google Scholar] [CrossRef]
  21. Lv, L.; Yang, H.-D.; Fan, H.; Yang, L.; Chen, Q.-W.; Zhou, J.-P. La2Ti2O7 Nanosheets Synthesized under Magnetic Field for Ofloxacin Ferrophotocatalytic Degradation. J. Environ. Chem. Eng. 2022, 10, 108088. [Google Scholar] [CrossRef]
  22. Hou, W.-M.; Ku, Y. Synthesis and Characterization of La2Ti2O7 Employed for Photocatalytic Degradation of Reactive Red 22 Dyestuff in Aqueous Solution. J. Alloys Compd. 2011, 509, 5913–5918. [Google Scholar] [CrossRef]
  23. Raciulete, M.; Anastasescu, C.; Papa, F.; Atkinson, I.; Bradu, C.; Negrila, C.; Eftemie, D.-I.; Culita, D.C.; Miyazaki, A.; Bratan, V.; et al. Band-Gap Engineering of Layered Perovskites by Cu Spacer Insertion as Photocatalysts for Depollution Reaction. Catalysts 2022, 12, 1529. [Google Scholar] [CrossRef]
  24. Temerov, F.; Baghdadi, Y.; Rattner, E.; Eslava, S. A Review on Halide Perovskite-Based Photocatalysts: Key Factors and Challenges. ACS Appl. Energy Mater. 2022, 5, 14605–14637. [Google Scholar] [CrossRef]
  25. Huang, H.; Pradhan, B.; Hofkens, J.; Roeffaers, M.B.J.; Steele, J.A. Solar-Driven Metal Halide Perovskite Photocatalysis: Design, Stability, and Performance. ACS Energy Lett. 2020, 5, 1107–1123. [Google Scholar] [CrossRef]
  26. Li, M.-J.; Hsu, T.-H.; Huang, C.-L. Electrical Characterization of Sol-Gel La2Ti2O7 Films for Resistive Random Access Memory Applications. Mater. Sci. Semicond. Process 2023, 158, 107370. [Google Scholar] [CrossRef]
  27. Qiao, B.; Jiang, Y.; Yao, T.; Tao, A.; Yan, X.; Gao, C.; Li, X.; Ohta, H.; Chen, C.; Ma, X.-L.; et al. Microstructure and Electrical Properties of La2Ti2O7 Thin Films on SrTiO3 Substrates. Appl. Surf. Sci. 2022, 587, 151599. [Google Scholar] [CrossRef]
  28. Li, J.; Zhao, Y.; Xia, M.; An, H.; Bai, H.; Wei, J.; Yang, B.; Yang, G. Highly Efficient Charge Transfer at 2D/2D Layered P-La2Ti2O7/Bi2WO6 Contact Heterojunctions for Upgraded Visible-Light-Driven Photocatalysis. Appl. Catal. B 2020, 261, 118244. [Google Scholar] [CrossRef]
  29. Cai, X.; Mao, L.; Fujitsuka, M.; Majima, T.; Kasani, S.; Wu, N.; Zhang, J. Effects of Bi-Dopant and Co-Catalysts upon Hole Surface Trapping on La2Ti2O7 Nanosheet Photocatalysts in Overall Solar Water Splitting. Nano Res. 2022, 15, 438–445. [Google Scholar] [CrossRef]
  30. Gomez-Romero, P.; Pokhriyal, A.; Rueda-García, D.; Bengoa, L.N.; González-Gil, R.M. Hybrid Materials: A Metareview. Chem. Mater. 2024, 36, 8–27. [Google Scholar] [CrossRef]
  31. Yang, S.; Qu, H.; Zhou, X. Designing Visible-Light-Responsive La2Ti2O7 Photocatalysts by Surface Co-Doping of Nonmetal and Metal Combinations. New J. Chem. 2023, 47, 22469–22480. [Google Scholar] [CrossRef]
  32. Marcela Prieto-Zuleta, L.; Giovanny Villabona-Leal, E.; Joazet Ojeda-Galván, H.; Bao, Y.; Ocampo-Pérez, R.; Oros Ruíz, S.; Song, S.; Gabriel Rodríguez, Á.; Ricardo Navarro-Contreras, H.; Quintana, M. Enhanced Photocatalytic Activity of 0D/2D/2D MNPs@La2Ti2O7@BiOBr (M = Ag, Au) Heterojunction in the Degradation of Organic Contaminants and the Production of Hydrogen by Simulated Solar Irradiation. Appl. Surf. Sci. 2025, 684, 161894. [Google Scholar] [CrossRef]
  33. Bai, L.; Cai, Z.; Xu, Q. A Plasmonic Photocatalyst of Ag@AgBr/La2Ti2O7 with Enhanced Photocatalytic Activity under Visible Light Irradiation. Appl. Phys. A 2021, 127, 595. [Google Scholar] [CrossRef]
  34. Wang, Y.; Hu, J.; Zhai, C.; Gao, H.; Liu, Z.-Q.; Du, Y.; Zhu, M. CdS Quantum Dots Sensitized 2D La2Ti2O7 Nanosheets as Support for Visible Light-Assisted Electrocatalytic Methanol Oxidation in Alkaline Medium. Energy Technol. 2019, 7, 1800539. [Google Scholar] [CrossRef]
  35. Goncearenco, E.; Morjan, I.P.; Dutu, E.; Scarisoreanu, M.; Fleaca, C.; Gavrila-Florescu, L.; Dumitrache, F.; Banici, A.M.; Teodorescu, V.S.; Anastasescu, C.; et al. The Effect of Noble Metal Addition on the Properties of Oxide Semiconductors Nanoparticles. J. Solid. State Chem. 2022, 307, 122817. [Google Scholar] [CrossRef]
  36. Hsieh, H.-C.; Chen, R.-C.; Huang, Y.-K.; Sheu, H.-S.; Chuang, Y.-C.; Wang, C.-W.; Lee, C.-S. Enhancing Catalytic Performance in Oxidative Steam Reforming of Ethanol: The Role of Ruthenium Ion Substitution in Layered Perovskite La2Ti2–xRuxO7±δ Catalysts. J. Phys. Chem. C 2024, 128, 19570–19585. [Google Scholar] [CrossRef]
  37. Goncearenco, E.; Morjan, I.P.; Fleaca, C.T.; Dumitrache, F.; Dutu, E.; Scarisoreanu, M.; Teodorescu, V.S.; Sandulescu, A.; Anastasescu, C.; Balint, I. Titania Nanoparticles for Photocatalytic Degradation of Ethanol under Simulated Solar Light. Beilstein J. Nanotechnol. 2023, 14, 616–630. [Google Scholar] [CrossRef]
  38. Sandulescu, A.; Anastasescu, C.; Papa, F.; Raciulete, M.; Vasile, A.; Spataru, T.; Scarisoreanu, M.; Fleaca, C.; Mihailescu, C.N.; Teodorescu, V.S.; et al. Advancements on Basic Working Principles of Photo-Driven Oxidative Degradation of Organic Substrates over Pristine and Noble Metal-Modified TiO2. Model Case of Phenol Photo Oxidation. Catalysts 2021, 11, 487. [Google Scholar] [CrossRef]
  39. Dong, Y.; Li, G.; Xu, D.; Wang, Q.; Yang, T.; Pang, S.; Zhang, G.; Lv, L.; Xia, Y.; Ren, Z.; et al. A Novel Hierarchical Heterostructure of Hollow La2Ti2O7/In2O3 with Strong Interface Interaction for Photocatalytic Antibiotic Degradation. Chem. Eng. J. 2022, 446, 136705. [Google Scholar] [CrossRef]
  40. Wang, H.; Zhang, Y.; Ma, Z.; Zhang, W. Role of PEG2000 on Sol-Gel Preparation of Porous La2Ti2O7 for Enhanced Photocatalytic Activity on Ofloxacin Degradation. Mater. Sci. Semicond. Process 2019, 91, 151–158. [Google Scholar] [CrossRef]
  41. Chaillot, D.; Miehé-Brendlé, J.; Bennici, S. Insights on the Influence of the Precursors on the Sol–Gel Synthesis of Hybrid Organic–Inorganic Saponite-like Materials. Comptes Rendus Chim. 2019, 22, 258–268. [Google Scholar] [CrossRef]
  42. Chang, C.; Rad, S.; Gan, L.; Li, Z.; Dai, J.; Shahab, A. Review of the Sol–Gel Method in Preparing Nano TiO2 for Advanced Oxidation Process. Nanotechnol. Rev. 2023, 12, 20230150. [Google Scholar] [CrossRef]
  43. Fuierer, P.A.; Newnham, R.E. La2Ti2O7 Ceramics. J. Am. Ceram. Soc. 1991, 74, 2876–2881. [Google Scholar] [CrossRef]
  44. Takahashi, J.; Ohtsuka, T. Vibrational Spectroscopic Study of Structural Evolution in the Coprecipitated Precursors to La2Sn2O7 and La2Ti2O7. J. Am. Ceram. Soc. 1989, 72, 426–431. [Google Scholar] [CrossRef]
  45. Suresh, M.; Prasadarao, A.V.; Komarneni, S. Mixed Hydroxide Precursors for La2Ti2O7 and Nd2Ti2O7 by Homogeneous Precipitation. J. Electroceram. 2001, 6, 147–151. [Google Scholar] [CrossRef]
  46. Li, K.; Wang, Y.; Wang, H.; Zhu, M.; Yan, H. Hydrothermal Synthesis and Photocatalytic Properties of Layered La2Ti2O7 Nanosheets. Nanotechnology 2006, 17, 4863–4867. [Google Scholar] [CrossRef]
  47. Kim, H.G.; Hwang, D.W.; Bae, S.W.; Jung, J.H.; Lee, J.S. Photocatalytic Water Splitting over La2Ti2O7 Synthesized by the Polymerizable Complex Method. Catal. Lett. 2003, 91, 193–198. [Google Scholar] [CrossRef]
  48. Shcherbakova, L.G.; Mamsurova, L.G.; Sukhanova, G.E. Lanthanide Titanates. Russ. Chem. Rev. 1979, 48, 228–242. [Google Scholar] [CrossRef]
  49. Balachandran, U.; Eror, N.G. X-Ray Diffraction and Vibrational-Spectroscopy Study of the Structure of La2Ti2O7. J. Mater. Res. 1989, 4, 1525–1528. [Google Scholar] [CrossRef]
  50. Favacho, V.S.S.; Melo, D.M.A.; Costa, J.E.L.; Silva, Y.K.R.O.; Braga, R.M.; Medeiros, R.L.B.A. Perovskites Synthesized by Soft Template-Assisted Hydrothermal Method: A Bibliometric Analysis and New Insights. Int. J. Hydrogen Energy 2024, 78, 1391–1428. [Google Scholar] [CrossRef]
  51. Walton, R.I. Perovskite Oxides Prepared by Hydrothermal and Solvothermal Synthesis: A Review of Crystallisation, Chemistry, and Compositions. Chem. Eur. J. 2020, 26, 9041–9069. [Google Scholar] [CrossRef] [PubMed]
  52. Hu, L.; Wang, J.; Li, W.; Long, Y. Microwave-Assisted Homogeneous Precipitation Route for the Synthesis of La2Ti2O7:Tb3+ Nanophosphors. Mater. Lett. 2023, 350, 134984. [Google Scholar] [CrossRef]
  53. Lu, X.; Tang, P.; Zhang, J.; Xu, H.; Wen, S.; Lou, Y.; Liu, Y.; Chen, H. Preparation of La2Ti2O7 by Sol-Gel Method and Its Photocatalytic Activity. Integr. Ferroelectr. 2023, 234, 108–114. [Google Scholar] [CrossRef]
  54. Navas, D.; Fuentes, S.; Castro-Alvarez, A.; Chavez-Angel, E. Review on Sol-Gel Synthesis of Perovskite and Oxide Nanomaterials. Gels 2021, 7, 275. [Google Scholar] [CrossRef] [PubMed]
  55. Choukchou Braham, Z.; Kermad, A.; El Korso, S.; Ramdani, M.R.; Boudjemaa, A.; Bachari, K.; Choukchou Braham, A. Enhancing Structural Properties of Simple Perovskite Materials Based on Zirconate and Titanate Prepared by Sol-Gel Method. Mater. Chem. Phys. 2023, 310, 128460. [Google Scholar] [CrossRef]
  56. Bokov, D.; Turki Jalil, A.; Chupradit, S.; Suksatan, W.; Javed Ansari, M.; Shewael, I.H.; Valiev, G.H.; Kianfar, E. Nanomaterial by Sol-Gel Method: Synthesis and Application. Adv. Mater. Sci. Eng. 2021, 2021, 102014. [Google Scholar] [CrossRef]
  57. Parashar, M.; Shukla, V.K.; Singh, R. Metal Oxides Nanoparticles via Sol–Gel Method: A Review on Synthesis, Characterization and Applications. J. Mater. Sci. Mater. Electron. 2020, 31, 3729–3749. [Google Scholar] [CrossRef]
  58. Esposito, S. “Traditional” Sol-Gel Chemistry as a Powerful Tool for the Preparation of Supported Metal and Metal Oxide Catalysts. Materials 2019, 12, 668. [Google Scholar] [CrossRef]
  59. Brinker, J.C.; Scherer, G.W. Sol-Gel Science; Academic Press: Cambridge, MA, USA, 1990; ISBN 9780080571034. [Google Scholar]
  60. Dantelle, G.; Beauquis, S.; Le Dantec, R.; Monnier, V.; Galez, C.; Mugnier, Y. Solution-Based Synthesis Routes for the Preparation of Noncentrosymmetric 0-D Oxide Nanocrystals with Perovskite and Nonperovskite Structures. Small 2022, 18, 2200992. [Google Scholar] [CrossRef]
  61. Danks, A.E.; Hall, S.R.; Schnepp, Z. The Evolution of ‘Sol–Gel’ Chemistry as a Technique for Materials Synthesis. Mater. Horiz. 2016, 3, 91–112. [Google Scholar] [CrossRef]
  62. Anastasescu, C.; Negrila, C.; Angelescu, D.G.; Atkinson, I.; Anastasescu, M.; Spataru, N.; Zaharescu, M.; Balint, I. Particularities of Photocatalysis and Formation of Reactive Oxygen Species on Insulators and Semiconductors: Cases of SiO2, TiO2 and Their Composite SiO2–TiO2. Catal. Sci. Technol. 2018, 8, 5657–5668. [Google Scholar] [CrossRef]
  63. Preda, S.; Anastasescu, C.; Balint, I.; Umek, P.; Sluban, M.; Negrila, C.C.; Angelescu, D.G.; Bratan, V.; Rusu, A.; Zaharescu, M. Charge Separation and ROS Generation on Tubular Sodium Titanates Exposed to Simulated Solar Light. Appl. Surf. Sci. 2019, 470, 1053–1063. [Google Scholar] [CrossRef]
  64. Swami, R.; Bokolia, R.; Sreenivas, K. Effects of Sintering Temperature on Structural, Electrical and Ferroelectric Properties of La2Ti2O7 Ceramics. Ceram. Int. 2020, 46, 26790–26799. [Google Scholar] [CrossRef]
  65. Abe, Y.; Laine, R.M. Photocatalytic Plate-like La2Ti2O7 Nanoparticles Synthesized via Liquid-feed Flame Spray Pyrolysis (LF-FSP) of Metallo-organic Precursors. J. Am. Ceram. Soc. 2020, 103, 4832–4839. [Google Scholar] [CrossRef]
  66. Zhang, S.; Xu, J.; Lu, C.; Ouyang, R.; Ma, J.; Zhong, X.; Fang, X.; Xu, X.; Wang, X. Preparation Method Investigation and Structure Identification by XRD and Raman Techniques for A2B2O7 Composite Oxides. J. Am. Ceram. Soc. 2024, 107, 3475–3496. [Google Scholar] [CrossRef]
  67. Chen, C.; Gao, Z.; Yan, H.; Reece, M.J. Crystallographic Structure and Ferroelectricity of (AxLa1−x)2Ti2O7(A = Sm and Eu) Solid Solutions with High TC. J. Am. Ceram. Soc. 2016, 99, 523–530. [Google Scholar] [CrossRef]
  68. Ocal, C.; Ferrer, S. Low Temperature Diffusion of Pt and Au Atoms through Thin TiO2 Films on a Ti Substrate. Surf. Sci. 1987, 191, 147–156. [Google Scholar] [CrossRef]
  69. Sunding, M.F.; Hadidi, K.; Diplas, S.; Løvvik, O.M.; Norby, T.E.; Gunnæs, A.E. XPS Characterisation of in Situ Treated Lanthanum Oxide and Hydroxide Using Tailored Charge Referencing and Peak Fitting Procedures. J. Electron. Spectros Relat. Phenom. 2011, 184, 399–409. [Google Scholar] [CrossRef]
  70. Teodorescu, C.M.; Esteva, J.M.; Karnatak, R.C.; El Afif, A. An Approximation of the Voigt I Profile for the Fitting of Experimental X-Ray Absorption Data. Nucl. Instrum. Methods Phys. Res. A 1994, 345, 141–147. [Google Scholar] [CrossRef]
  71. 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]
  72. Alshehri, A.; Narasimharao, K. PtOx-TiO2 Anatase Nanomaterials for Photocatalytic Reformation of Methanol to Hydrogen: Effect of TiO2 Morphology. J. Mater. Res. Technol. 2020, 9, 14907–14921. [Google Scholar] [CrossRef]
  73. Vasile, A.; Papa, F.; Bratan, V.; Munteanu, C.; Teodorescu, M.; Atkinson, I.; Anastasescu, M.; Kawamoto, D.; Negrila, C.; Ene, C.D.; et al. Water Denitration over Titania-Supported Pt and Cu by Combined Photocatalytic and Catalytic Processes: Implications for Hydrogen Generation Properties in a Photocatalytic System. J. Environ. Chem. Eng. 2022, 10, 107129. [Google Scholar] [CrossRef]
  74. Ma, Z.; Li, Y.; Lv, Y.; Sa, R.; Li, Q.; Wu, K. Synergistic Effect of Doping and Compositing on Photocatalytic Efficiency: A Case Study of La2Ti2O7. ACS Appl. Mater. Interfaces 2018, 10, 39327–39335. [Google Scholar] [CrossRef]
  75. Li, X.; Xie, J.; Jiang, C.; Yu, J.; Zhang, P. Review on Design and Evaluation of Environmental Photocatalysts. Front. Environ. Sci. Eng. 2018, 12, 14. [Google Scholar] [CrossRef]
  76. Chen, L.; Tang, J.; Song, L.-N.; Chen, P.; He, J.; Au, C.-T.; Yin, S.-F. Heterogeneous Photocatalysis for Selective Oxidation of Alcohols and Hydrocarbons. Appl. Catal. B 2019, 242, 379–388. [Google Scholar] [CrossRef]
  77. Bellardita, M.; Loddo, V.; Augugliaro, V.; Palmisano, L.; Yurdakal, S. Selective Photocatalytic and Photoelectrocatalytic Synthesis of Valuable Compounds in Aqueous Medium. Catal. Today 2024, 432, 114587. [Google Scholar] [CrossRef]
  78. Pavel, M.; Anastasescu, C.; State, R.-N.; Vasile, A.; Papa, F.; Balint, I. Photocatalytic Degradation of Organic and Inorganic Pollutants to Harmless End Products: Assessment of Practical Application Potential for Water and Air Cleaning. Catalysts 2023, 13, 380. [Google Scholar] [CrossRef]
  79. Xiong, L.; Tang, J. Strategies and Challenges on Selectivity of Photocatalytic Oxidation of Organic Substances. Adv. Energy Mater. 2021, 11, 2003216. [Google Scholar] [CrossRef]
  80. Lai, Y.; Zeng, Y.; Chen, X.; Wang, T.; Yang, X.; Guo, Q. Photochemistry of Ethanol on Rutile TiO2 (110): Breaking Two Bonds with One Hole. J. Phys. Chem. C 2023, 127, 1863–1869. [Google Scholar] [CrossRef]
  81. Li, X.; Chen, Y.; Tao, Y.; Shen, L.; Xu, Z.; Bian, Z.; Li, H. Challenges of Photocatalysis and Their Coping Strategies. Chem. Catal. 2022, 2, 1315–1345. [Google Scholar] [CrossRef]
  82. Guo, Q.; Zhou, C.; Ma, Z.; Yang, X. Fundamentals of TiO2 Photocatalysis: Concepts, Mechanisms, and Challenges. Adv. Mater. 2019, 31, 1901997. [Google Scholar] [CrossRef]
  83. Mandal, A.; Kargupta, K. Cu-Doped 2D-Bi2MoO6 Nanoribbon/RGO Photocatalysts for Selective Ethanol Production by Photocatalytic CO2 Reduction. ACS Appl. Nano Mater. 2025, 8, 3471–3486. [Google Scholar] [CrossRef]
  84. Cai, X.; Zhu, M.; Elbanna, O.A.; Fujitsuka, M.; Kim, S.; Mao, L.; Zhang, J.; Majima, T. Au Nanorod Photosensitized La2Ti2O7 Nanosteps: Successive Surface Heterojunctions Boosting Visible to Near-Infrared Photocatalytic H2 Evolution. ACS Catal. 2018, 8, 122–131. [Google Scholar] [CrossRef]
  85. Al-Madanat, O.; Curti, M.; Günnemann, C.; AlSalka, Y.; Dillert, R.; Bahnemann, D.W. TiO2 Photocatalysis: Impact of the Platinum Loading Method on Reductive and Oxidative Half-Reactions. Catal. Today 2021, 380, 3–15. [Google Scholar] [CrossRef]
  86. Lee, S.-K.; Mills, A. Platinum and Palladium in Semiconductor Photocatalytic Systems. Platin. Met. Rev. 2003, 47, 61–72. [Google Scholar] [CrossRef]
  87. Mamaghani, A.H.; Haghighat, F.; Lee, C.-S. Photocatalytic Oxidation Technology for Indoor Environment Air Purification: The State-of-the-Art. Appl. Catal. B 2017, 203, 247–269. [Google Scholar] [CrossRef]
  88. Shiraishi, Y.; Shiota, S.; Hirakawa, H.; Tanaka, S.; Ichikawa, S.; Hirai, T. Titanium Dioxide/Reduced Graphene Oxide Hybrid Photocatalysts for Efficient and Selective Partial Oxidation of Cyclohexane. ACS Catal. 2017, 7, 293–300. [Google Scholar] [CrossRef]
  89. Katsiev, K.; Harrison, G.; Alghamdi, H.; Alsalik, Y.; Wilson, A.; Thornton, G.; Idriss, H. Mechanism of Ethanol Photooxidation on Single-Crystal Anatase TiO2 (101). J. Phys. Chem. C 2017, 121, 2940–2950. [Google Scholar] [CrossRef]
  90. Kumar, A.; Mondal, K.; Rajakumar, B. A Combined Experimental and Theoretical Study to Determine the Kinetics of 2-Ethoxy Ethanol with OH Radical in the Gas Phase. J. Phys. Chem. A 2021, 125, 8869–8881. [Google Scholar] [CrossRef]
  91. Vorontsov, A.; Dubovitskaya, V.P. Selectivity of Photocatalytic Oxidation of Gaseous Ethanol over Pure and Modified TiO2. J. Catal. 2004, 221, 102–109. [Google Scholar] [CrossRef]
  92. Trogadas, P.; Coppens, M.-O. Nature-Inspired Electrocatalysts and Devices for Energy Conversion. Chem. Soc. Rev. 2020, 49, 3107–3141. [Google Scholar] [CrossRef] [PubMed]
  93. Wang, Z.; Yu, H.; Liu, Z. Oxygen Vacancies Defective La2Ti2O7 Nanosheets Enhanced Photocatalytic Activity of Hydrogen Evolution under Visible Light Irradiation. Molecules 2023, 28, 5792. [Google Scholar] [CrossRef]
  94. Fernández, L.; Bustos, F.; Correa, D.; Seguel, M.; Suarez, C.; Caro, C.; Leyton, P.; Cabello-Guzmán, G. A Photochemical Route in the Synthesis and Characterization of La2Ti2O7 and La2Ti2O7/AgO Films and Its Evaluation in Congo Red Degradation and as Antibacterial Control to Staphylococcus Aureus. Environ. Sci. Pollut. Res. 2023, 30, 107580–107597. [Google Scholar] [CrossRef]
Figure 1. The flowchart of the sample preparation by the sol–gel method.
Figure 1. The flowchart of the sample preparation by the sol–gel method.
Applsci 15 07665 g001
Figure 2. FT-IR spectra of the (a) LTA and (b) LTA-Pt.
Figure 2. FT-IR spectra of the (a) LTA and (b) LTA-Pt.
Applsci 15 07665 g002
Figure 3. Vibration bands in Raman spectra of thermally treated samples. The arrows and red frames highlight the increasing intensities of the 446, 488, and 668 cm−1 bands. The blue shift from the 342 cm−1 band in the LTA sample to 337 cm−1 in the LTA-Pt sample is indicated by red numbers.
Figure 3. Vibration bands in Raman spectra of thermally treated samples. The arrows and red frames highlight the increasing intensities of the 446, 488, and 668 cm−1 bands. The blue shift from the 342 cm−1 band in the LTA sample to 337 cm−1 in the LTA-Pt sample is indicated by red numbers.
Applsci 15 07665 g003
Figure 4. XPS survey spectra (black line—LTA sample; blue line—LTA-Pt sample).
Figure 4. XPS survey spectra (black line—LTA sample; blue line—LTA-Pt sample).
Applsci 15 07665 g004
Figure 5. XPS spectra and the deconvolutions for the LTA sample for (a) Ti 2p, (b) O 1s, (c) La 3d, and (d) C 1s.
Figure 5. XPS spectra and the deconvolutions for the LTA sample for (a) Ti 2p, (b) O 1s, (c) La 3d, and (d) C 1s.
Applsci 15 07665 g005
Figure 6. XPS spectra and the deconvolutions for the LTA-Pt sample for (a) Ti 2p, (b) O 1s, (c) La 3d, (d) Pt 4f, and (e) C 1s.
Figure 6. XPS spectra and the deconvolutions for the LTA-Pt sample for (a) Ti 2p, (b) O 1s, (c) La 3d, (d) Pt 4f, and (e) C 1s.
Applsci 15 07665 g006
Figure 7. SEM images of the nanopowders: (a) LTA and (b) LTA-Pt.
Figure 7. SEM images of the nanopowders: (a) LTA and (b) LTA-Pt.
Applsci 15 07665 g007
Figure 8. TEM images of the nanopowders: (a) LTA and (b) LTA-Pt.
Figure 8. TEM images of the nanopowders: (a) LTA and (b) LTA-Pt.
Applsci 15 07665 g008
Figure 9. Platinum particles across the surface of the particle agglomerates in sample LTA-Pt.
Figure 9. Platinum particles across the surface of the particle agglomerates in sample LTA-Pt.
Applsci 15 07665 g009
Figure 10. Particle size distribution histograms for (a) LTA sample and (b) LTA-Pt sample.
Figure 10. Particle size distribution histograms for (a) LTA sample and (b) LTA-Pt sample.
Applsci 15 07665 g010
Figure 11. XRD patterns of the powders thermally treated at 900 °C for 2 h: (a) LTA and (b) LTA-Pt.
Figure 11. XRD patterns of the powders thermally treated at 900 °C for 2 h: (a) LTA and (b) LTA-Pt.
Applsci 15 07665 g011
Figure 12. N2 adsorption–desorption isotherms and pore size distributions (insert in the figures) of (a) LTA and (b) LTA-Pt samples.
Figure 12. N2 adsorption–desorption isotherms and pore size distributions (insert in the figures) of (a) LTA and (b) LTA-Pt samples.
Applsci 15 07665 g012
Figure 13. UV–Vis absorption spectra of samples.
Figure 13. UV–Vis absorption spectra of samples.
Applsci 15 07665 g013
Figure 14. Photogeneration of superoxide anion radical (·O2) under AM1.5 light exposure over the powders: (a) the time-course of the XTT-Formazan presence for the system containing LTA sample, (b) the time-course of the XTT-Formazan presence for the system containing LTA-Pt sample and (c) the relative amount of ·O2 produced by the powders during the irradiation time (120 min).
Figure 14. Photogeneration of superoxide anion radical (·O2) under AM1.5 light exposure over the powders: (a) the time-course of the XTT-Formazan presence for the system containing LTA sample, (b) the time-course of the XTT-Formazan presence for the system containing LTA-Pt sample and (c) the relative amount of ·O2 produced by the powders during the irradiation time (120 min).
Applsci 15 07665 g014
Figure 15. Comparative photoluminescence spectra of LTA and LTA-Pt samples for λexc = 290 nm in water containing ethanol (3 mL water and 30 μL ethanol).
Figure 15. Comparative photoluminescence spectra of LTA and LTA-Pt samples for λexc = 290 nm in water containing ethanol (3 mL water and 30 μL ethanol).
Applsci 15 07665 g015
Figure 16. The experimental setup used for the photodegradation of ethanol in the gaseous phase under simulated solar light.
Figure 16. The experimental setup used for the photodegradation of ethanol in the gaseous phase under simulated solar light.
Applsci 15 07665 g016
Figure 17. The results obtained for the gaseous phase oxidative degradation of ethanol using La2Ti2O7 catalysts exposed to simulated solar light: (a) amount of CH3CH2OH photo-degradated in time, (b) conversion of CH3CH2OH, (c) amount of CH3CHO produced in time, (d) selectivity to CH3CHO, (e) amount of HCOOH produced in time, (f) selectivity to HCOOH, (g) amount of CO2 produced in time and (h) selectivity to CO2.
Figure 17. The results obtained for the gaseous phase oxidative degradation of ethanol using La2Ti2O7 catalysts exposed to simulated solar light: (a) amount of CH3CH2OH photo-degradated in time, (b) conversion of CH3CH2OH, (c) amount of CH3CHO produced in time, (d) selectivity to CH3CHO, (e) amount of HCOOH produced in time, (f) selectivity to HCOOH, (g) amount of CO2 produced in time and (h) selectivity to CO2.
Applsci 15 07665 g017
Figure 18. Evolution of H2 from oxidative degradation of ethanol on synthesized materials exposed to simulated sunlight.
Figure 18. Evolution of H2 from oxidative degradation of ethanol on synthesized materials exposed to simulated sunlight.
Applsci 15 07665 g018
Figure 19. The reaction rate constant “k” (measured in min−1).
Figure 19. The reaction rate constant “k” (measured in min−1).
Applsci 15 07665 g019
Table 1. Assignment of vibration bands in FT-IR spectra of thermally treated samples.
Table 1. Assignment of vibration bands in FT-IR spectra of thermally treated samples.
Wavenumbers (cm−1)Assignments and Vibration Mode
LTALTA-Pt
806806Ti–O–Ti bridges in TiO6 octahedra from titanate
776772Ti–O stretching in TiO6 octahedra from titanate
620621La–O stretching and Ti–O vibrations
547549La–O stretching and Ti–O vibrations
491490La–O stretching vibrations and Ti–O–Ti bridges in TiO6 octahedra
464464La–O stretching vibrations and Ti–O bending modes
Table 2. Assignment of Raman shifts in thermally treated samples.
Table 2. Assignment of Raman shifts in thermally treated samples.
No.Synthesized SamplesLiterature for La–O and Ti–O Bonds [64,66]Assignment to Raman Vibrations
LTA-PtLTA
1109111112La–O stretching vibration
2129131131Ti–O bond stretching vibration for the six-fold coordinated Ti in rutile TiO2 (B1g mode)
3153154151La–O stretching vibration
4179181180La–O stretching vibration
5213213214La–O stretching vibration
6228227227La–O stretching vibration
7237240238La–O stretching vibration
8260262261La–O stretching vibration
9275274275La–O stretching vibration
10284285284La–O stretching vibration
11317316317La–O stretching vibration
12337342336La–O stretching vibration
13348352350La–O stretching vibration
14367367367La–O stretching vibration
15401402400La–O stretching vibration
16424426425La–O stretching vibration
17445446446Stretching vibrations of Ti–O–Ti octahedra (Eg mode)
18-488--
19517517517-
20536538533-
21555557556La–O stretching vibration
22605605606Stretching vibrations of Ti–O–Ti octahedra (A1g mode)
23-668--
24794792790Stretching vibrations of Ti–O–Ti octahedra
25809808808Ti–O bond stretching vibration for the six-fold coordinated Ti in rutile TiO2 (B2g mode)
Table 3. Compositional table—Binding energies with the attributions of the different components with each component’s relative intensity and percentage of total integral intensity obtained from XPS data and deconvolutions.
Table 3. Compositional table—Binding energies with the attributions of the different components with each component’s relative intensity and percentage of total integral intensity obtained from XPS data and deconvolutions.
SampleElement TypeElement ComponentRelative IntensityBinding EnergyAttributionAtomic Percentage
(%)(eV) (%)
LTAC 1s162.2284.6C–C 7.2
229.5286.27C=O
38.3289.07O–C=O, La carbonates
O 1s145.8529.34TiO2 (surf.)/La–O–Ti53.8
226.7530.72TiOx,TiO2 bulk/TiO2
323.3531.94C=O, OH
44.2533.48C-O
La 3d12.4833.82La3+ multiplet peak 125.0
249.9836.46antibonding
325.2840.44La3+ multiplet peak 2
422.5847.5satellite
Ti 2p166.3458.06TiO2 (surf.)/La–O–Ti14.0
233.7459.99TiOx, TiO2-bulk/TiO2
LTA-PtC 1s176.2284.6C–C9.6
217.4285.84C–O
36.4288.34O–C=O, La carbonates
O 1s167.6529.95TiO2/La–O–Ti67
227.3531.73C=O, OH
35.1532.8C–O, Pt oxide
La 3d131.1833.78La3+ multiplet peak 112.4
220.2835.48antibonding
334.8838.31La3+ multiplet peak 1
413.9847.90satellite
Ti 2p1 458.32TiO2/La–O–Ti9.8
Pt 4f177.470.36Pt1.2
222.671.55PtO
Table 4. Elemental composition in the weight and atomic percentage of the samples.
Table 4. Elemental composition in the weight and atomic percentage of the samples.
LTALTA-Pt
ElementWeight (%)Atomic (%)ElementWeight (%)Atomic (%)
OK13.2446.65OK14.2950.24
TiK23.5127.68TiK19.8023.27
LaL63.2425.67LaL64.0725.96
PtL1.840.53
Table 5. Phase identification, lattice parameters, and crystallite sizes of the thermally treated samples.
Table 5. Phase identification, lattice parameters, and crystallite sizes of the thermally treated samples.
SampleCrystalline PhaseLattice Parameters (Å)Crystallite Size (nm)
abc
LTALa2Ti2O77.81613.0225.53914
LTA-PtLa2Ti2O77.82913.0345.55315
Table 6. Elemental composition of the analyzed samples.
Table 6. Elemental composition of the analyzed samples.
SampleCompositionValues (Mass %)Line
LTA-PtTi26.46Ti-Kα
La72.50La-Lα
Pt1.03Pt-Lα
Table 7. The specific surface areas (SBET), total pore volumes (Vtotal), and average pore diameters (dBJH) of the samples.
Table 7. The specific surface areas (SBET), total pore volumes (Vtotal), and average pore diameters (dBJH) of the samples.
SampleSBET (m2/g)Vtotal (cm3/g)dBJH (nm)
LTA10.10.10629.6
LTA-Pt8.80.08730.2
Table 8. Ethanol degradation in gaseous phase measured after 3 h of reaction.
Table 8. Ethanol degradation in gaseous phase measured after 3 h of reaction.
CatalystCIN (μmoles)COUT (μmoles)Conversion of CH3CH2OH (%)
Ethanol
(CH3CH2OH)
Ethanol
(CH3CH2OH)
Acetaldehyde
(CH3CHO)
Formic Acid
(HCOOH)
Carbon Dioxide
(CO2)
LTA243.15208.3911.180.295.1314.29
LTA-Pt243.45171.0236.471.0613.6729.75
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

Ilie, A.; Predoană, L.; Anastasescu, C.; Preda, S.; Hosu, I.S.; Costescu, R.M.; Culiță, D.C.; Brătan, V.; Balint, I.; Zaharescu, M. Layered Perovskite La2Ti2O7 Obtained by Sol–Gel Method with Photocatalytic Activity. Appl. Sci. 2025, 15, 7665. https://doi.org/10.3390/app15147665

AMA Style

Ilie A, Predoană L, Anastasescu C, Preda S, Hosu IS, Costescu RM, Culiță DC, Brătan V, Balint I, Zaharescu M. Layered Perovskite La2Ti2O7 Obtained by Sol–Gel Method with Photocatalytic Activity. Applied Sciences. 2025; 15(14):7665. https://doi.org/10.3390/app15147665

Chicago/Turabian Style

Ilie, Alexandra, Luminița Predoană, Crina Anastasescu, Silviu Preda, Ioana Silvia Hosu, Ruxandra M. Costescu, Daniela C. Culiță, Veronica Brătan, Ioan Balint, and Maria Zaharescu. 2025. "Layered Perovskite La2Ti2O7 Obtained by Sol–Gel Method with Photocatalytic Activity" Applied Sciences 15, no. 14: 7665. https://doi.org/10.3390/app15147665

APA Style

Ilie, A., Predoană, L., Anastasescu, C., Preda, S., Hosu, I. S., Costescu, R. M., Culiță, D. C., Brătan, V., Balint, I., & Zaharescu, M. (2025). Layered Perovskite La2Ti2O7 Obtained by Sol–Gel Method with Photocatalytic Activity. Applied Sciences, 15(14), 7665. https://doi.org/10.3390/app15147665

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