Next Article in Journal
Hysteresis in Heat Capacity of MWCNTs Caused by Interface Behavior
Next Article in Special Issue
Facilitated Photocatalytic Degradation of Rhodamine B over One-Step Synthesized Honeycomb-Like BiFeO3/g-C3N4 Catalyst
Previous Article in Journal
Nanopore Detection Assisted DNA Information Processing
Previous Article in Special Issue
Ag3PO4-Deposited TiO2@Ti3C2 Petals for Highly Efficient Photodecomposition of Various Organic Dyes under Solar Light
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Effect of La3+ on the Methylene Blue Dye Removal Capacity of the La/ZnTiO3 Photocatalyst, a DFT Study

by
Ximena Jaramillo-Fierro
1,*,
Guisella Cuenca
2 and
John Ramón
2
1
Departamento de Química, Facultad de Ciencias Exactas y Naturales, Universidad Técnica Particular de Loja, San Cayetano Alto, Loja 1101608, Ecuador
2
Ingeniería Química, Facultad de Ciencias Exactas y Naturales, Universidad Técnica Particular de Loja, San Cayetano Alto, Loja 1101608, Ecuador
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(18), 3137; https://doi.org/10.3390/nano12183137
Submission received: 18 July 2022 / Revised: 30 August 2022 / Accepted: 7 September 2022 / Published: 10 September 2022

Abstract

:
Theoretically, lanthanum can bond with surface oxygens of ZnTiO3 to form La-O-Ti bonds, resulting in the change of both the band structure and the electron state of the surface. To verify this statement, DFT calculations were performed using a model with a dispersed lanthanum atom on the surface (101) of ZnTiO3. The negative heat segmentation values obtained suggest that the incorporation of La on the surface of ZnTiO3 is thermodynamically stable. The bandgap energy value of La/ZnTiO3 (2.92 eV) was lower than that of ZnTiO3 (3.16 eV). TDOS showed that the conduction band (CB) and the valence band (VB) energy levels of La/ZnTiO3 are denser than those of ZnTiO3 due to the participation of hybrid levels composed mainly of O2p and La5d orbitals. From the PDOSs, Bader’s charge analysis, and ELF function, it was established that the La-O bond is polar covalent. MB adsorption on La/ZnTiO3 (−200 kJ/mol) was more favorable than on ZnTiO3 (−85 kJ/mol). From the evidence of this study, it is proposed that the MB molecule first is adsorbed on the surface of La/ZnTiO3, and then the electrons in the VB of La/ZnTiO3 are photoexcited to hybrid levels, and finally, the MB molecule oxidizes into smaller molecules.

1. Introduction

Among the various industries, the textile industry is characterized by the generation of significant amounts of wastewater, in which dyes are some of the main components [1,2]. The dye methylene blue (MB) is widely used in textile products, making it one of the most common water contaminants [3]. Although many strategies, such as filtration and adsorption, have been proposed for the decontamination of aqueous systems, an economical and environmentally friendly procedure to purify large amounts of wastewater is still required [4,5]. Today, advanced oxidation processes (AOPs), including the Fenton method, Fenton-like method, and photoelectric catalytic processes, have gradually emerged due to their efficiency in removing organic contaminants [6]. Photocatalysis is also an AOP; therefore, it represents an interesting alternative for the degradation of organic pollutants, taking advantage of highly reactive substances as well as a light source [7,8]. In fact, the use of photocatalytic materials that have the capacity to produce reactive oxygen species (ROS) in the presence of sunlight could be one of the most accessible alternatives, since the sun is a suitable source of energy to promote chemical reactions without cost [9].
Technological development has allowed the design of various semiconductors with potential application in photocatalytic processes, including metallic compounds, such as oxides, chalcogenides, nitrides, and oxyhalides, among others. Photocatalytic semiconductors absorb photons to generate electrons (e) and holes (h+), which are then required to promote oxidation–reduction chemical reactions. An efficient photocatalyst should generally have the following attributes: appropriate bandgap energy, large light absorption range, and high charge migration capacity, as well as good stability, strong catalytic activity, and, from an economic view, high sustainability and low cost [10].
Zinc and titanium oxides are of great interest due to their ready availability, relatively low cost, minimal toxicity, good stability, and excellent ability to produce strongly oxidizing radical species [11,12,13,14]. Simultaneous preparation of the ZnO/TiO2 compound generally allows for the formation of three other well-known compounds, including perovskite ZnTiO3 (hexagonal, cubic), spinel Zn2TiO4 (cubic, tetragonal), and defect spinel Zn2Ti3O8 (cubic) [15,16,17]. Among them, the ZnTiO3 perovskite type has received much attention in photocatalytic reactions due to its relatively low cost, wide bandgap, compositional flexibility, unique chemical and physical properties such as high structural and thermal stability, and electron transport properties [18,19,20,21].
ZnTiO3 is a novel polar oxide containing Zn2+ (3d10) and Ti4+ (3d0), which have colombic repulsion with each other [22,23]; therefore, ZnTiO3 is a potential candidate as a ferroelectric, nonlinear optical, and piezoelectric material [24,25]. Furthermore, this semiconductor shows promise for the fabrication of dye-sensitized solar cells and for the treatment of harmful organic compounds by photocatalytic remediation [26,27,28,29,30].
Doping methodology can be a powerful approach to change the structural and catalytic properties of semiconductor oxides [31,32]. The dopant metal is often adapted to respond to visible light, so hybrid energy levels can form between the valence band (VB) and the conduction band of such oxides [33]. Therefore, proper doping can prevent photoinduced electron–hole (e/h+) pair recombination and improve the photoactivity of semiconductor oxides, including perovskite ZnTiO3 [34,35].
The photoactivity of semiconductors can be improved through the doping technique using different group of elements, such as metals (Au, Cu, Ag, Co, etc.) [36], non-metals (Mg, N, C, S, etc.) [37], and rare earth metals (Ce, Eu, La, etc.) [38]. Rare earth metals such as lanthanide ions are potential candidates for doping since they have 4f levels in their electronic configuration, which could act as effective reservoirs of electrons to trap them in the conduction band (CB) of the photocatalyst [39,40]. In addition, doping with lanthanide elements can improve light sensitivity and facilitate the transport and diffusion of reactive species and products by increasing the adsorption capacity of the adsorbate [40].
Several experimental and theoretical investigations have shown that the optical and electronic properties of materials can be improved by incorporating rare earth (RE) dopant ions [41], especially the lanthanum ion, due to its electron-trapping effect provided by the adaptable chemical valence sites (La2+ and La3+). Lanthanum (La) is one of the rare earth metal elements widely investigated for various purposes, due to its attractive properties and natural abundance (100%) [42]. Lanthanum is an effective dopant for modulating the band structure and enhancing the photocatalytic properties of various semiconductor oxides, including ZnTiO3 [43,44,45].
Various authors indicate that the La-doped semisconductor has completely different physical–chemical properties compared to the undoped semiconductor [46]. The presence of La in the ZnTiO3 lattice generates electron capture and therefore prevents their recombination with holes to limit the formation of electron–hole pairs (e/h+), which leads to an improvement in the photochemical efficiency of this semiconductor [47]. Moreover, the presence of La in the ZnTiO3 lattice reduces its bandgap and changes its optical behavior to the visible region, which improves the photoactivity of the semiconductor under sunlight [48].
However, a higher adsorption capacity due to a higher concentration of lanthanum is not a condition for a higher photocatalytic activity, which may be limited by a lower separation rate of the photoinduced (e/h+) pairs. In fact, an excessively high concentration of lanthanum could increase the amount of oxygen vacancies on the surface, which can become photoinduced electron–hole (e/h+) pair recombination centers [49]. Since electron–hole separation has a more important protagonist than adsorption capacity in photocatalytic processes, several studies have reported that a concentration of 1–2% (w/w) of lanthanum in the lattice of a photocatalyst facilitates charge transfer between the VB or CB of the photocatalyst and the d/f levels of the La3+ ions that were incorporated into the respective lattice. An increase in lanthanum concentration would result in a high accumulation of nanoparticles, which can protect the photocatalyst surface from light absorption and lead to lower photocatalytic activity.
The literature suggests that the La3+ ion cannot replace the Ti4+ or Zn2+ ions of the lattice in the ZnTiO3 body, since the ionic radius of the La3+ ion (1.15 Å) is much larger than that of Ti4+ (0.68 Å) or the Zn2+ ion (0.74 Å) [44], and the charge of La3+ does not coincide with the charge of Ti4+ or Zn2+ [40,46]. However, the La3+ ion can be well dispersed on the surface of ZnTiO3 semiconductor particles, forming Ti-O-La bonds with ionic-covalent characteristics [50].
The dispersion of La3+ on the surface of ZnTiO3 can cause lattice distortion in the surface layer and the generation of defects that block the growth of the catalyst crystallites and increase the number of oxygen vacancies in the catalyst [47,49,51]. On the other hand, the great diameter of La3+ is useful for improving the specific surface area (SSA) of ZnTiO3, and since La3+ shows a strong electron-withdrawing effect, the presence of this ion could also contribute to the formation of Lewis acid sites [52], which would improve the adsorption of organic contaminants on the photocatalyst surface [40].
The experimental results obtained previously showed that the incorporation of La in the lattice of the coupled semiconductor ZnTiO3/TiO2 (ZTO) induced the narrowing of its bandgap and also promoted the photocatalytic activity of the coupled semiconductor, probably due to the ability of La3+ ions to catch electrons in the CB of the ZTO and consequently increase the lifetime of photogenerated charge carriers. In addition, the La-ZTO nanocomposite sample was found to have a higher SSA and smaller average crystallite size compared to the corresponding pure ZTO. The average crystallite size of the nanocomposite reduced with the lanthanum incorporation, probably because when La3+ ions occupy the regular sites of the ZTO lattice, they form structural defects that block crystallite growth. On the other hand, UV–Vis diffuse reflection spectroscopy (UV–Vis XRD) studies showed that, compared to the pure ZTO, the absorption edge of the La-ZTO exhibits a slight redshift, which could be associated to the charge transition between the 4f electrons of lanthanum and the CB of the coupled semiconductor ZTO [53].
Despite these results, to date there is no systematic study at the molecular level of the effect of the La3+ ion on the electronic properties of the semiconductor La/ZnTiO3 nor of the synergistic effect of this ion on the adsorption capacity and photoactivity of the semiconductor for its application in the removal of methylene blue (MB) dye under solar light. Density Functional Theory (DFT) is a widely used computational method for calculating the electronic structure of an isolated molecule at the quantum level. DFT allows obtaining results with the desired chemical precision as long as a sufficiently large set of bases is used, the electronic correlation is adequately described, and relativistic effects are correctly included in the calculation [54]. Therefore, the objective of this study is to use computational calculations of the Density Functional Theory (DFT) to explain the changes produced in the electronic structure of ZnTiO3 due to doping with lanthanum and thus propose a mechanism for adsorption and photocatalytic degradation of the MB molecule on the surface (101) of both ZnTiO3 and La/ZnTiO3.
The results presented in this computational study are an original contribution to current knowledge, since they explain for the first time at the molecular level the effect of the lanthanum ion on the electronic properties of La/ZnTiO3, as well as the improvement of adsorption and photocatalytic properties of the La-doped catalyst for its potential use in removing MB dye under solar light.

2. Materials and Methods

Density Functional Theory (DFT) study was developed using the Vienna Ab Initio Simulation Package (VASP) version 6.0 (VASP Software GmbH, Vienna, Austria) [55,56]. For modeling and visualization of the corresponding molecules and structures, the molecular modeling program BioVia Materials Studio, version 5.5 (BioVia, San Diego, CA, USA) was utilized. The ionic potential of inner nuclei and electrons was described by pseudopotentials following projector augmented wave (PAW) approach [57]. All calculations were executed using the Perdew–Burke–Ernzerhof (PBE) generalized gradient approximation (GGA) functional to describe the electronic exchange-correlation interactions [58]. The plane wave cutoff energy was brought to a value of 500 eV. The Kohn–Sham equations [59] were solved in a self-consistent way allowing the variation in energy between cycles to be less than 10−5 eV.
The MB molecule adsorption on the surface of La/ZnTiO3 was modeled using the following previously optimized parameters: hexagonal ZnTiO3 with a cell = 5.15 Å × 5.15 Å × 13.94 Å <90° × 90° × 120°> [60]. Composite properties were estimated by sampling the first Brillouin zone by means of Monkhorst–Pack [61] k-point meshes of 3 × 2 × 1. All calculations were non-spin polarized. To improve the convergence of the total energy, the Gaussian smearing method was used with σ = 0.10 eV. All atomic positions relaxed completely until the respective forces were less than 0.001 eV/Å.
The valence electron configuration for the lanthanum (La) atom was 5s25p65d16s2. The functional GGA+U was also used in the calculation of the structures, since the DFT+U (GGA+U) approach has proven to be useful in correcting some of the deficiencies and also shows promise in studies of d-electron systems [62]. Hubbard U values were established at 2.5 eV and 6.0 eV for Ti and La atoms, respectively [26,44].
The bulk of ZnTiO3 was cleaved at the stable surface (101) [63,64,65] to study MB adsorption. The La/ZnTiO3 (101) slab model consisted of a p(2 × 3) supercell, with 36 Zn atoms, 36 Ti atoms, 108 O atoms, and 1 La atom. The values of the surface energies (γs) of the La/ZnTiO3 structure with a vacuum distance of 20 Å were estimated using the following equation [50,66]:
γ s = ( E s l a b n × E b u l k ) 2 A
where E s l a b is the total energy of the slab material (eV), E b u l k is the total energy of the bulk material (eV), n is the number of atoms involved in the slab, and A is the surface area (Å2).
On the other hand, the adsorption energy ( Δ E a d s ) of the MB molecule on the surface (101) of both ZnTiO3 and La/ZnTiO3 oxides was estimated using the following equation [44,67]:
Δ E a d s = E s o r b / s u r f E s u r f E s o r b
where E s o r b / s u r f is the energy of the supersystem produced by the adsorbed molecule on the surface (eV), E s u r f is the energy of the surface (eV), and E s o r b is the energy of the isolated molecule in vacuum (eV).
In this study, the adsorption of the La atom on the surface (101) of ZnTiO3 was investigated as follows. First, the clean surface of the semiconductor was subjected to the relaxation process. During this process, one layer of surface atoms was allowed to relax freely while the other atoms were immobilized. Then, a La atom was adsorbed on the surface (101) of ZnTiO3, which coordinated with the O atoms on the surface. The surface with a coordinated La atom also underwent relaxation. On the other hand, the adsorption of the methylene blue (MB) molecule on the surface (101) of La/ZnTiO3 was investigated by placing the MB molecule with different orientations. First, the P0 orientation was tested by placing the MB molecule in a vertical position, with the terminal methyl groups oriented to the coordinated lanthanum atom on the semiconductor surface. For the P1 and P3 orientations, the MB molecule was placed completely parallel, bringing the S and N heteroatom closer to the coordinated lanthanum atom on the semiconductor surface. In contrast, for the P2 and P4 orientations, the MB molecule was placed partially parallel, with the S or N heteroatom close to the coordinated La atom on the semiconductor surface, respectively.
Furthermore, to investigate the influence of La on molecular adsorption stability, values of heat of segregation ( Δ G s e g ) relaxation were calculated. The heat segregation ( Δ G s e g ) can be obtained by means of the following equation [44]:
Δ G s e g = 1 n ( E M B / o x i d e : n L a E M B / o x i d e n μ H e t + n μ L a )
where E M B / o x i d e : n L a and E M B / o x i d e are the total energies of the surfaces with and without La, and n is the number of the La atoms on the surface. μ is the chemical potential of the heteroatom (N or S) of the MB ring. In general, a more negative value of Δ G s e g is evidence that the surface is thermodynamically more stable.

3. Results

3.1. Optimization of La/ZnTiO3

The structure of ZnTiO3 with a subjacent hexagonal pattern with a rhombohedral-centered hexagonal Bravais lattice and R-3(148) space group [24,26] was optimized in a previous study [68]. The surface (101) of both ZnTiO3 and La/ZnTiO3 after relaxation is shown in Figure 1a and 1b, respectively.
As can be seen, the La atom formed bonds with three O atoms on the ZnTiO3 surface. The three surface O atoms moved toward the La atom, away from the adjacent Ti atoms just below, consequently increasing their initial bond lengths by about 0.09 Å to generate the new La-O bonds. The O-Ti bond lengths on the ZnTiO3 and La/ZnTiO3 surfaces are comparatively shown in Table S1.
La-O bond lengths ranged from 2.32 Å to 2.38 Å with an average of 2.35 Å and La-O-Ti angles ranged from 116.25° to 133.26° with an average of 122.18°, so there are no obvious symmetry elements inside the La polyhedron. The location of the shortest and longest La-O bonds in the structure confirms that it is not easy to adjust a sphere to the oxygen sites. Similar geometric distortions have been reported by other authors [69]. Significant bond lengths and angles on the surface of La/ZnTiO3 are shown in Table 1.
On the other hand, the surface energy of the surface (101) of the La/ZnTiO3 structure was obtained by Equation (1). The surface energy value (γs) of the surface (101) of La/ZnTiO3 with a vacuum distance of 20 Å was estimated to be 0.069 eV/Å2 (1.10 J/m2). This value is similar to the previously reported surface energy value for the surface (101) of ZnTiO3, which was 0.076 eV/Å2 (1.21 J/m2).
The adsorption energy (ΔEads) of the La atom on the surface (101) of ZnTiO3 oxide after relaxation was calculated using Equation (2), where E s o r b / s u r f is the energy of the supersystem generated by the adsorbed La atom on the ZnTiO3 surface (eV), Esurf is the energy of the clean oxide (eV), and Esorb is the energy of the isolated La atom in a vacuum (eV). The adsorption energy value of the La atom on the surface (101) of ZnTiO3 was calculated to be −852.46 kJ/mol. The adsorption of the La atom involved the formation of new chemical bonds, which suggests that a chemisorption process occurred [70].

3.2. Electronic Structure of La/ZnTiO3

The set of lines and points of high symmetry in the first Brillouin zone [71] and the results obtained from the estimation of the electronic band structure of La/ZnTiO3 are shown in Figure 2. In this figure, it is also shown that La/ZnTiO3 has two impurity energy levels (green lines) located just above the valence band maximum (VBM) of pure ZnTiO3. Furthermore, Figure 2 shows that the indirect bandgap energy value of the ZnTiO3 and La/ZnTiO3 structures calculated by the GGA+U method were 3.16 eV and 2.92 eV, respectively.
On the other hand, the density of states (DOSs) was determined to provide more information about the bonds present in the La/ZnTiO3 structure [22]. Figure 3 shows the total density of states (TDOSs) of La/ZnTiO3. The energy levels constituting the conduction band (CB) and the valence band (VB) are slightly denser than those previously reported for pure ZnTiO3, probably due to the participation of hybrid levels, composed mainly of O2p and La5d orbitals [44].
The TDOS of La/ZnTiO3 has two principal zones: a lower valence band (VB) zone, from about −18 to 0 eV, and an upper conduction band (CB) zone from about 2 to 6 eV. The VB is affected by the contribution of Zn and O, while the CB is affected by the contribution of Ti and La. In the La/ZnTiO3 structure, the valence band maximum (VBM) is bordered by the O atom, while the Ti atom establishes the conduction band maximum (CBM).
Furthermore, Figure 4a–d show the partial density of states (PDOSs) of La/ZnTiO3. From these figures, it is evident that in the energy range from −18 to −16 eV, the TDOS arises principally from the O2s orbital, with a small contribution from the Zn3d orbital and the Ti3p, Ti3d, and Ti4s orbitals. In the energy region from −6 to 0 eV, TDOS arises principally from the Zn3d, Ti3d, and O2p orbitals, and there are obvious hybridizations between the Zn3d–O2p and Ti3d–O2p orbitals. Above the Fermi level, in the energy region from 2 to 6 eV, the total DOS arises principally from the Ti3d, Zn3d, and O2p orbitals. These results agree with those informed in the literature [24]. Figure 4c,d show the PDOS of O and La atoms on the surface of La/ZnTiO3. Two types of hybridizations are evidenced from the PDOS results. First, the overlap of the La5d and O2p states is observed in the energy region from −6 to 0.1 eV. The wide range of states indicates that the La5d and O2p orbitals become delocalized, which could indicate ionic bonding. Second, there is also an overlap of La5p and O2s states in the range −18 to −16 eV and around −13 eV. These latter peaks are relatively narrow, suggesting that the La5p and O2s orbitals are both localized. The presence of localized electrons is a common characteristic of covalent bonding. The results show that the La-O bond is probably a combination of the ionic and covalent bonds (polar covalent bond). These results agree with those informed in the literature [50].
In order to verify the chemical features of the La-O bond described above, population analysis of La/ZnTiO3 was also performed using the Bader’s method [72]. This analysis is useful since the ionicity of a bond can be described in terms of charge transfer between the atoms that form the chemical bond [73,74]. In the La/ZnTiO3 structure, the net charge of Ti (+2.5e) was 1.5e less than its formal +4e charge, while the Zn atom had a positive charge of +1.3e and the O atom had a negative charge of −1.3e, being in both cases 0.7e less than their respective formal charges +2e and −2e. Finally, the net charge of La (+2.2e) was 0.8e lower than its formal charge of +3e. These results agree with those described in other studies [75]. The coordinates and the Bader’s charge analysis of the optimized ZnTiO3 and La/ZnTiO3 surfaces are provided in Table S2. According to the literature, a nonzero Bader charge transfer indicates a bond ionicity of zero; likewise, a very small charge transfer of about 0.05e between bonding atoms indicates a weak bond ionicity. Finally, a large charge transfer (e.g., 1.5e) indicates significant bond ionicity [24,73]. In this study, an increase in the magnitude of the net charge of the three surface oxygen atoms that gave rise to the new La-O bonds was observed, as well as a decrease in the magnitude of the net charge of the respective lanthanum atom. Therefore, in agreement with the literature, the results of the Bader’s analysis suggest the slight ionization of the La-O bond due to a possible charge transfer between the atoms that form the chemical bond [73,74].
Charge difference analysis was used to measure the charge redistribution on the ZnTiO3 surface due to lanthanum doping [73]. The results of the charge redistribution in the La/ZnTiO3 structure are shown in Figure 5. In this figure, the evident interaction between the three surface oxygen atoms of ZnTiO3 and the lanthanum atom suggests that the adhesion of La-O would be mainly influenced by the charge transfer between the La and the surface oxygens. The cyan and yellow surfaces correspond to the regions of charge gain and loss, respectively, which supports Bader’s analysis above.
Moreover, the electron localization function (ELF) was utilized to better understand the La-O chemical bond [76]. In ELF analysis, when the region of maximum density (RMD) is more symmetrically distributed around the nucleus, a more ionic or van der Waals interaction occurs. If the covalent character of a bond enhances, RMD migration between centers becomes more obvious until a completely symmetric geometry is achieved in the ideal covalent case. In Figure 6, the ELF section for the surface (101) of La/ZnTiO3 shows the interaction of the La atom with three O atoms on the surface. The figure shows that the RMDs are in the line that joins the nuclei and can be separated from the nuclei themselves by a path; in addition, the RMDs do not circumscribe the nucleus. Therefore, a polar covalence would be generated in the La-O bonds [77].

3.3. MB Adsorption on Surface (101) of ZnTiO3 and La/ZnTiO3

The tested orientations of the methylene blue (MB) molecule on the surface (101) of La/ZnTiO3 are shown in Figure 7. In this figure it is evident that the MB molecule is adsorbed when placed in the P1, P2, P3, and P4 orientations. However, no interaction was observed in the Po orientation where the MB molecule was placed in a vertical position.
To investigate the influence of La on molecular adsorption stability, adsorption values ( Δ E a d s ) after relaxation were calculated. The Δ E a d s of the MB molecule on the surface (101) of the La/ZnTiO3 oxide were calculated using Equation (2), where E s o r b / s u r f is the energy of the supersystem generated by the MB molecule on the oxide surface (eV), E s u r f is the energy of the clean oxide (eV), and E s o r b is the energy of the isolated MB molecule in a vacuum (eV). Likewise, heat segregation ( Δ G s e g ) was obtained by means of Equation (3). Table 2 lists the adsorption energy ( Δ E a d s ) and heat segregation ( Δ G s e g ) of the different interfaces after relaxation.
Since the molecular adsorption process of MB on the surface of La/ZnTiO3 with the molecule located in the P4 orientation was energetically more favored than in the other orientations, we studied the molecular adsorption process of the MB molecule on the surface of ZnTiO3 only with the P4 orientation. As shown in Figure 8, the MB molecule is more strongly adsorbed on the La/ZnTiO3 surface than on the ZnTiO3 surface. The average distances from the H atoms of the MB molecule (HMB) to the surface plane of ZnTiO3 are HMB-O(oxide) = 2.27 Å and HMB-O(oxide) = 2.41 Å. The average distance from the nitrogen atom of the MB molecule (NMB) to the lanthanum atom on the surface plane of La/ZnTiO3 is NMB-La(oxide) = 2.56 Å.
The adsorption energy value of the MB molecule on the clean surface was −64.06 kJ/mol. With the addition of La on the semiconductor surface, the adsorption energy was more stable for the adsorption of the MB molecule through the N-heteroatom of its aromatic ring (around −200 kJ/mol) compared to the adsorption of the MB molecule through the S heteroatom of its aromatic ring (about −85 kJ/mol). At the same time, the heat segregation values were around −58.19 eV and −57.00 eV for MB adsorption on the La/ZnTiO3 surface through the N and S heteroatoms of the MB ring, respectively. Negative values suggest that the incorporation of La on the semiconductor surface is thermodynamically stable. Therefore, it can be assumed that the presence of La increases the surface binding strength with more stable adsorption energy. Furthermore, the results imply that the P3 and P4 orientations were energetically more stable compared to the P1 and P2 orientations.
Bader’s charge analysis was also used to semiquantitatively assess charge transfer. The results of this analysis for the certain atoms at the surface of the semiconductors are listed in Table 3.
From the Bader’s charge analysis, it is evident that there was charge transfer between the MB molecule and neighboring atoms on the ZnTiO3 and La/ZnTiO3 surfaces. Furthermore, it was shown that the direction of charge transfer for the ZnTiO3 surface is from the MB molecule towards the catalyst, while for the La/ZnTiO3 surface the direction is the opposite. Likewise, Table 3 shows that there are areas of charge depletion around the H and La atoms and areas of charge accumulation around the O and N atoms, which suggests that the molecular adsorption of MB on the surfaces of ZnTiO3 and La/ZnTiO3 occurs through the formation of H-O hydrogen bonds and N-La ionic bonds, respectively.

4. Discussion

4.1. Optimization of La/ZnTiO3

Theoretically, lanthanum (La), due to its ionic radius, is difficult to intercalate in the ZnTiO3 lattice, but it can bond with the oxygen of the surface lattice to generate the La-O-Ti bond. In a previous experimental study, we found that La dispersion on the ZnTiO3 surface contributed to the decrease of X-ray diffraction (XRD) peak intensities [78]. This was probably due to the generation of La-O-Ti bonds on the surface of ZnTiO3 crystallites, which inhibited their growth by limiting direct contact with adjacent crystallites [79].
In this theoretical study, we use a model with a La atom (~2% w/w) intercalated only in the upper layer of ZnTiO3 to clarify the dispersion mechanism of La on the semiconductor surface. In Figure 1, where the optimized surfaces (101) of ZnTiO3 and La/ZnTiO3 are shown comparatively, the generation of La-O-Ti bonds on the surface of ZnTiO3 can be seen, which supports the experimental results obtained previously. The generation of La-O-Ti bonds not only allows the stabilization of small crystalline particles [79], but it can also generate the change in the structure of bands and states of the surface electrons. To also verify this argument, the DFT calculation was carried out using the optimized La/ZnTiO3 model.

4.2. Electronic Structure of La/ZnTiO3

To better understand the effect of lanthanum doping on the electronic structure of ZnTiO3, the electronic band structure diagrams of this semiconductor in both its pure and La-doped forms were plotted. In Figure 2, we can see that the energy levels of ZnTiO3 are less dense than those shown for La/ZnTiO3, probably because the latter presents hybrid levels or levels of impurities due to the presence of lanthanum. This figure shows that La/ZnTiO3 has two impurity energy levels located just above the valence band maximum (VBM) of pure ZnTiO3. Likewise, in Figure 2, another evident characteristic of La/ZnTiO3 is that the Fermi level (EF) crosses the impurity energy levels mentioned above, which, according to the literature, would mean that the EF is not completely full of electrons in the ground state. The hybrid levels or levels of impurities would be forming a surface acceptor in the La/ZnTiO3 due to the continuous state that is generated by the smaller distance and overlap between the energy levels of the impurities and the VBM. In this way, the surface acceptor, acting as a trap for photoexcited holes, could allow electrons in the VB to be first excited to isolated impurity energy levels and then excited to the CB. Throughout this photochemical process, the electrons would need less photon energy to be excited, so doping with La would allow the photocatalyst to obtain a better response to visible light. If the energy levels of the impurities are taken into account, the bandgap would be greatly reduced, and the acceptor effect would be more evident as the concentration of lanthanum increases. However, several authors have shown the efficacy of using La ≤ 1–2 wt.% to dope semiconductors and improve their photocatalytic activity in the UV and visible light regions [80]. This is probably due to the fact that the use of high concentrations of lanthanum as a dopant could increase the amount of oxygen vacancies and consequently increase the recombination centers of photoinduced electrons and holes; in addition, the high agglomeration of dopant particles could block the active sites on the catalyst surface and decrease its photoactivity [81,82].
To determine the indirect bandgap value of La/ZnTiO3, a Hubbard approximation term was implemented in order to accurately explain the electronic structure of this semiconductor [47]. In a previous study, we reported an indirect bandgap value of 3.16 eV for the ZnTiO3 structure. In this study, the indirect bandgap value for the La/ZnTiO3 structure was estimated to be 2.92 eV. As can be seen in Figure 2, the bandgap of the semiconductor decreases when La disperses on its surface, and this theoretical result supports the experimental results that we reported previously. The bandgap plays a fundamental role in the photocatalytic activity of semiconductors because it participates in the determination of the e-/h+ recombination rate [83]. Therefore, La/ZnTiO3 could be more photoactive than ZnTiO3 due to the smaller separation between the occupied and unoccupied bands [34]. Table 4 shows the comparison of the bandgap energy values of ZnTiO3 and La/ZnTiO3 calculated in this study with other experimental energy values reported in the literature.
Regarding the electronic nature of the La/ZnTiO3 structure, Figure 3 and Figure 4 show that the hybridization of Zn-O bonds and Ti-O bonds occurs principally in their 3d and 2p orbitals, while the hybridization of La-O bonds occurs mainly in their 5d and 2p orbitals. The charge distribution of the Ti-O, La-O, and Zn-O bonds indicates that the empty 3d orbital in Ti4+ (3d0) and the nearly empty 5d orbital in La3+ (5d1) more easily generate a covalent bond with O atoms than the completely occupied 3d orbital in Zn2+ (3d10).
On the other hand, Figure 5 shows the respective contour plots of the charge density difference of ZnTiO3 and La/ZnTiO3. At the free surface, the electronic structure changes due to the existence of the La atom. ELF analysis allowed a better description of the La-O chemical bond. According to the literature, the areas of charge depletion around the La atom and the areas of charge accumulation around the three O atoms could suggest ionic bonds between the La and O atoms [50]. However, the spherical shells around the nuclei clearly showed that these atoms were separated but still polarized, suggesting a polar covalent bond [76].

4.3. MB Adsorption on Surface (101) of ZnTiO3 and La/ZnTiO3

Numerous experimental studies of methylene blue removal from aqueous solutions have confirmed that this dye can be adsorbed on ZnTiO3 surfaces without difficulty, due to the electrostatic attraction between the surface oxygen atoms of the semiconductor and the positive regions of the MB molecule. The results reported in a previous study demonstrated that the MB molecule oriented perpendicular to the ZnTiO3 surface was the most favored (Eads = −282 kJ/mol). In fact, the MB molecule with orientation completely parallel to the surface was slightly bent away from it due to electrostatic repulsion. According to the optimized configurations, the adsorption of MB on ZnTiO3 occurred in a bidentate chelating mode through the formation of hydrogen bonds, promoting a highly stable adsorbate–surface interaction [68]. In contrast, the results of this study show that the MB molecule with partially parallel orientation and with the N heteroatom close to La dispersed on the ZnTiO3 surface and adsorbed with higher negative energy (Eads = −201.5 kJ/mol) than in the other orientations. Although theoretically the MB molecular adsorption process on the surface (101) of ZnTiO3 is energetically more favorable than on the (101) surface of La/ZnTiO3, the experimental evidence that we previously reported suggests that the dispersion of La on the semiconductor surface is an important requirement to continue with the next photocatalytic processes.
In the literature, insufficient computational studies were found about the molecular adsorption of MB on lanthanum dispersed on semiconductors; therefore, in Table 5, the results achieved in this investigation are contrasted with those results reported in the literature for the molecular adsorption of several dyes on ZnTiO3, which were calculated by the GGA/PBE method.

Proposed Photocatalytic Mechanism

Current investigation has found that diverse counterions can facilitate the dispersion of active sites on the surface of a catalyst. This dispersion increases as the diameter of the counterions increases. La3+ is a large diameter counterion and is therefore useful for increasing the SSA of the ZnTiO3 catalyst and for dispersing its active sites. La3+ has an important electron-withdrawing effect, which helps to generate Lewis acid centers, which provide the material catalytic stability in an aqueous reaction medium [84]. Consequently, the modification of La3+ can effectively modify both the SSA as the type of active acid center of a catalyst [85].
In this study, the use of La/ZnTiO3 allowed an effective degradation of the methylene blue solution, probably for the following reasons. First, the higher SSA of La/ZnTiO3 compared to ZnTiO3 would increase the adsorption capacity of the catalyst and provide a more active adsorption center for the target molecule. Second, the dispersion of La3+ on the surface of ZnTiO3 could enhance the transfer of photoinduced electrons from the bulk to the surface and thus prevent recombination of (e/h+) pairs under the radiation effect [38].
Since it has been suggested that the existence of La ions on the surface of ZnTiO3 influences the photocatalytic activity of this semiconductor by altering the recombination rate of the (e/h+) pair, the photocatalytic process in which La/ZnTiO3 participates could begin when the electrons (e) of the photocatalyst are photoexcited and immediately transferred from the valence band (VB) to the conduction band (CB), leaving a hole (h+) in the VB (reaction R1) due to the formation of a pair (e/h+). These (e/h+) pairs have the ability to recombine immediately (reaction R2); some of these pairs can even migrate to the semiconductor surface and react separately with species that are adsorbed on the surface, such as H2O, OH, O2, and other molecules (R), including the MB dye. The holes generated in the VB of the semiconductor are capable of oxidizing adsorbed water molecules or hydroxyl ions to generate strongly reactive hydroxyl radicals (reactions R3 and R4). The lifetime of free (h+) could be prolonged by allowing more of these holes to diffuse to the ZnTiO3 surface to create more reactive radicals that allow oxidizing adsorbed molecules on the surface. On the other hand, a La3+ ion with completely empty 4f and 5d orbitals can confine photoexcited electrons in the CB of ZnTiO3 (reaction R5); therefore, the La3+ ion becomes a La2+ ion by capturing an (e), and considering that the generated La2+ ion is unstable, the captured electrons could be transferred to oxygenated molecules adsorbed on the surface of ZnTiO3, generating hydroxyl radicals (OH˙) and superoxide radical anions (O2˙) through a sequence of reactions (reactions R6–R9). It is important to mention that both OH˙ and O2˙ radicals can easily oxidize organic molecules. In fact, O2˙- has a enough reduction potential to oxidize organic compounds with strong electron-donating groups; nevertheless, the OH˙ radical can remove H atoms or attack unsaturated C-C bonds [38]. Consequently, a model compound containing unsaturated C-C bonds, such as MB, is very likely to be photo-oxidized or attacked by the OH˙ radicals formed (reaction R10). Moreover, direct oxidation of the MB molecule could also occur if it reacts with the holes (reaction R11) [79]. The following reactions suggest the likely route for MB dye photodegradation on the La/ZnTiO3 surface [49]:
( ZnTiO 3 ) hv   ZnTiO 3 + e CB + h VB +
e CB + h VB + heat
H 2 O ads + h VB +     ( H + +   OH ) ads + h VB + OH ads
OH ads + h VB +   OH ads
La 3 + + e CB     La 2 +
La 2 + + ( O 2 ) ads     La 3 + +   O 2
O 2 +   H +   HO 2
2 HO 2   H 2 O 2 +   O 2
H 2 O 2 + e BC   OH +   OH
R +   OH ads   R ads +   H 2 O degradation   products
R ads + h VB +   R ads + degradation   products

5. Conclusions

ZnTiO3 is a semiconductor with a wide bandgap; therefore, the biggest problem with using ZnTiO3 for MB dye photodegradation lies in the fact that it is only active under UV light radiation. The dispersion of lanthanum (La) on the surface of ZnTiO3 could address this problem, since it allows hybrid levels in the VB and the CB to form, decreases the bandgap width, changes surface electron states, reduces the potential energy for MB adsorption, and decreases the energy barriers for MB photodegradation under solar irradiation. The dispersion of La on the surface of ZnTiO3 also distorts the surface lattice of the semiconductor, restricting the growth of crystallites to obtain nanometer-sized particles with a higher specific surface area.
La atom could not be intercalated in ZnTiO3 body; however, it had the ability to bond with oxygen on the ZnTiO3 surface to form a La-O-Ti bond. Furthermore, the electronic structures showed that La-O bonds had both ionic and covalent characteristics; thus, they are polar covalent bonds. Likewise, the lower bandgap energy of La/ZnTiO3 (2.92 eV) compared to the bandgap energy of ZnTiO3 (3.16 eV) was verified.
Experimental evidence indicates that the La/ZnTiO3 photocatalyst exhibits better MB removal capacity from aqueous solutions than the ZnTiO3 photocatalyst. In this study, the molecular adsorption process of MB on the surface of La/ZnTiO3 with the molecule located in the P4 orientation was energetically more favored than in the other orientations. In addition, the negative values of heat cleavage obtained in this study suggest that the incorporation of La on the surface of ZnTiO3 is thermodynamically stable. Therefore, based on the literature and evidence from this study, it is suggested that the enhanced removal of MB using La/ZnTiO3 would occur through the following mechanism. First, the MB molecule is adsorbed on the surface of La/ZnTiO3. This process can happen at room temperature without irradiation. Second, electrons in the VB of La/ZnTiO3 can be photoexcited to hybrid levels, mainly composed of La5d and O2p orbitals, through the band transition [44]. Likewise, the electrons in the CB can also be photoexcited to the catalyst surface through the intraband transition to finally oxidize the MB molecule.
In conclusion, the results of this theoretical study showed that the La/ZnTiO3 compound could effectively improve MB dye removal from aqueous solutions, which is consistent with previously obtained experimental results. The results presented in this article confirm the feasibility of using these photocatalysts as excellent dye adsorbents for future environmental applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12183137/s1, Table S1: Relevant interatomic distances for La/ZnTiO3 surface; Table S2: Bader’s charge analysis of the optimized ZnTiO3 and La/ZnTiO3 surfaces.

Author Contributions

Conceptualization, X.J.-F.; methodology, X.J.-F., G.C. and J.R.; software, X.J.-F.; validation, X.J.-F.; formal analysis, X.J.-F.; investigation, X.J.-F., G.C. and J.R.; resources, X.J-F.; data curation, X.J.-F.; writing—original draft preparation, X.J.-F.; writing—review and editing, X.J.-F.; supervision, X.J.-F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Universidad Técnica Particular de Loja, Ecuador.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Acknowledgments

The authors thank the computing time contributed by the Servidor de Cálculo of the Universidad Técnica Particular de Loja (Ecuador).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yaseen, D.A.; Scholz, M. Textile dye wastewater characteristics and constituents of synthetic effluents: A critical review. Int. J. Environ. Sci. Technol. 2019, 16, 1193–1226. [Google Scholar] [CrossRef]
  2. Al-Mamun, M.R.; Kader, S.; Islam, M.S.; Khan, M.Z.H. Photocatalytic activity improvement and application of UV-TiO2 photocatalysis in textile wastewater treatment: A review. J. Environ. Chem. Eng. 2019, 7, 103248. [Google Scholar] [CrossRef]
  3. Jaramillo-Fierro, X.; González, S.; Montesdeoca-Mendoza, F.; Medina, F. Structuring of ZnTiO3/TiO2 adsorbents for the removal of methylene blue, using zeolite precursor clays as natural additives. Nanomaterials 2021, 11, 898. [Google Scholar] [CrossRef]
  4. Wang, G.; Li, G.; Huan, Y.; Hao, C.; Chen, W. Acrylic acid functionalized graphene oxide: High-efficient removal of cationic dyes from wastewater and exploration on adsorption mechanism. Chemosphere 2020, 261, 127736. [Google Scholar] [CrossRef] [PubMed]
  5. Ambigadevi, J.; Kumar, P.S.; Vo, D.V.N.; Haran, S.H.; Raghavan, T.N.S. Recent developments in photocatalytic remediation of textile effluent using semiconductor based nanostructured catalyst: A review. J. Environ. Chem. Eng. 2021, 9, 104881. [Google Scholar] [CrossRef]
  6. Shi, X.; Wang, L.; Zuh, A.A.; Jia, Y.; Ding, F.; Cheng, H.; Wang, Q. Photo-Fenton reaction for the degradation of tetracycline hydrochloride using a FeWO4/BiOCl nanocomposite. J. Alloys Compd. 2022, 903, 163889. [Google Scholar] [CrossRef]
  7. Wang, L.; Ma, X.; Huang, G.; Lian, R.; Huang, J.; She, H.; Wang, Q. Construction of ternary CuO/CuFe2O4/g-C3N4 composite and its enhanced photocatalytic degradation of tetracycline hydrochloride with persulfate under simulated sunlight. J. Environ. Sci. 2022, 112, 59–70. [Google Scholar] [CrossRef]
  8. Zhu, S.; Wang, D. Photocatalysis: Basic Principles, Diverse Forms of Implementations and Emerging Scientific Opportunities. Adv. Energy Mater. 2017, 7, 1700841. [Google Scholar] [CrossRef]
  9. Nosaka, Y.; Nosaka, A.Y. Generation and Detection of Reactive Oxygen Species in Photocatalysis. Chem. Rev. 2017, 117, 11302–11336. [Google Scholar] [CrossRef]
  10. Manuel, A.; Shankar, K. Hot Electrons in TiO2–Noble Metal Nano-Heterojunctions: Fundamental Science and Applications in Photocatalysis. Nanomaterials 2021, 11, 1249. [Google Scholar] [CrossRef]
  11. Upadhyay, G.K.; Rajput, J.K.; Pathak, T.K.; Kumar, V.; Purohit, L.P. Synthesis of ZnO:TiO2 nanocomposites for photocatalyst application in visible light. Vacuum 2019, 160, 154–163. [Google Scholar] [CrossRef]
  12. Lee, C.G.; Na, K.H.; Kim, W.T.; Park, D.C.; Yang, W.H.; Choi, W.Y. TiO2/ZnO nanofibers prepared by electrospinning and their photocatalytic degradation of methylene blue compared with TiO2 nanofibers. Appl. Sci. 2019, 9, 3404. [Google Scholar] [CrossRef]
  13. Sinha, D.; De, D.; Goswami, D.; Mondal, A.; Ayaz, A. ZnO and TiO2 Nanostructured Dye sensitized Solar Photovoltaic Cell. Mater. Today Proc. 2019, 11, 782–788. [Google Scholar] [CrossRef]
  14. Zalani, N.M.; Kaleji, B.K.; Mazinani, B. Synthesis and characterisation of the mesoporous ZnO-TiO2 nanocomposite; Taguchi optimisation and photocatalytic methylene blue degradation under visible light. Mater. Technol. 2020, 35, 281–289. [Google Scholar] [CrossRef]
  15. Siwińska-Stefańska, K.; Kubiak, A.; Piasecki, A.; Goscianska, J.; Nowaczyk, G.; Jurga, S.; Jesionowski, T. TiO2-ZnO Binary Oxide Systems: Comprehensive Characterization and Tests of Photocatalytic Activity. Materials 2018, 11, 841. [Google Scholar] [CrossRef]
  16. Tahay, P.; Khani, Y.; Jabari, M.; Bahadoran, F.; Safari, N.; Zamanian, A. Synthesis of cubic and hexagonal ZnTiO3 as catalyst support in steam reforming of methanol: Study of physical and chemical properties of copper catalysts on the H2 and CO selectivity and coke formation. Int. J. Hydrogen Energy 2020, 45, 9484–9495. [Google Scholar] [CrossRef]
  17. Pantoja-Espinoza, J.C.; Domínguez-Arvizu, J.L.; Jiménez-Miramontes, J.A.; Hernández-Majalca, B.C.; Meléndez-Zaragoza, M.J.; Salinas-Gutiérrez, J.M.; Herrera-Pérez, G.M.; Collins-Martínez, V.H.; López-Ortiz, A. Comparative study of Zn2Ti3O8 and ZnTiO3 photocatalytic properties for hydrogen production. Catalysts 2020, 10, 1372. [Google Scholar] [CrossRef]
  18. Baamran, K.S.; Tahir, M. Thermodynamic investigation and experimental analysis on phenol steam reforming towards enhanced H2 production over structured Ni/ZnTiO3 nanocatalyst. Energy Convers. Manag. 2019, 180, 796–810. [Google Scholar] [CrossRef]
  19. Ranjith, K.S.; Uyar, T. ZnO-TiO2 composites and ternary ZnTiO3 electrospun nanofibers: The influence of annealing on the photocatalytic response and reusable functionality. CrystEngComm 2018, 20, 5801–5813. [Google Scholar] [CrossRef]
  20. Sowmyashree, A.S.; Somya, A.; Kumar, C.B.P.; Rao, S. Novel nano corrosion inhibitor, integrated zinc titanate nano particles: Synthesis, characterization, thermodynamic and electrochemical studies. Surf. Interfaces 2021, 22, 100812. [Google Scholar] [CrossRef]
  21. Edalatfar, M.; Yazdani, F.; Salehi, M.B. Synthesis and identification of ZnTiO3 nanoparticles as a rheology modifier additive in water-based drilling mud. J. Pet. Sci. Eng. 2021, 201, 108415. [Google Scholar] [CrossRef]
  22. Zhu, B.; Yang, Q.; Zhang, W.; Cui, S.; Yang, B.; Wang, Q.; Li, S.; Zhang, D. A high sensitivity dual-mode optical thermometry based on charge compensation in ZnTiO3:M (M = Eu3+, Mn4+) hexagonal prisms. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2022, 274, 121101. [Google Scholar] [CrossRef] [PubMed]
  23. Mofokeng, S.J.; Noto, L.L.; Dhlamini, M.S. Photoluminescence properties of ZnTiO3:Eu3+ phosphor with enhanced red emission by Al3+ charge compensation. J. Lumin. 2020, 228, 117569. [Google Scholar] [CrossRef]
  24. Zhang, J.; Xu, B.; Wang, Y.S.; Qin, Z.; Ke, S.H. First-principles investigation of the ferroelectric, piezoelectric and nonlinear optical properties of LiNbO3-type ZnTiO3. Sci. Rep. 2019, 9, 17632. [Google Scholar] [CrossRef]
  25. Djellabi, R.; Ordonez, M.F.; Conte, F.; Falletta, E.; Bianchi, C.L.; Rossetti, I. A Review of Advances in Multifunctional XTiO3 Perovskite-type Oxides as piezo-photocatalysts for Environmental Remediation and Energy Production. J. Hazard. Mater. 2021, 421, 126792. [Google Scholar] [CrossRef]
  26. Obodo, K.O.; Noto, L.L.; Mofokeng, S.J.; Ouma, C.N.M.; Braun, M.; Dhlamini, M.S. Influence of Tm, Ho and Er dopants on the properties of Yb activated ZnTiO3 perovskite: A density functional theory insight. Mater. Res. Express 2018, 5, 106202. [Google Scholar] [CrossRef]
  27. Sarkar, M.; Sarkar, S.; Biswas, A.; De, S.; Kumar, P.R.; Mothi, E.M.; Kathiravan, A. Zinc titanate nanomaterials—photocatalytic studies and sensitization of hydantoin derivatized porphyrin dye. Nano-Struct. Nano-Objects 2020, 21, 100412. [Google Scholar] [CrossRef]
  28. Kurajica, S.; Minga, I.; Blazic, R.; Muzina, K.; Tominac, P. Adsorption and Degradation Kinetics of Methylene Blue on As-prepared and Calcined Titanate Nanotubes. Athens J. Sci. 2018, 5, 7–22. [Google Scholar] [CrossRef]
  29. Sahu, A.; Chaurashiya, R.; Hiremath, K.; Dixit, A. Nanostructured zinc titanate wide band gap semiconductor as a photoelectrode material for quantum dot sensitized solar cells. Sol. Energy 2018, 163, 338–346. [Google Scholar] [CrossRef]
  30. Bhagwat, U.O.; Wu, J.J.; Asiri, A.M.; Anandan, S. Synthesis of ZnTiO3@TiO2 Heterostructure Nanomaterial as a Visible light Photocatalyst. Chem. Sel. 2019, 4, 6106–6112. [Google Scholar] [CrossRef]
  31. Rafieh, A.I.; Ekanayake, P.; Tan, A.L.; Lim, C.M. Effects of ionic radii of co-dopants (Mg, Ca, Al and La) in TiO2 on performance of dye-sensitized solar cells. Sol. Energy 2017, 141, 249–255. [Google Scholar] [CrossRef]
  32. Faisal, M.; Jalalah, M.; Harraz, F.A.; El-Toni, A.M.; Labis, J.P.; Al-Assiri, M.S. A novel Ag/PANI/ZnTiO3 ternary nanocomposite as a highly efficient visible-light-driven photocatalyst. Sep. Purif. Technol. 2021, 256, 117847. [Google Scholar] [CrossRef]
  33. Ozturk, B.; Soylu, G.S.P. Promoting role of transition metal oxide on ZnTiO3-TiO2 nanocomposites for the photocatalytic activity under solar light irradiation. Ceram. Int. 2016, 42, 11184–11192. [Google Scholar] [CrossRef]
  34. Khaki, M.R.D.; Shafeeyan, M.S.; Raman, A.A.A.; Daud, W.M.A.W. Enhanced UV–Visible photocatalytic activity of Cu-doped ZnO/TiO2 nanoparticles. J. Mater. Sci. Mater. Electron. 2018, 29, 5480–5495. [Google Scholar] [CrossRef]
  35. Li, X.; Xiong, J.; Huang, J.; Feng, Z.; Luo, J. Novel g-C3N4/h′ZnTiO3-a′TiO2 direct Z-scheme heterojunction with significantly enhanced visible-light photocatalytic activity. J. Alloys Compd. 2019, 774, 768–778. [Google Scholar] [CrossRef]
  36. Abirami, R.; Kalaiselvi, C.R.; Kungumadevi, L.; Senthil, T.S.; Kang, M. Synthesis and characterization of ZnTiO3 and Ag doped ZnTiO3 perovskite nanoparticles and their enhanced photocatalytic and antibacterial activity. J. Solid State Chem. 2020, 281, 121019. [Google Scholar] [CrossRef]
  37. Ruzimuradov, O.; Hojamberdiev, M.; Fasel, C.; Riedel, R. Fabrication of lanthanum and nitrogen—Co-doped SrTiO3– TiO2 heterostructured macroporous monolithic materials for photocatalytic degradation of organic dyes under visible light. J. Alloys Compd. 2017, 699, 144–150. [Google Scholar] [CrossRef]
  38. Chaker, H.; Ameur, N.; Saidi-Bendahou, K.; Djennas, M.; Fourmentin, S. Modeling and Box-Behnken design optimization of photocatalytic parameters for efficient removal of dye by lanthanum-doped mesoporous TiO2. J. Environ. Chem. Eng. 2021, 9, 2213–3437. [Google Scholar] [CrossRef]
  39. Mazierski, P.; Lisowski, W.; Grzyb, T.; Winiarski, M.J.; Klimczuk, T.; Mikołajczyk, A.; Flisikowski, J.; Hirsch, A.; Kołakowska, A.; Puzyn, T.; et al. Enhanced photocatalytic properties of lanthanide-TiO2 nanotubes: An experimental and theoretical study. Appl. Catal. B Environ. 2017, 205, 376–385. [Google Scholar] [CrossRef]
  40. Prakash, J.; Kumar, A.; Dai, H.; Janegitz, B.C.; Krishnan, V.; Swart, H.C.; Sun, S. Novel rare earth metal–doped one-dimensional TiO2 nanostructures: Fundamentals and multifunctional applications. Mater. Today Sustain. 2021, 13, 100066. [Google Scholar] [CrossRef]
  41. Jian, S.; Tian, Z.; Hu, J.; Zhang, K.; Zhang, L.; Duan, G.; Yang, W.; Jiang, S. Enhanced visible light photocatalytic efficiency of La-doped ZnO nanofibers via electrospinning-calcination technology. Adv. Powder Mater. 2022, 1, 100004. [Google Scholar] [CrossRef]
  42. Shwetharani, R.; Sakar, M.; Chandan, H.R.; Balakrishna, R.G. Observation of simultaneous photocatalytic degradation and hydrogen evolution on the lanthanum modified TiO2 nanostructures. Mater. Lett. 2018, 218, 262–265. [Google Scholar] [CrossRef]
  43. Dal’Toé, A.T.O.; Colpani, G.L.; Padoin, N.; Fiori, M.A.; Soares, C. Lanthanum doped titania decorated with silver plasmonic nanoparticles with enhanced photocatalytic activity under UV-visible light. Appl. Surf. Sci. 2018, 441, 1057–1071. [Google Scholar] [CrossRef]
  44. Yan, Z.; Yang, X.; Gao, G.; Gao, R.; Zhang, T.; Tian, M.; Su, H.; Wang, S. Understanding of Photocatalytic Partial Oxidation of Methanol to Methyl Formate on Surface Doped La(Ce)-Tio2: Experiment and Dft Calculation. SSRN Electron. J. 2022, 411, 31–40. [Google Scholar] [CrossRef]
  45. Song, K.; Min, T.; Seo, J.; Ryu, S.; Lee, H.; Wang, Z.; Choi, S.Y.; Lee, J.; Eom, C.B.; Oh, S.H. Electronic and Structural Transitions of LaAlO3/SrTiO3 Heterostructure Driven by Polar Field-Assisted Oxygen Vacancy Formation at the Surface. Adv. Sci. 2021, 8, 2002073. [Google Scholar] [CrossRef]
  46. Coelho, L.L.; Hotza, D.; Estrella, A.S.; de Amorim, S.M.; Puma, G.L.; Moreira, R.d.P.M. Modulating the photocatalytic activity of TiO2 (P25) with lanthanum and graphene oxide. J. Photochem. Photobiol. A Chem. 2019, 372, 1–10. [Google Scholar] [CrossRef]
  47. Priyanka, K.P.; Revathy, V.R.; Rosmin, P.; Thrivedu, B.; Elsa, K.M.; Nimmymol, J.; Balakrishna, K.M.; Varghese, T. Influence of La doping on structural and optical properties of TiO2 nanocrystals. Mater. Charact. 2016, 113, 144–151. [Google Scholar] [CrossRef]
  48. Wu, A.; Wang, D.; Wei, C.; Zhang, X.; Liu, Z.; Feng, P.; Ou, X.; Qiang, Y.; Garcia, H.; Niu, J. A comparative photocatalytic study of TiO2 loaded on three natural clays with different morphologies. Appl. Clay Sci. 2019, 183, 105352. [Google Scholar] [CrossRef]
  49. Eskandarloo, H.; Badiei, A.; Behnajady, M.A.; Tavakoli, A.; Ziarani, G.M. Ultrasonic-assisted synthesis of Ce doped cubic-hexagonal ZnTiO3 with highly efficient sonocatalytic activity. Ultrason. Sonochem. 2016, 29, 258–269. [Google Scholar] [CrossRef]
  50. Guo, W.; She, Z.; Yang, S.; Xue, H.; Zhang, X. Understanding the influence of Lu, La and Ga active elements on the bonding properties of Sn/SiO2 interfaces from first principle calculations. Ceram. Int. 2020, 46, 24737–24743. [Google Scholar] [CrossRef]
  51. Khan, S.; Cho, H.; Kim, D.; Han, S.S.; Lee, K.H.; Cho, S.H.; Song, T.; Choi, H. Defect engineering toward strong photocatalysis of Nb-doped anatase TiO2: Computational predictions and experimental verifications. Appl. Catal. B Environ. 2017, 206, 520–530. [Google Scholar] [CrossRef]
  52. Mazierski, P.; Mikolajczyk, A.; Bajorowicz, B.; Malankowska, A.; Zaleska-Medynska, A.; Nadolna, J. The role of lanthanides in TiO2-based photocatalysis: A review. Appl. Catal. B Environ. 2018, 233, 301–317. [Google Scholar] [CrossRef]
  53. Ako, R.T.; Ekanayake, P.; Tan, A.L.; Young, D.J. La modified TiO2 photoanode and its effect on DSSC performance: A comparative study of doping and surface treatment on deep and surface charge trapping. Mater. Chem. Phys. 2016, 172, 105–112. [Google Scholar] [CrossRef]
  54. Zhan, C.G. Development and application of first-principles electronic structure approach for molecules in solution based on fully polarizable continuum model. Wuli Huaxue Xuebao/Acta Physico. Chim. Sin. 2011, 27, 1–10. [Google Scholar] [CrossRef]
  55. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B Condens. Matter Mater. Phys. 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  56. Wang, V.; Xu, N.; Liu, J.C.; Tang, G.; Geng, W.T. VASPKIT: A user-friendly interface facilitating high-throughput computing and analysis using VASP code. Comput. Phys. Commun. 2021, 267, 108033. [Google Scholar] [CrossRef]
  57. Sujith, C.P.; Joseph, S.; Mathew, T.; Mathew, V. First-principles investigation of structural, electronic and optical properties of quasi-one-dimensional barium cadmium chalcogenides Ba2CdX3 (X = S, Se, Te) using HSE06 and GGA-PBE functionals. J. Phys. Chem. Solids 2022, 161, 110488. [Google Scholar] [CrossRef]
  58. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  59. Kohn, W.; Sham, L.J. Quantum density oscillations in an inhomogeneous electron gas. Phys. Rev. 1965, 137, A1697. [Google Scholar] [CrossRef]
  60. Jaramillo-Fierro, X.; Hernández, K.; González, S. Cu(C3H3N3S3)3 Adsorption onto ZnTiO3/TiO2 for Coordination-Complex Sensitized Photochemical Applications. Materials 2022, 15, 3252. [Google Scholar] [CrossRef]
  61. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  62. German, E.; Faccio, R.; Mombrú, A.W. Comparison of standard DFT and Hubbard-DFT methods in structural and electronic properties of TiO2 polymorphs and H-titanate ultrathin sheets for DSSC application. Appl. Surf. Sci. 2018, 428, 118–123. [Google Scholar] [CrossRef]
  63. Cherifi, K.; Cheknane, A.; Hilal, H.S.; Benghia, A.; Rahmoun, K.; Benyoucef, B. Investigation of triphenylamine-based sensitizer characteristics and adsorption behavior onto ZnTiO3 perovskite (1 0 1) surfaces for dye-sensitized solar cells using first-principle calculation. Chem. Phys. 2020, 530, 110595. [Google Scholar] [CrossRef]
  64. Nor, N.U.M.; Mazalan, E.; Risko, C.; Crocker, M.; Amin, N.A.S. Unveiling the structural, electronic, and optical effects of carbon-doping on multi-layer anatase TiO2 (1 0 1) and the impact on photocatalysis. Appl. Surf. Sci. 2022, 586, 152641. [Google Scholar] [CrossRef]
  65. Samanta, P.K.; English, N.J. Opto-electronic properties of stable blue photosensitisers on a TiO2 anatase-101 surface for efficient dye-sensitised solar cells. Chem. Phys. Lett. 2019, 731, 136624. [Google Scholar] [CrossRef]
  66. Chang, X.; Li, X.; Xue, Q. Sensing mechanism of acetone adsorption on charged ZnO and ZnSe surfaces: Insights from DFT calculations. Mater. Today Commun. 2022, 31, 103238. [Google Scholar] [CrossRef]
  67. Lai, W.; Zhang, K.; Shao, P.; Yang, L.; Ding, L.; Pavlostathis, S.G.; Shi, H.; Zou, L.; Liang, D.; Luo, X. Optimization of adsorption configuration by DFT calculation for design of adsorbent: A case study of palladium ion-imprinted polymers. J. Hazard. Mater. 2019, 379, 120791. [Google Scholar] [CrossRef]
  68. Jaramillo-Fierro, X.; Capa, L.F.; Medina, F.; González, S. Dft study of methylene blue adsorption on ZnTiO3 and TiO2 surfaces (101). Molecules 2021, 26, 3780. [Google Scholar] [CrossRef]
  69. Paterson, A.L.; Hanson, M.A.; Werner-Zwanziger, U.; Zwanziger, J.W. Relating 139La Quadrupolar Coupling Constants to Polyhedral Distortion in Crystalline Structures. J. Phys. Chem. C 2015, 119, 25508–25517. [Google Scholar] [CrossRef]
  70. Maldonado, F.; Villamagua, L.; Rivera, R. DFT Analysis of the Adsorption of Phenol on the Nonpolar (1010) ZnO Surface. J. Phys. Chem. C 2019, 123, 12296–12304. [Google Scholar] [CrossRef]
  71. Hinuma, Y.; Pizzi, G.; Kumagai, Y.; Oba, F.; Tanaka, I. Band structure diagram paths based on crystallography. Comput. Mater. Sci. 2017, 128, 140–184. [Google Scholar] [CrossRef]
  72. Yu, M.; Trinkle, D.R. Accurate and efficient algorithm for Bader charge integration. J. Chem. Phys. 2011, 134, 064111. [Google Scholar] [CrossRef] [PubMed]
  73. Zhang, H.; Huang, W.; Wang, W.C.; Shi, X.Q. Ionicity of bonding in elemental solids. J. Phys. Commun. 2018, 2, 115009. [Google Scholar] [CrossRef]
  74. Kumar, P.S.V.; Raghavendra, V.; Subramanian, V. Bader’s Theory of Atoms in Molecules (AIM) and its Applications to Chemical Bonding. J. Chem. Sci. 2016, 128, 1527–1536. [Google Scholar] [CrossRef]
  75. Koch, D.; Golub, P.; Manzhos, S. Stability of charges in titanium compounds and charge transfer to oxygen in titanium dioxide. J. Phys. Conf. Ser. 2018, 1136, 12017. [Google Scholar] [CrossRef]
  76. Savin, A.; Nesper, R.; Wengert, S.; Fässler, T.F. ELF: The Electron Localization Function. Angew. Chemie Int. Ed. Engl. 1997, 36, 1808–1832. [Google Scholar] [CrossRef]
  77. Wen, C.; Zhu, Y.J.; Kanbara, T.; Zhu, H.Z.; Xiao, C.F. Effects of I and F codoped TiO2 on the photocatalytic degradation of methylene blue. Desalination 2009, 249, 621–625. [Google Scholar] [CrossRef]
  78. Jaramillo-Fierro, X.; González, S.; Medina, F. La-doped ZnTiO3 /TiO2 nanocomposite supported on ecuadorian diatomaceous earth as a highly efficient photocatalyst driven by solar light. Molecules 2021, 26, 6232. [Google Scholar] [CrossRef]
  79. Wang, B.; Zhang, G.; Sun, Z.; Zheng, S.; Frost, R.L. A comparative study about the influence of metal ions (Ce, la and V) doping on the solar-light-induced photodegradation toward rhodamine B. J. Environ. Chem. Eng. 2015, 3, 1444–1451. [Google Scholar] [CrossRef]
  80. Armaković, S.J.; Grujić-Brojčin, M.; Šćepanović, M.; Armaković, S.; Golubović, A.; Babić, B.; Abramović, B.F. Efficiency of La-doped TiO2 calcined at different temperatures in photocatalytic degradation of β-blockers. Arab. J. Chem. 2019, 12, 5355–5369. [Google Scholar] [CrossRef]
  81. Surendar, T.; Kumar, S.; Shanker, V. Influence of La-doping on phase transformation and photocatalytic properties of ZnTiO3nanoparticles synthesized via modified sol-gel method. Phys. Chem. Chem. Phys. 2014, 16, 728–735. [Google Scholar] [CrossRef] [PubMed]
  82. Elsellami, L.; Lachheb, H.; Houas, A. Synthesis, characterization and photocatalytic activity of Li-, Cd-, and La-doped TiO2. Mater. Sci. Semicond. Process. 2015, 36, 103–114. [Google Scholar] [CrossRef]
  83. Cherifi, K.; Cheknane, A.; Benghia, A.; Hilal, H.S.; Rahmoun, K.; Benyoucef, B.; Goumri-Said, S. Exploring N3 ruthenium dye adsorption onto ZnTiO3 (101) and (110) surfaces for dye sensitized solar cell applications: Full computational study. Mater. Today Energy 2019, 13, 109–118. [Google Scholar] [CrossRef]
  84. Shu, Q.; Liu, X.; Huo, Y.; Tan, Y.; Zhang, C.; Zou, L. Construction of a Brönsted-Lewis solid acid catalyst La-PW-SiO2/SWCNTs based on electron withdrawing effect of La(III) on π bond of SWCNTs for biodiesel synthesis from esterification of oleic acid and methanol. Chin. J. Chem. Eng. 2022, 44, 351–362. [Google Scholar] [CrossRef]
  85. Kumar, V.V.; Naresh, G.; Sudhakar, M.; Tardio, J.; Bhargava, S.K.; Venugopal, A. Role of Brønsted and Lewis acid sites on Ni/TiO2 catalyst for vapour phase hydrogenation of levulinic acid: Kinetic and mechanistic study. Appl. Catal. A Gen. 2015, 505, 217–223. [Google Scholar] [CrossRef]
Figure 1. Optimized surface (101) of (a) ZnTiO3 and (b) La/ZnTiO3.
Figure 1. Optimized surface (101) of (a) ZnTiO3 and (b) La/ZnTiO3.
Nanomaterials 12 03137 g001
Figure 2. Band structure of ZnTiO3 and La/ZnTiO3 along the high symmetry directions in the Brillouin zone.
Figure 2. Band structure of ZnTiO3 and La/ZnTiO3 along the high symmetry directions in the Brillouin zone.
Nanomaterials 12 03137 g002
Figure 3. Total density of states (TDOS) of La/ZnTiO3.
Figure 3. Total density of states (TDOS) of La/ZnTiO3.
Nanomaterials 12 03137 g003
Figure 4. Partial density of states (PDOS) of (a) Zn, (b) Ti, (c) O, and (d) La.
Figure 4. Partial density of states (PDOS) of (a) Zn, (b) Ti, (c) O, and (d) La.
Nanomaterials 12 03137 g004aNanomaterials 12 03137 g004b
Figure 5. Charge density difference after La doping on ZnTiO3.
Figure 5. Charge density difference after La doping on ZnTiO3.
Nanomaterials 12 03137 g005
Figure 6. Representation of ELF with contour lines of ZnTiO3 and La/ZnTiO3 surfaces.
Figure 6. Representation of ELF with contour lines of ZnTiO3 and La/ZnTiO3 surfaces.
Nanomaterials 12 03137 g006
Figure 7. Methylene blue (MB) molecule in (a) P0, (b) P1, (c) P2, (d) P3, and (e) P4 orientations on the La/ZnTiO3 surface.
Figure 7. Methylene blue (MB) molecule in (a) P0, (b) P1, (c) P2, (d) P3, and (e) P4 orientations on the La/ZnTiO3 surface.
Nanomaterials 12 03137 g007
Figure 8. Methylene blue (MB) molecule adsorbed on the (a) ZnTiO3 and (b) La/ZnTiO3 surfaces.
Figure 8. Methylene blue (MB) molecule adsorbed on the (a) ZnTiO3 and (b) La/ZnTiO3 surfaces.
Nanomaterials 12 03137 g008
Table 1. Significant bond lengths and angles for La/ZnTiO3 surface.
Table 1. Significant bond lengths and angles for La/ZnTiO3 surface.
Bond Length (Å)Angle (°)
AtomsLa/ZnTiO3AtomsLa/ZnTiO3
La-O42.32La-O4-Ti17117.04
La-O62.32La-O6-Ti26116.25
La-O82.38La-O8-Ti23133.26
Table 2. Adsorption energy and heat segregation for different surfaces.
Table 2. Adsorption energy and heat segregation for different surfaces.
PositionBondEads (kJ/mol)Gseg (kJ/mol)
ZnTiO3O-H−64.06-
La/ZnTiO3: P0-+353.43-
La/ZnTiO3: P1La-S−82.81−57.00
La/ZnTiO3: P2La-S−85.44−57.00
La/ZnTiO3: P3La-N−199.57−58.19
La/ZnTiO3: P4La-N−201.50−58.19
Table 3. Bader’s charge analysis for the selected atoms at the surfaces.
Table 3. Bader’s charge analysis for the selected atoms at the surfaces.
Absorption SystemAtomTotal Electron (-e)
before Adsorption
Total Electron (-e)
after Adsorption
Transfer Charge
(-e)
MB absorbed on ZnTiO3H10.890.88+0.02
H20.880.86+0.01
O17.117.15−0.03
O27.117.14−0.03
MB absorbed on La/ZnTiO3N17.457.77−0.33
La8.858.87+0.02
Table 4. Calculated bandgap energy of La/ZnTiO3 and other experimental energy values reported in the literature.
Table 4. Calculated bandgap energy of La/ZnTiO3 and other experimental energy values reported in the literature.
AdsorbentMethodBandgap (eV)Reference
ZnTiO3/TiO2Experimental3.07[78]
La/ZnTiO3/TiO2Experimental3.04[78]
ZnTiO3Experimental3.54[81]
La/ZnTiO3 (1%)Experimental3.37[81]
La/ZnTiO3 (2%)Experimental2.92[81]
La/ZnTiO3 (3%)Experimental3.35[81]
La/ZnTiO3 (4%)Experimental3.01[81]
La/ZnTiO3 (5%)Experimental3.12[81]
ZnTiO3VASP (GGA/PBE+U)3.16[68]
La/ZnTiO3VASP (GGA/PBE+U)2.98This study
Table 5. Calculated molecular adsorption energy values of MB and other dyes on surfaces (101) of La/ZnTiO3 and ZnTiO3.
Table 5. Calculated molecular adsorption energy values of MB and other dyes on surfaces (101) of La/ZnTiO3 and ZnTiO3.
AdsorbentDyeSoftware UsedAdsorption (kJ/mol)References
ZnTiO3s-Cu-TTCVASP−296.56[60]
ZnTiO3TPA-1CASTEP−136.39[63]
ZnTiO3TPA-2CASTEP−157.47[63]
ZnTiO3TPA-3CASTEP−561.33[63]
ZnTiO3TPA-4CASTEP−228.19[63]
ZnTiO3 (H)MBVASP−126.76[68]
ZnTiO3 (SP)MBVASP−282.05[68]
ZnTiO3 (P4)MBVASP−64.06This study
La/ZnTiO3 (P1)MBVASP−82.81This study
La/ZnTiO3 (P2)MBVASP−85.44This study
La/ZnTiO3 (P3)MBVASP−199.57This study
La/ZnTiO3 (P4)MBVASP−201.50This study
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jaramillo-Fierro, X.; Cuenca, G.; Ramón, J. The Effect of La3+ on the Methylene Blue Dye Removal Capacity of the La/ZnTiO3 Photocatalyst, a DFT Study. Nanomaterials 2022, 12, 3137. https://doi.org/10.3390/nano12183137

AMA Style

Jaramillo-Fierro X, Cuenca G, Ramón J. The Effect of La3+ on the Methylene Blue Dye Removal Capacity of the La/ZnTiO3 Photocatalyst, a DFT Study. Nanomaterials. 2022; 12(18):3137. https://doi.org/10.3390/nano12183137

Chicago/Turabian Style

Jaramillo-Fierro, Ximena, Guisella Cuenca, and John Ramón. 2022. "The Effect of La3+ on the Methylene Blue Dye Removal Capacity of the La/ZnTiO3 Photocatalyst, a DFT Study" Nanomaterials 12, no. 18: 3137. https://doi.org/10.3390/nano12183137

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