Next Article in Journal
Hypoglycemic and Anti-Inflammatory Effects of Triterpene Glycoside Fractions from Aeculus hippocastanum Seeds
Next Article in Special Issue
Recent Advances in Biopolymer-Based Dye Removal Technologies
Previous Article in Journal
Catalytic Activity of Thermolyzed [Co(NH3)6][Fe(CN)6] in CO Hydrogenation Reaction
Previous Article in Special Issue
Adsorption and Desorption Properties of Polyethylenimine/Polyvinyl Chloride Cross-Linked Fiber for the Treatment of Azo Dye Reactive Yellow 2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

DFT Study of Methylene Blue Adsorption on ZnTiO3 and TiO2 Surfaces (101)

by
Ximena Jaramillo-Fierro
1,2,*,
Luis Fernando Capa
3,
Francesc Medina
1 and
Silvia González
2
1
Departamento d’Enginyería Química, Universitat Rovira i Virgili, Av Països Catalans, 2643007 Tarragona, Spain
2
Departamento de Química y Ciencias Exactas, Universidad Técnica Particular de Loja, San Cayetano Alto, 1101608 Loja, Ecuador
3
Maestría en Química Aplicada, Universidad Técnica Particular de Loja, San Cayetano Alto, 1101608 Loja, Ecuador
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(13), 3780; https://doi.org/10.3390/molecules26133780
Submission received: 25 May 2021 / Revised: 16 June 2021 / Accepted: 19 June 2021 / Published: 22 June 2021
(This article belongs to the Special Issue Recent Advances in Dyes Removal Technologies)

Abstract

:
The search for alternative materials with high dye adsorption capacity, such as methylene blue (MB), remains the focus of current studies. This computational study focuses on oxides ZnTiO3 and TiO2 (anatase phase) and on their adsorptive properties. Computational calculations based on DFT methods were performed using the Viena Ab initio Simulation Package (VASP) code to study the electronic properties of these oxides. The bandgap energy values calculated by the Hubbard U (GGA + U) method for ZnTiO3 and TiO2 were 3.17 and 3.21 eV, respectively, which are consistent with the experimental data. The most favorable orientation of the MB adsorbed on the surface (101) of both oxides is semi-perpendicular. Stronger adsorption was observed on the ZnTiO3 surface (−282.05 kJ/mol) than on TiO2 (–10.95 kJ/mol). Anchoring of the MB molecule on both surfaces was carried out by means of two protons in a bidentate chelating (BC) adsorption model. The high adsorption energy of the MB dye on the ZnTiO3 surface shows the potential value of using this mixed oxide as a dye adsorbent for several technological and environmental applications.

1. Introduction

Over the past decade, Ti- and Zn-based oxides have received much attention due to their competitive cost, non-toxicity, excellent stability, availability, and ability to produce highly oxidizing radicals [1,2,3,4]. Titanium oxide (TiO2) and zinc oxide (ZnO) are two well-known semiconductors that have been widely used to construct electron transport channels due to their appropriate bandgaps, efficient electron mobilities, and simple synthesis methods [5,6,7,8]. In addition, these oxides are promising semiconductors to eliminate organic pollutants with incomparable efficiency due to their tunable surface and structural functionality [9]. The ZnO–TiO2 composite system has even more superior properties than the individual oxides due to the high separation rate of photogenerated carriers and the wide optical response range [10]. Several syntheses and characterization studies of the ZnO–TiO2 system have shown that there are three compounds in this binary system, including ZnTiO3 (cubic, hexagonal), Zn2TiO4 (cubic, tetragonal), and Zn3Ti3O8 (cubic) [11,12,13].
ZnTiO3 has many similar physical properties to TiO2 and ZnO, including high electron mobility [14,15,16]. This ternary oxide has been widely used because of its outstanding properties and potential scientific and technical applications [17]. ZnTiO3 has been investigated in a variety of applications as an antibacterial, catalyst, nanofiber, white pigment, microwave dielectric, gas sensor, nonlinear optical, corrosion inhibitor, and luminescent material [18,19,20,21,22,23], but its application in adsorption has not been sufficiently studied, despite the fact that the literature indicates that due to its great specific area, it could have an important potential as an adsorbent [24,25].
ZnTiO3 is a polar oxide of the LiNbO3-type (LN-type) with both cations coordinated octahedrally in a three-dimensional framework of the octahedron perovskite (Pv) that shares corners [26]. In this structure, both cations move along the trigonal axis c, thus producing a spontaneous polarization reinforced by a second-order Jahn–Teller (SOJT) distortion due to Ti4+ (d0) [27]. The paraelectric parent structure of ZnTiO3 is the ilmenite (Il)-type phase (hexagonal space group R-3), which is the stable phase under ambient conditions [28]. The crystalline and phase transformation behaviors of ZnTiO3 have systematically been investigated by various authors regarding several synthesis methods, Ti:Zn precursor molar ratios, and calcination temperatures [29,30,31,32,33,34,35]. Furthermore, the literature agrees that obtaining ZnTiO3 as a pure phase at a low processing temperature is a challenge in materials chemistry [36,37,38].
In a previous experimental paper, we reported the synthesis and characterization of the ZnTiO3/TiO2 nanocomposite. This heterostructure was indexed to a hexagonal phase with space group R-3(148) for ZnTiO3 and a tetragonal phase with space group I41/amd(141) for TiO2 (anatase) [39,40]. In these studies, we also reported the ability of ZnTiO3/TiO2 to remove the methylene blue (MB) dye in aqueous systems, and it was contrasted with the results obtained for pure TiO2 (anatase). The results showed that the heterostructure has better photocatalytic adsorption and degradation capacity than anatase alone, probably due to a synergistic effect. This synergistic effect between semiconductors has been extensively studied, showing that the presence of a second semiconductor can provide special active sites to enhance the adsorption and photocatalysis of various compounds [41].
Although several properties of ZnTiO3 have been extensively studied experimentally, a proper description of its electronic, optical, and adsorptive properties remains an active research area from a theoretical point of view [42] since, up to date, this ternary oxide has scarcely been studied with quantum methods [43]. Therefore, the computational study of the molecular interaction between ZnTiO3 and methylene blue could contribute to clarifying the adsorption and degradation mechanism of this dye, favoring the development of materials for the treatment of waters contaminated with MB.
The elimination of MB in wastewater is an extremely important task in environmental protection because it has caused serious contamination in many countries of the world [44]. Methylene blue, known as methylthioninium chloride is a basic cationic dye widely used in the printing, plastics, paper, leather, food, pharmaceutical, and textile industries [45,46,47]. The discharge of wastewater effluents from these industries, with a high content of MB without efficient degradation, results in harmful effects for humans and animals [48].
Various technologies have been used to treat wastewater contaminated with dyes [49]. Among these techniques, adsorption is easy to perform without pretreatment and is highly selective for removing dyes [50]. In addition, adsorption has been found to be superior to other techniques for wastewater treatment in terms of initial cost, simplicity of design, ease of operation, and insensitivity to toxic substances [24]. Although there are several experimental studies of MB adsorption on different surfaces, some uncertainties remain due to lack of understanding at the molecular level of the MB adsorption mechanism on the ZnTiO3 surface.
As is well known, computational calculations of the electronic structure in an isolated molecule can achieve the desired chemical precision as long as a sufficiently large basis set is used, the electronic correlation is sufficiently described, and the relativistic effects in the calculation are adequately included [51]. Therefore, in this study, Density Functional Theory (DFT) computational calculations were used to characterize the electronic structure of ZnTiO3 and TiO2 (anatase) and also to investigate the feasibility of using both oxides as MB adsorbents. The results presented in this paper clarify the previously obtained experimental results and confirm that ZnTiO3 is an excellent adsorbent and that it has high potential for future technological and environmental applications.

2. Results

2.1. Optimization and Electronic Structure of ZnTiO3 and TiO2

The adsorption of the methylene blue molecule on the surface of both ZnTiO3 and TiO2 was modeled using the following parameters: hexagonal ZnTiO3 with a cell = 5.148 Å × 5.148 Å × 13.937 Å <90° × 90° × 120°> and tetragonal TiO2 with a cell = 3.821 Å × 3.821 Å × 9.697 Å <90° × 90° × 90°>, as shown in Figure 1. The coordinates of the optimized ZnTiO3 and TiO2 structures are detailed in Table S1 and the corresponding optimization energy values are included in Figure S2.
The selection of the high symmetry points and lines in the first Brillouin zone [52] and the results of the calculation of the electronic band structure of ZnTiO3 and TiO2 are shown in Figure 2a,b, respectively.
Figure 2a,b show that the indirect bandgap energy values of the ZnTiO3 and TiO2 structures calculated by the exchange–correlation functional in the generalized gradient approximation (GGA-PBE) method were 2.20 and 2.31 eV, respectively. However, the indirect bandgap values were also calculated by the GGA + U method, that is, incorporating the Hubbard U approximation term. Indirect bandgap calculations using GGA + U resulted in 3.16 eV (U = 2.5) and 3.21 eV (U = 4.0) for ZnTO3 and TiO2, respectively. These results are in good agreement with the experimental results reported in the literature: 3.18 eV for ZnTiO3 [15] and 3.20 eV for TiO2 [53].
The total and partial density of states (DOS) of ZnTiO3 and TiO2 are illustrated in Figure 3 and Figure 4, respectively. Figure 3a shows that the total density of state (TDOS) of ZnTiO3 has two main zones: an upper conduction band (CB) zone from 2.5 to 6.2 eV, and a lower valence band (VB) zone, from −6.0 to −0.2 eV. The CB is dominated by the contribution of Ti, while the VB is dominated by the contribution of Zn and O. Figure 3b–d show the partial density of state (PDOS) of ZnTiO3. As can be seen in Figure 3b, the main contribution of Ti in the CB is through the 3d orbital. On the other hand, Figure 3c shows that Zn contributes mainly to the VB through the 3d orbital while O interacts with Zn in this band through the 2p orbital shown in Figure 3d.
Likewise, Figure 4a shows that the total density of state (TDOS) of TiO2 has two main zones: an upper conduction band (CB) zone from −0.1 to 3.8 eV, and a lower valence band (VB) zone, from −7.6 to −2.8 eV. The CB is dominated by the contribution of Ti, while the VB is dominated by the contribution of O. Figure 4b,c show the partial density of state (PDOS) of TiO2. As can be seen in Figure 4b, the main contribution of Ti in the CB is through the 3d orbital. On the other hand, Figure 4c shows that O contributes mainly to the VB through the 2p orbital.
In both structures, ZnTiO3 and TiO2, the valence band maximum (VBM) is bordered by the oxygen atom, while the Ti atom determines the conduction band maximum (CBM). Consequently, ZnTiO3 has an energy bandgap quite similar to that of TiO2, due to the fact that ZnTiO3 involves both ZnO and TiO2. Our calculated results agree with the literature [5,18,54].
In order to further understand the chemical bonding of hexagonal ZnTiO3 and tetragonal TiO2, the population analyses were estimated by the Bader method. For ZnTiO3, the net charge of Ti (+2.6e) was 1.4e, much smaller than its +4e formal charge, whereas the Zn atom had a positive charge of +1.4e and the O atom had a negative charge of −1.3e, which are less than their +2e and −2e formal charges by 0.6e and 0.7e, respectively. For TiO2, the net charges of the Ti and O atoms were similar to those calculated for the Ti and O atoms of ZnTiO3. These results agree with those reported by other authors [55]. Since the charges on the different bonds can reflect the covalent and ionic properties of the molecule, we concluded that the Ti–O bond is typically covalent for both ZnTiO3 and TiO2 and that the Zn–O bond for ZnTiO3 is typically ionic; these results coincide with those reported in the literature [42,56].

2.2. Adsorption of the MB Dye on the Structures

The orientations of the MB molecule on the ZnTiO3 and TiO2 surfaces are shown in Figure 5. Figure 5a shows the horizontal orientation of the MB molecule on the ZnTiO3 surface, while Figure 5b,c show the semi-perpendicular orientation of the MB molecule on the ZnTiO3 and TiO2 surfaces, respectively. The adsorption of the MB molecule on the ZnTiO3 surface with the molecule placed in semi-perpendicular orientation (Eads = −2.916 eV) was more energetically favored than in the horizontal orientation (Eads = −1.310 eV). Therefore, we studied the adsorption of MB on the TiO2 surface only with the semi-perpendicular orientation. The calculated adsorption energy for the TiO2 surface (Eads = −0.113 eV) was less favorable than the calculated adsorption energy for the ZnTiO3 surface.
The anchoring modes of the MB molecule on the ZnTiO3 and TiO2 surfaces are shown in Figure 6. Adsorption of the dye on the ZnTiO3 and TiO2 surfaces occurs in a bidentate chelating (BC) adsorption model [57] with two protons oriented toward the nearest surface oxygen [58].
The calculated adsorption energy value indicates that the MB molecule is strongly adsorbed on the ZnTiO3 surface. The average distances from the hydrogen atoms of the MB molecule (HMB) to the surface plane of ZnTiO3 are O(oxide)-HMB = 2.34 Å and O(oxide)-HMB = 2.52 Å. Moreover, the adsorption energy value of the MB molecule on the TiO2 surface indicates a weaker interaction than on ZnTiO3 (Eads = −0.113 eV). The average distances from the hydrogen atoms of the molecule to the plane of the TiO2 surface are O(oxide)-HMB = 2.68 Å and O(oxide)-HMB = 2.69 Å.

3. Discussion

3.1. Optimization and Electronic Structure of ZnTiO3 and TiO2

The description of the electronic structure of materials involving transition metals with DFT is often complicated due to correlation effects involving 3d electrons [59]. However, since the transition metals of the oxides studied in this paper are formally in the 3d0 or 3d10, neither TiO2 nor ZnTiO3 are strongly correlated materials; consequently, a simple DFT approach allowed us to accurately calculate the conduction and valence bands of ZnTiO3 and TiO2.
The literature shows that ZnTiO3 has a relatively wide bandgap (Eg = 2.73–3.70 eV), the value of which depends on the synthesis conditions [60,61]. In our study, the ZnTiO3 and TiO2 structures presented indirect bandgap values of 2.20 and 2.31 eV, respectively, which were calculated by the GGA-PBE method. In contrast with the experimental data, 3.18 eV for ZnTiO3 [15] and 3.20 eV for TiO2 [53], the theoretical values are lower, and this may be due to the widely known DFT-underestimation of the bandgap in most materials [1,15]. Therefore, a Hubbard U approximation term was adopted to accurately describe the electronic structures [12,62]. The new indirect bandgap values calculated by GGA + U were 3.16 and 3.21 eV for ZnTiO3 and TiO2, respectively, which are consistent with the aforementioned experimental data. As can be seen, mixed oxide ZnTiO3 has lower bandgap energy than TiO2. According to the literature, this occurs due to the replacement of Ti (3d0) atoms with Zn (3d10) that induce O 2p-Zn 3d10 repulsion [21,63]. Table 1 shows the comparison of the bandgap energy values of the ZnTiO3 and TiO2 calculated in this study with other energy values reported in the literature.
The main character of the electronic structure of ZnTiO3 originates mainly from the hybridization between the Ti-3d and O-2p states. The Zn-3d and O-2p hybridization and the Ti-3d and O-2p hybridization as well as nonbonding O-2p states are observed at the upper valence bands (VBs). The localized Zn-3d states indicate weak Zn-3d and O-2p hybridization. The states at the lower conduction bands (CBs) are attributed to antibonding states from Ti-3d and O-2p. These results agree with those reported by other authors [42]. Similarly, the main character of the electronic structure of TiO2 originates from the hybridization between the Ti-3d and O-2p states. This hybridization, as well as nonbonding O-2p states, is observed at the upper valence bands (VBs), while the states at the lower conduction bands (CBs) are attributed to antibonding states from Ti-3d and O-2p. In both structures, the lowest-energy states are due to the isolated Ti-3s, Ti-3p, and O-2s states. These results also agree with those reported by other authors [5,54,65].

3.2. Adsorption of the MB Dye on the Oxide Models

Several experimental studies of MB removal in aqueous systems have shown that this cationic dye can be easily adsorbed on the ZnTiO3 and TiO2 surfaces, due to electrostatic attraction [49]. Therefore, the surface oxygen atoms of both oxides probably generate negatively charged sites that easily attract positive regions of the MB molecule, favoring molecular adsorption.
Our results indicated that the semi-perpendicular orientation of the MB molecule with respect to both oxide surfaces is more favored. In fact, the MB molecule oriented parallel to the surface shows a strong preference for the methyl group of the molecule, while the aromatic ring bends slightly away from the surface due to electrostatic repulsion between the N and S atoms from the aromatic ring and surface oxygen. This is consistent with several studies reporting good adsorption results for dye molecules oriented perpendicular to the adsorbent surface [57,66,67]. Greathouse et al. mentioned that at very high concentrations, this dye forms aggregates that are adsorbed vertically to the surface [68]. Our results suggested that this orientation of the MB molecule on the surface is caused by the balance between electrostatic repulsion between adjacent ions and the strong hydrophobic MB–MB and MB–surface interactions [69].
The MB molecule was adsorbed on the ZnTiO3 surface (101) with higher negative energy (Eads = −2.92 eV) than on the TiO2 surface (101) (Eads = −0.12 eV), and the average adsorption distances (O–H) were 2.43 and 2.68 Å for ZnTiO3 and TiO2, respectively. According to the optimized configurations, in both cases, the approach is from the two H atoms of the methyl group of the MB molecule to an O atom of each surface. In agreement with the literature, hydrogen bonding can increase the stability of the interaction of a dye with the ZnTiO3 and TiO2 surfaces during the adsorption process [70,71]. The Bader charge analysis in Table S1 indicates no significant electronic exchange between the MB molecule and the oxides surface [72]. The results obtained in this theoretical study suggest that MB adsorption is more stable (higher negative adsorption energy) on the ZnTiO3 surface than on the TiO2 surface, which is consistent with previous experimental studies.
Pastore et al. suggested three typical coordination schemes: monodentate, bidentate chelating, and bidentate bridging [57]. In our study, we found that the MB molecule is adsorbed on the ZnTiO3 and TiO2 surfaces (101) in a bidentate chelating mode, which, according to several authors, produces more stable adsorption with more exothermic adsorption energy [70,73]. The shape of the dye molecules anchored to the oxide significantly affects the molecular adsorption energy. Therefore, calculating the adsorption energy value gives insight into the adherence strength and shape of molecule–surface bonding [66]. The higher the adsorption energy, the higher the retention of the dye on the oxide surface, this being a desirable condition to apply to subsequent photosensitive processes.
In the literature, not enough computational studies of MB adsorption on semiconductors were found; consequently, in Table 2, the results obtained in the present study are compared with those results reported in the literature for the adsorption of other dyes on ZnTiO3 (101) and TiO2 (101).
Other research studies should analyze and apply these results to implement adsorptive systems for dyes or similar molecules.

4. Materials and Methods

All Density Functional Theory (DFT) calculations were performed using the Viena Ab initio Simulation Package (VASP), version 5.3.3 [15,74]. The Perdew–Burke–Ernzerhof (PBE) exchange–correlation functional in the generalized gradient approximation (GGA) proposed by Perdew et al. [75] was employed. The augmented plane wave (PAW) method was used to describe the electron–ion interactions [15]. The cutoff energy to the plane waves was set to 500 eV. The Kohn–Sham equations [76,77] were solved self-consistently until the energy variation between cycles was less than 10−5 eV. The first Brillouin zone was sampled using Monkhorst–Pack [78] k-point meshes to calculate the bulk properties of ZnTiO3 and TiO2, in particular, 3 × 7 × 5 and 3 × 3 × 1, respectively. All atomic positions were fully relaxed until the forces on each atom were below 0.01 eV/Å. The computational parameters were selected seeking the best balance between computational cost and precision. The tested values were as follows: energy cutoff points = 450, 475, 500, and 515 eV; force convergence criterion for ionic relaxation = 0.08, 0.04, 0.02, 0.01, and 0.005 eV/Å and number of k-points corresponding to k-spacing in each axe = 0.35, 0.30, 0.25, 0.20, and 0.15. The parameters were optimized until the difference between the energy values of the system was lower than 10−4 eV.
The Gaussian smearing method with σ = 0.10 eV was applied to band occupations in order to improve total energy convergence [15]. A Hubbard U approximation term was adopted to describe the strong on-site Coulomb repulsion in order to accurately explain the electronic structures [62], which is not correctly described by the PBE functional [58]. Population analyses were estimated using Bader’s charge analysis code, which provided important information on bonding behaviors from the atomic charge values [79,80,81]. All calculations were non-spin polarized and all molecular models were created and visualized using BioVia Material Studio, version 5.5.
To study MB adsorption, an optimized molecular structure was used [68]. The bulk of both ZnTiO3 and TiO2 crystals was cleaved on the surface (101), since it is the most stable surface according to the literature [54,57,62,66,82]. The slab model of ZnTiO3 (101) was a supercell p(2 × 3) with three atomic layers, which includes 36 Zn atoms, 36 Ti atoms, and 108 O atoms. On the other hand, the TiO2 (101) surface model has seven atomic layers with a p(3 × 3) structure of the original unit cell, which includes 168 Ti atoms and 336 O atoms.
An appropriate vacuum thickness of each structure was chosen by calculating the surface energy. For both ZnTiO3 and TiO2 surface models, a vacuum of 20 Å was added. The surface energies (γs) were calculated using the following equation [83]:
γ s = ( E s l a b n × E b u l k ) 2 A
where Eslab is the total energy of the slab material (eV), Ebulk is the total energy of the bulk material (eV), n is the number of atoms contained in the slab, and A is the surface area (Å2). The values for the surface energies (γs) of the ZnTiO3 and TiO2 structures with a vacuum distance of 20 Å were 0.076 eV/Å2 (7.30 kJ/Å2) and 0.062 eV/Å2 (5.98 kJ/Å2), respectively.
Adsorption calculation was initiated with the MB molecule placed close to the surface of each oxide in at least one of the following orientations, horizontal (H) and semi-perpendicular (SP), with respect to the surface.
The adsorption energy (ΔEads) of the MB molecule on the surface of both ZnTiO3 and TiO2 oxides was calculated using the following equation [84]:
Δ E a d s = E M B / o x i d e E o x i d e E M B
where EMB/oxide is the energy of the supersystem formed by the adsorbed molecule on the surface (eV), Eoxide is the energy of the clean oxide (eV), and EMB is the energy of the isolated molecule in vacuum (eV).

5. Conclusions

The aim of this comparative study was to use molecular simulation to address unresolved issues related to the adsorption mechanism of methylene blue on both ZnTiO3 and TiO2. DFT calculations of MB adsorption on the surface (101) of both ZnTiO3 and TiO2 indicated that adsorption on ZnTiO3 was stronger than on TiO2. The semi-perpendicular orientation was the most probable molecular approach to the oxide surfaces. Electrostatic repulsion due to the proximity of adjacent S and N atoms when MB was in high concentration was overcome by the much stronger interactions between the methyl groups and the surface oxygen atoms of ZnTiO3 and TiO2.
Finally, we computationally corroborated the feasibility of using ZnTiO3 as an MB adsorbent material, as experimentally found. Theoretically, we forecast the appealing prospect for this material according to the adsorption energy and the large bandgap calculated by DFT, which is in addition to the experimental results that we reported in a previous paper. Our study verifies that ZnTiO3 has better MB adsorption energy than TiO2 in the anatase phase, which is important to enhance a subsequent degradation process. The large bandgap obtained by DFT calculations also shows that ZnTiO3 can potentially be used as a photocatalyst, allowing for complete degradation of the dye after being adsorbed. Therefore, considering only the band structure, ZnTiO3 fully meets the necessary requirements to be a photocatalyst. As already mentioned, however, in addition to the band structure, the adsorption capacity is also very important for photocatalytic materials. In this way, ZnTiO3 constitutes an efficient alternative material for various technological and environmental applications.

Supplementary Materials

The following are available online. Figure S1: Aromatic ring of MB bent slightly on the ZnTiO3 surface (101); Figure S2: Optimization energies of ZnTiO3, TiO2 and MB; Table S1: Coordinates of the optimized structures; Table S2: Bader’s charge analysis of the methylene blue molecule.

Author Contributions

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

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Acknowledgments

The authors would like to thank Henry Quezada for the computer support. This paper was financially supported by Universitat Rovira I Virgili (Spain) and Universidad Técnica Particular de Loja (Ecuador).

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. 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. [Google Scholar] [CrossRef]
  2. Li, L.; Zhang, X.; Zhang, W.; Wang, L.; Chen, X.; Gao, Y. Microwave-assisted synthesis of nanocomposite Ag/ZnO-TiO2 and photocatalytic degradation Rhodamine B with different modes. Colloids Surf. A Physicochem. Eng. Asp. 2014, 457, 134–141. [Google Scholar] [CrossRef]
  3. 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]
  4. Grabis, J.; Letlena, A.; Sīpola, I. Preparation and properties of photocalysts in ZnO/TiO2 system. Key Eng. Mater. 2019, 800, 170–174. [Google Scholar] [CrossRef]
  5. 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]
  6. 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] [Green Version]
  7. Irani, M.; Mohammadi, T.; Mohebbi, S. Photocatalytic degradation of methylene blue with zno nanoparticles; a joint experimental and theoretical study. J. Mex. Chem. Soc. 2017, 60, 218–225. [Google Scholar] [CrossRef]
  8. 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]
  9. 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] [Green Version]
  10. Chen, F.; Yu, C.; Wei, L.; Fan, Q.; Ma, F.; Zeng, J.; Yi, J.; Yang, K.; Ji, H. Fabrication and characterization of ZnTiO3/Zn2Ti3O8/ZnO ternary photocatalyst for synergetic removal of aqueous organic pollutants and Cr(VI) ions. Sci. Total Environ. 2020, 706, 136026. [Google Scholar] [CrossRef]
  11. 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. 2019, 35, 281–289. [Google Scholar] [CrossRef]
  12. Dutta, D.P.; Singh, A.; Tyagi, A. Ag doped and Ag dispersed nano ZnTiO3: Improved photocatalytic organic pollutant degradation under solar irradiation and antibacterial activity. J. Environ. Chem. Eng. 2014, 2, 2177–2187. [Google Scholar] [CrossRef]
  13. Fu, R.; Wang, Q.; Gao, S.; Wang, Z.; Huang, B.; Dai, Y.; Lu, J. Effect of different processes and Ti/Zn molar ratios on the structure, morphology, and enhanced photoelectrochemical and photocatalytic performance of Ti3+ self-doped titanium-zinc hybrid oxides. J. Power Sources 2015, 285, 449–459. [Google Scholar] [CrossRef] [Green Version]
  14. 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]
  15. Yu, J.; Li, N.; Zhu, L.; Xu, X. Application of ZnTiO3 in quantum-dot-sensitized solar cells and numerical simulations using first-principles theory. J. Alloys Compd. 2016, 681, 88–95. [Google Scholar] [CrossRef]
  16. 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]
  17. Baamran, K.S.; Tahir, M. Ni-embedded TiO2-ZnTiO3 reducible perovskite composite with synergistic effect of metal/support towards enhanced H2 production via phenol steam reforming. Energy Convers. Manag. 2019, 200, 112064. [Google Scholar] [CrossRef]
  18. Liu, Q.-J.; Zhang, N.-C.; Liu, F.-S.; Wang, H.-Y.; Liu, Z.-T. Theoretical study of structural, elastic, electronic properties, and dispersion of optical functions of hexagonal ZnTiO3. Phys. Status Solidi Basic Res. 2013, 250, 1810–1815. [Google Scholar] [CrossRef]
  19. Yan, X.; Zhao, C.-L.; Zhou, Y.-L.; Wu, Z.-J.; Yuan, J.-M.; Li, W.-S. Synthesis and characterization of ZnTiO3 with high photocatalytic activity. Trans. Nonferrous Met. Soc. China Engl. Ed. 2015, 25, 2272–2278. [Google Scholar] [CrossRef]
  20. Li, M.; Jiao, B. Synthesis and photoluminescence properties of ZnTiO3:Eu3+ red phosphors via sol-gel method. J. Rare Earths 2015, 33, 231–238. [Google Scholar] [CrossRef]
  21. Lv, J.; Tang, M.; Quan, R.; Chai, Z. Synthesis of solar heat-reflective ZnTiO3 pigments with novel roof cooling effect. Ceram. Int. 2019, 45, 15768–15771. [Google Scholar] [CrossRef]
  22. Sowmyashree, A.; Somya, A.; Kumar, C.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]
  23. Tavakoli-Azar, T.; Mahjoub, A.R.; Sadjadi, M.S.; Farhadyar, N.; Sadr, M.H. Improving the photocatalytic performance of a perovskite ZnTiO3 through ZnTiO3@S nanocomposites for degradation of Crystal violet and Rhodamine B pollutants under sunlight. Inorg. Chem. Commun. 2020, 119, 108091. [Google Scholar] [CrossRef]
  24. Raveendra, R.; Prashanth, P.; Krishna, R.H.; Bhagya, N.; Nagabhushana, B.; Naika, H.R.; Lingaraju, K.; Nagabhushana, H.; Prasad, B.D. Synthesis, structural characterization of nano ZnTiO3 ceramic: An effective azo dye adsorbent and antibacterial agent. J. Asian Ceram. Soc. 2014, 2, 357–365. [Google Scholar] [CrossRef] [Green Version]
  25. 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]
  26. Sahu, A.; Chaurasiya, 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]
  27. Inaguma, Y.; Aimi, A.; Shirako, Y.; Sakurai, D.; Mori, D.; Kojitani, H.; Akaogi, M.; Nakayama, M. High-pressure synthesis, crystal structure, and phase stability relations of a LiNbO3-type polar titanate ZnTiO3 and its reinforced polarity by the second-order Jahn-Teller effect. J. Am. Chem. Soc. 2013, 136, 2748–2756. [Google Scholar] [CrossRef]
  28. Ruiz-Fuertes, J.; Winkler, B.; Bernert, T.; Bayarjargal, L.; Morgenroth, W.; Koch-Müller, M.; Refson, K.; Milman, V.; Tamura, N. Ferroelectric soft mode of polar ZnTiO3 investigated by Raman spectroscopy at high pressure. Phys. Rev. B Condens. Matter Mater. Phys. 2015, 91, 214110. [Google Scholar] [CrossRef] [Green Version]
  29. Pawar, R.C.; Kang, S.; Park, J.H.; Kim, J.-H.; Ahn, S.; Lee, C.S. Evaluation of a multi-dimensional hybrid photocatalyst for enrichment of H2 evolution and elimination of dye/non-dye pollutants. Catal. Sci. Technol. 2017, 7, 2579–2590. [Google Scholar] [CrossRef]
  30. Salavati-Niasari, M.; Soofivand, F.; Sobhani-Nasab, A.; Shakouri-Arani, M.; Faal, A.Y.; Bagheri, S. Synthesis, characterization, and morphological control of ZnTiO3 nanoparticles through sol-gel processes and its photocatalyst application. Adv. Powder Technol. 2016, 27, 2066–2075. [Google Scholar] [CrossRef] [Green Version]
  31. Ke, S.; Cheng, X.; Wang, Q.; Wang, Y.; Pan, Z. Preparation of a photocatalytic TiO2/ZnTiO3 coating on glazed ceramic tiles. Ceram. Int. 2014, 40, 8891–8895. [Google Scholar] [CrossRef]
  32. Acosta-Silva, Y.D.J.; Castanedo-Perez, R.; Torres-Delgado, G.; Méndez-López, A.; Zelaya-Angel, O. Analysis of the photocatalytic activity of CdS+ZnTiO3 nanocomposite films prepared by sputtering process. Superlattices Microstruct. 2016, 100, 148–157. [Google Scholar] [CrossRef]
  33. Yadav, B.; Yadav, A.; Singh, S.; Singh, K. Nanocrystalline zinc titanate synthesized via physicochemical route and its application as liquefied petroleum gas sensor. Sens. Actuators B Chem. 2013, 177, 605–611. [Google Scholar] [CrossRef]
  34. 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. Hydrog. Energy 2020, 45, 9484–9495. [Google Scholar] [CrossRef]
  35. 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]
  36. Bhagwat, U.O.; Wu, J.J.; Asiri, A.M.; Anandan, S. Synthesis of ZnTiO3@TiO2 heterostructure nanomaterial as a visible light photocatalyst. ChemistrySelect 2019, 4, 6106–6112. [Google Scholar] [CrossRef]
  37. Lei, S.; Fan, H.; Ren, X.; Fang, J.; Ma, L.; Liu, Z. Novel sintering and band gap engineering of ZnTiO3 ceramics with excellent microwave dielectric properties. J. Mater. Chem. C 2017, 5, 4040–4047. [Google Scholar] [CrossRef]
  38. 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]
  39. Jaramillo-Fierro, X.; Pérez, S.G.; Jaramillo, X.; Cabello, F.M. Synthesis of the ZnTiO3/TiO2 nanocomposite supported in ecuadorian clays for the adsorption and photocatalytic removal of methylene blue dye. Nanomaterials 2020, 10, 1891. [Google Scholar] [CrossRef]
  40. 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]
  41. Zhang, Z.; Yates, J.J.T. Band bending in semiconductors: Chemical and physical consequences at surfaces and interfaces. Chem. Rev. 2012, 112, 5520–5551. [Google Scholar] [CrossRef] [PubMed]
  42. Xiao-Chao, Z.; Cai-Mei, F.; Zhen-Hai, L.; Pei-De, H. Electronic structures and optical properties of ilmenite-type hexagonal ZnTiO3. Wuli Huaxue Xuebao/Acta Phys. Chim. Sin. 2011, 27, 47–51. [Google Scholar] [CrossRef]
  43. Conesa, J.C. Band structures and nitrogen doping effects in zinc titanate photocatalysts. Catal. Today 2013, 208, 11–18. [Google Scholar] [CrossRef] [Green Version]
  44. Gil, A.; Assis, F.; Albeniz, S.; Korili, S. Removal of dyes from wastewaters by adsorption on pillared clays. Chem. Eng. J. 2011, 168, 1032–1040. [Google Scholar] [CrossRef]
  45. Laysandra, L.; Sari, M.W.M.K.; Soetaredjo, F.E.; Foe, K.; Putro, J.N.; Kurniawan, A.; Ju, Y.-H.; Ismadji, S. Adsorption and photocatalytic performance of bentonite-titanium dioxide composites for methylene blue and rhodamine B decoloration. Heliyon 2017, 3, e00488. [Google Scholar] [CrossRef] [Green Version]
  46. Mehrabi, M.; Javanbakht, V. Photocatalytic degradation of cationic and anionic dyes by a novel nanophotocatalyst of TiO2/ZnTiO3/αFe2O3 by ultraviolet light irradiation. J. Mater. Sci. Mater. Electron. 2018, 29, 9908–9919. [Google Scholar] [CrossRef]
  47. Subramaniam, M.N.; Goh, P.S.; Abdullah, N.; Lau, W.J.; Ng, B.C.; Ismail, A.F. Adsorption and photocatalytic degradation of methylene blue using high surface area titanate nanotubes (TNT) synthesized via hydrothermal method. J. Nanoparticle Res. 2017, 19, 220. [Google Scholar] [CrossRef]
  48. Abdellah, M.; Nosier, S.; El-Shazly, A.; Mubarak, A. Photocatalytic decolorization of methylene blue using TiO2/UV system enhanced by air sparging. Alex. Eng. J. 2018, 57, 3727–3735. [Google Scholar] [CrossRef]
  49. Sutanto, N.; Saharudin, K.A.; Sreekantan, S.; Kumaravel, V.; Akil, H.M. Heterojunction catalysts g-C3N4/-3ZnO-c-Zn2Ti3O8 with highly enhanced visible-light-driven photocatalytic activity. J. Sol-Gel Sci. Technol. 2020, 93, 354–370. [Google Scholar] [CrossRef]
  50. Zhao, Q.; Zhang, L.; Wang, X.; Jia, X.; Xu, P.; Zhao, M.; Dai, R. Simultaneous efficient adsorption and photocatalytic degradation of methylene blue over iron(III)-based metal–organic frameworks: A comparative study. Transit. Met. Chem. 2019, 44, 789–797. [Google Scholar] [CrossRef]
  51. Chang-Guo, Z. Development and application of first-principles electronic structure approach for molecules in solution based on fully polarizable continuum model. Wuli Huaxue Xuebao/Acta Phys. Chim. Sin. 2011, 27, 1–10. [Google Scholar] [CrossRef]
  52. 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] [Green Version]
  53. Zhang, R.; Zhao, J.; Yang, Y.; Lu, Z.; Shi, W. Understanding electronic and optical properties of La and Mn co-doped anatase TiO2. Comput. Condens. Matter 2016, 6, 5–17. [Google Scholar] [CrossRef] [Green Version]
  54. Qi, K.; Selvaraj, R.; Al Fahdi, T.; Al-Kindy, S.; Kim, Y.; Wang, G.-C.; Tai, C.-W.; Sillanpää, M. Enhanced photocatalytic activity of anatase-TiO2 nanoparticles by fullerene modification: A theoretical and experimental study. Appl. Surf. Sci. 2016, 387, 750–758. [Google Scholar] [CrossRef]
  55. 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, 012017. [Google Scholar] [CrossRef]
  56. 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, 1–14. [Google Scholar] [CrossRef]
  57. Pastore, M.; De Angelis, F. Computational modelling of TiO2 surfaces sensitized by organic dyes with different anchoring groups: Adsorption modes, electronic structure and implication for electron injection/recombination. Phys. Chem. Chem. Phys. 2012, 14, 920–928. [Google Scholar] [CrossRef]
  58. 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]
  59. Cramer, C.J.; Truhlar, D. Density functional theory for transition metals and transition metal chemistry. Phys. Chem. Chem. Phys. 2009, 11, 10757–10816. [Google Scholar] [CrossRef]
  60. Wang, J.; Wang, D.; Zhang, X.; Zhao, C.; Zhang, M.; Zhang, Z.; Wang, J. An anti-symmetric dual (ASD) Z-scheme photocatalytic system: (ZnIn2S4/Er3+:Y3Al5O12@ZnTiO3/CaIn2S4) for organic pollutants degradation with simultaneous hydrogen evolution. Int. J. Hydrogen Energy 2019, 44, 6592–6607. [Google Scholar] [CrossRef]
  61. Wang, Y.-W.; Yuan, P.-H.; Fan, C.-M.; Wang, Y.; Ding, G.-Y.; Wang, Y.-F. Preparation of zinc titanate nanoparticles and their photocatalytic behaviors in the photodegradation of humic acid in water. Ceram. Int. 2012, 38, 4173–4180. [Google Scholar] [CrossRef]
  62. 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]
  63. McLeod, J.A.; Moewes, A.; Zatsepin, D.A.; Kurmaev, E.Z.; Wypych-Puszkarz, A.; Bobowska, I.; Opasinska, A.; Cholakh, S.O. Predicting the band gap of ternary oxides containing 3d10 and 3d0 metals. Phys. Rev. B Condens. Matter Mater. Phys. 2012, 86, 195207. [Google Scholar] [CrossRef] [Green Version]
  64. Zhu, T.; Gao, S.-P. The stability, electronic structure, and optical property of tio 2 polymorphs. J. Phys. Chem. C 2014, 118, 11385–11396. [Google Scholar] [CrossRef] [Green Version]
  65. Hossain, F.M.; Evteev, A.V.; Belova, I.V.; Nowotny, J.; Murch, G.E. First-principles calculations of a corrugated anatase TiO2 surface. Comput. Mater. Sci. 2012, 51, 78–82. [Google Scholar] [CrossRef]
  66. 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]
  67. Zhu, H.-C.; Li, C.-F.; Fu, Z.-H.; Wei, S.-S.; Zhu, X.-F.; Zhang, J. Increasing the open-circuit voltage and adsorption stability of squaraine dye binding onto the TiO2 anatase (1 0 1) surface via heterocyclic anchoring groups used for DSSC. Appl. Surf. Sci. 2018, 455, 1095–1105. [Google Scholar] [CrossRef]
  68. Greathouse, J.A.; Geatches, D.L.; Pike, D.Q.; Greenwell, H.; Johnston, C.T.; Wilcox, J.; Cygan, R.T. Methylene blue adsorption on the basal surfaces of kaolinite: Structure and thermodynamics from quantum and classical molecular simulation. Clays Clay Miner. 2015, 63, 185–198. [Google Scholar] [CrossRef]
  69. Zhong, L.; Hu, Y.; Xing, D. Adsorption orientation of methylene blue (MB+) on the silver colloid: SERS and DFT studies. In Proceedings of the 2009 Conference on Lasers & Electro Optics & The Pacific Rim Conference on Lasers and Electro-Optics, Shanghai, China, 30 August–3 September 2009; pp. 1–2. [Google Scholar] [CrossRef]
  70. Orellana, W. D-π-A dye attached on TiO2(101) and TiO2(001) surfaces: Electron transfer properties from ab initio calculations. Sol. Energy 2021, 216, 266–273. [Google Scholar] [CrossRef]
  71. Qin, H.-C.; Qin, Q.-Q.; Luo, H.; Wei, W.; Liu, L.-X.; Li, L.-C. Theoretical study on adsorption characteristics and environmental effects of dimetridazole on TiO2 surface. Comput. Theor. Chem. 2019, 1150, 10–17. [Google Scholar] [CrossRef]
  72. Kuganathan, N.; Chroneos, A. Hydrogen adsorption on Ru-encapsulated, -doped and -supported surfaces of C60. Surfaces 2020, 3, 30. [Google Scholar] [CrossRef]
  73. Mathew, S.; Yella, A.; Gao, P.; Humphry-Baker, R.; Curchod, B.F.E.; Ashari-Astani, N.; Tavernelli, I.; Rothlisberger, U.; Nazeeruddin, K.; Graetzel, M. Dye-sensitized solar cells with 13% efficiency achieved through the molecular engineering of porphyrin sensitizers. Nat. Chem. 2014, 6, 242–247. [Google Scholar] [CrossRef] [Green Version]
  74. 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]
  75. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  76. Kohn, W.; Sham, L. Quantum density oscillations in an inhomogeneous electron gas. Phys. Rev. 1965, 137, A1697–A1705. [Google Scholar] [CrossRef]
  77. Wang, Y.A.; Xiang, P. From the Hohenberg-Kohn theory to the Kohn-Sham equations. In Recent Progress in Orbital-free Density Functional Theory; World Scientific: Singapore, 2013; pp. 3–12. [Google Scholar]
  78. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  79. Voityuk, A.A.; Stasyuk, A.J.; Vyboishchikov, S.F. A simple model for calculating atomic charges in molecules. Phys. Chem. Chem. Phys. 2018, 20, 23328–23337. [Google Scholar] [CrossRef]
  80. Yu, M.; Trinkle, D.R. Accurate and efficient algorithm for Bader charge integration. J. Chem. Phys. 2011, 134, 064111. [Google Scholar] [CrossRef] [Green Version]
  81. 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]
  82. Martinez, J.; Sinnott, S.B.; Phillpot, S.R. Adhesion and diffusion at TiN/TiO2 interfaces: A first principles study. Comput. Mater. Sci. 2017, 130, 249–256. [Google Scholar] [CrossRef] [Green Version]
  83. 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]
  84. Pillai, R.S.; Khan, I.; Titus, E. C2-Hydrocarbon Adsorption in Nano-porous Faujasite: A DFT Study. Mater. Today Proc. 2015, 2, 436–445. [Google Scholar] [CrossRef]
Figure 1. Optimized structures of (a) ZnTiO3 and (b) TiO2.
Figure 1. Optimized structures of (a) ZnTiO3 and (b) TiO2.
Molecules 26 03780 g001
Figure 2. Band structures of (a) ZnTiO3 and (b) TiO2 along the high symmetry directions in the Brillouin zone.
Figure 2. Band structures of (a) ZnTiO3 and (b) TiO2 along the high symmetry directions in the Brillouin zone.
Molecules 26 03780 g002
Figure 3. Density of states (DOSs) of ZnTiO3: (a) total, and partial: (b) Ti, (c) Zn and (d) O.
Figure 3. Density of states (DOSs) of ZnTiO3: (a) total, and partial: (b) Ti, (c) Zn and (d) O.
Molecules 26 03780 g003aMolecules 26 03780 g003b
Figure 4. Density of states (DOSs) of TiO2: (a) total, and partial: (b) Ti and (c) O.
Figure 4. Density of states (DOSs) of TiO2: (a) total, and partial: (b) Ti and (c) O.
Molecules 26 03780 g004
Figure 5. Methylene blue (MB) molecule in (a) horizontal and (b) semi-perpendicular orientation on the ZnTiO3 surface, and (c) semi-perpendicular orientation on the TiO2 surfaces.
Figure 5. Methylene blue (MB) molecule in (a) horizontal and (b) semi-perpendicular orientation on the ZnTiO3 surface, and (c) semi-perpendicular orientation on the TiO2 surfaces.
Molecules 26 03780 g005
Figure 6. Anchoring modes of the MB molecule on the surface of (a) ZnTiO3 and (b) TiO2.
Figure 6. Anchoring modes of the MB molecule on the surface of (a) ZnTiO3 and (b) TiO2.
Molecules 26 03780 g006
Table 1. Calculated bandgap energy of ZnTiO3 and TiO2 and other values reported in the literature.
Table 1. Calculated bandgap energy of ZnTiO3 and TiO2 and other values reported in the literature.
AdsorbentSoftware UsedBasis Set Used/Functional UsedBandgap (eV)Reference
ZnTiO3CASTEPGGA/SP-PBE3.14[1]
ZnTiO3CASTEPGGA+U3.28[1]
ZnTiO3MS-DMol3GGA/PBE3.10[5]
ZnTiO3MS-DMol3GGA/PPE-grime3.53[5]
ZnTiO3MS-DMol3GGA/PPE-TS3.12[5]
ZnTiO3Experimental3.18[15]
ZnTiO3VASPGGA/PBE2.96[15]
ZnTiO3CASTEPGGA/PW913.47[18]
ZnTiO3ABINITHSE064.25[56]
ZnTiO3ABINITGGA/NC3.25[56]
ZnTiO3ABINITLDA/NC3.05[56]
ZnTiO3ABINITGGA/ultrasoft2.96[56]
ZnTiO3ABINITLDA/ultrasoft2.86[56]
ZnTiO3VASPGGA/PBE+U3.16This study
ZnTiO3VASPGGA/PBE2.20This study
TiO2 VASPHSE063.20[62]
TiO2 VASPGGA/PBE2.55[58]
TiO2 VASPGGA/PBE+U3.11[58]
TiO2 Experimental3.20[53]
TiO2CASTEPGGA/PBE2.70[53]
TiO2 CASTEPGGA/PBE+U3.34[53]
TiO2 ABINITGGA/PBE2.08[64]
TiO2ABINITGW3.71[64]
TiO2 VASPGGA/PBE2.31This study
TiO2VASPGGA/PBE+U3.21This study
Table 2. Calculated adsorption energy of ZnTiO3 and TiO2 and other values reported in the literature.
Table 2. Calculated adsorption energy of ZnTiO3 and TiO2 and other values reported in the literature.
AdsorbentDyeSoftware UsedBasis Set/Functional UsedAdsorptionReferences
eVkJ/mol
ZnTiO3 (101)TPA-1CASTEPGGA/PBE−1.41−136.39[66]
ZnTiO3 (101)TPA-2CASTEPGGA/PBE−1.63−157.47[66]
ZnTiO3 (101)TPA-3CASTEPGGA/PBE−5.82−561.33[66]
ZnTiO3 (101)TPA-4CASTEPGGA/PBE−2.37−228.19[66]
ZnTiO3 (101) (H)MBVASPGGA/PBE−1.31−126.76This study
ZnTiO3 (101) (SP)MBVASPGGA/PBE−2.92−282.05This study
TiO2 (101)R4-BTVASPGGA/PBE−1.40−135.46[58]
TiO2 (101)R4-F2BTVASPGGA/PBE−1.39−134.50[58]
TiO2 (101)R4-BOVASPGGA/PBE−1.39−134.50[58]
TiO2 (101)R6-BzVASPGGA/PBE−1.40−135.46[58]
TiO2 (101)R6-BTVASPGGA/PBE−1.38−133.53[58]
TiO2 (101)R6-F2BTVASPGGA/PBE−1.37−132.56[58]
TiO2 (101)R6-B0VASPGGA/PBE−1.37−132.56[58]
TiO2 (101)R6-BzVASPGGA/PBE−1.38−133.53[58]
TiO2 (101) (SP)MBVASPGGA/PBE−0.12−11.61This 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.; Capa, L.F.; Medina, F.; González, S. DFT Study of Methylene Blue Adsorption on ZnTiO3 and TiO2 Surfaces (101). Molecules 2021, 26, 3780. https://doi.org/10.3390/molecules26133780

AMA Style

Jaramillo-Fierro X, Capa LF, Medina F, González S. DFT Study of Methylene Blue Adsorption on ZnTiO3 and TiO2 Surfaces (101). Molecules. 2021; 26(13):3780. https://doi.org/10.3390/molecules26133780

Chicago/Turabian Style

Jaramillo-Fierro, Ximena, Luis Fernando Capa, Francesc Medina, and Silvia González. 2021. "DFT Study of Methylene Blue Adsorption on ZnTiO3 and TiO2 Surfaces (101)" Molecules 26, no. 13: 3780. https://doi.org/10.3390/molecules26133780

Article Metrics

Back to TopTop