Next Article in Journal
Ni-Doped La0.6Sr0.4CoO3 Perovskite as an Efficient Electrocatalyst for Oxygen Reduction and Evolution Reactions in Alkaline Media
Previous Article in Journal
Zn-Cr Layered Double Hydroxides for Photocatalytic Transformation of CO2 under Visible Light Irradiation: The Effect of the Metal Ratio and Interlayer Anion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New BaTi0.96Cu0.02X0.02O3 (X = V, Nb) Photocatalysts for Dyes Effluent Remediation: Broad Visible Light Response

by
Ghayah M. Alsulaim
Department of Chemistry, College of Science, King Faisal University, Al-Ahsa 31982, Saudi Arabia
Catalysts 2023, 13(10), 1365; https://doi.org/10.3390/catal13101365
Submission received: 28 August 2023 / Revised: 27 September 2023 / Accepted: 10 October 2023 / Published: 12 October 2023
(This article belongs to the Section Photocatalysis)

Abstract

:
The problem of industrial dyes depollution has pushed the scientific research community to identify novel photocatalysts with high performance. Herein, new photocatalysts composed of BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 powders were prepared by solid-state reaction. The structural analysis of the samples confirmed the formation of the BaTiO3 structure. The splitting of (002) and (200) planes verified the formation of the tetragonal phase. The XRD peaks shifted, and the unit cell volume expansion verified the substitution of the Ti4+ site by Cu2+, V4+ and Nb5+ ions. The morphological measurements showed that the addition of (Cu, V) and (Cu, Nb) ions changes the particles’ morphology of BaTiO3, reducing its grains size. After the incorporation of (Cu, V) and (Cu, Nb) ions, the band gap of BaTiO3 was reduced from 3.2 to 2.84 and 2.72 eV, respectively. The modification of BaTiO3 by (Cu, Nb) ions induced superior photocatalytic properties for methyl green and methyl orange with degradation efficiencies of 97% and 94% during 60 and 90 min under sunlight irradiation, respectively. The total organic carbon results indicated that the BaTi0.96Cu0.02Nb0.02O3 catalyst has a high mineralization efficiency. In addition, it possesses a high stability during three cycles. The high photodegradation efficiency of Bi0.96La0.02Gd0.02FeO3 was related to the wide-ranging visible light absorption.

Graphical Abstract

1. Introduction

Human, animals, plants and aquatic life need a clean environment and pure water supply. Many industrial activities discharge large quantities of organic dyes pollutants into water resources [1,2]. These dyes are non-biodegradable, very hazardous and carcinogenic and can cause harm to all organisms even at low concentrations [3]. The paper, textile, paint, nylon, silk, cosmetics, plastics and leather industries are the main sources of dyes waste with nearly 200,000 tons of these organic dyes being dumped into water sources annually [4]. Photocatalysis is a potential green method to solve the dyes pollution problem through using a light energy source (UV or visible light) and active photocatalysts [1,2,3,4,5]. Hydrogen production by the photocatalytic water-splitting technique has been considered as a promising methodology to produce renewable energy [6]. In the photocatalysis process, a suitable light energy was used to activate the particles of the catalysts to produce charge carriers (electron–hole pairs) which later yielded very reactive radicals such as O2•− and ·OH species [4,7]. These reactive radicals can attack the organic dyes molecules and mineralize them to non-harmful products (CO2 and H2O) [4,5,6,7,8]. The dyes depollution issue pushes the scientific research community toward the synthesis of low-cost, stable, non-toxic and highly efficient photocatalysts [9,10,11]. Numerous photocatalytic studies have been carried out on metal oxides, sulfides and perovskite materials like ZnO [12], TiO2 [13], CuS [14] ZnS [15], Fe2O3 [16], WO3 [5], g-C3N4 [17] and BaTiO3 [18]. BaTiO3-based materials were candidates as promising catalysts for the photodegradation of organic contaminants and bacterial disinfection in wastewater [19,20]. The perovskite BaTiO3 semiconductor has a wide band gap energy varying from 3.2 to 3.4 eV [21]. Engineering the band gap energy of the BaTiO3 material to absorb the visible light spectrum is practically useful for the photocatalytic applications [22]. Bhat et al. [22] studied the effect of Rh doping on the photocatalytic activity of BaTiO3 material for the degradation of methylene blue dye (10 mg/L) under visible light irradiation. They found that 0.5 mol% Rh doping enhances the photocatalytic activity of BaTiO3 toward methylene blue dye with a photodegradation efficiency of 96% during 120 min of irradiation. The Mn-doped BaTiO3 nanotube as a photocatalyst exhibited improved visible light photocatalytic activity for methylene blue dye over 300 min [23]. The photodegradation properties of the 5% Ag-doped BaTiO3 catalyst was investigated for 20 mg/L rhodamine B dye under visible light irradiation [24]. After 105 min, this catalyst has shown a visible light photodegradation efficiency of 79%. Cu-doped BaTiO3 revealed photocatalytic activities of 98.2% for 10 mg/L methylene blue dye during 120 min and 99.4% for 10 mg/L rose Bengal dye within 45 min [18]. Amaechi et al. [25] showed that the 2% Fe3+-doped BaTiO3 nanocatalyst has a sunlight photodegradation efficiency of 75% for 20 mg/L methyl orange. Based on our knowledge, there is no available study on the influence of (Cu, V) and (Cu, Nb) codoping on the photodegradation properties of BaTiO3 for methyl green and methyl orange pollutants under natural sunlight and Xenon lamp photoreactor irradiation. The Cu, V and Nb elements as dopants have the ability to engineer the visible light response of BaTiO3. These dopants possess variable oxidation states which can inhibit the recombination of the charge carriers (electron–hole pairs) through acting as trap centers. Herein, new catalysts composed of BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 powders were prepared by the solid-state reaction method. The structural, morphological and optical properties of the synthesized samples were investigated and discussed. In addition, the photocatalytic properties of BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 powders for the degradation of methyl green and methyl orange dyes were measured under natural sunlight and Xenon lamp photoreactor irradiation. The obtained results indicated that the BaTi0.96Cu0.02Nb0.02O3 catalyst possesses high sunlight photodegradation and mineralization efficiencies for 20 mg/L methyl green and methyl orange dyes after 60 and 90 min, respectively. In addition, the BaTi0.96Cu0.02Nb0.02O3 catalyst exhibits a high photodegradation stability for both dyes during three cycles under sunlight.

2. Results and Discussion

2.1. X-ray Diffraction (XRD) Analysis

Figure 1 demonstrates the X-ray diffraction patterns of the pure BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 powders within a 2-theta angle of 10–80°. The pure BaTiO3 sample exhibits nine X-ray diffraction peaks at 2θ values equal to 22.401°, 31.722°, 39.101°, 44.69°, 45.421°, 51.112°, 56.299°, 66.031° and 75.102°, which were indexed to (100), (110), (111), (002), (200), (201), (211), (220) and (310) crystallographic planes of a tetragonal BaTiO3 structure (standard JCPDS, card file No. 05-0626, P4/mmm space group). The compositions of BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 powders display analogous X-ray diffraction peaks to that of the pure BaTiO3 sample, indicating also the formation of a tetragonal BaTiO3 structure. For all samples, the splitting of (002) and (200) planes within 2θ = 44–46° (inset Figure 1) verifies the formation of the tetragonal phase and ruled out the existence of a cubic BaTiO3 phase (both planes combine in one peak in case of a cubic BaTiO3 structure) [26]. As presented in Table 1, the ratio of (c/a) supported the formation of the tetragonal phase for all samples. The absence of any impurities can be deduced from the non-existence of any other X-ray diffraction peaks in these patterns. The X-ray diffraction peaks of all samples have high intensities, which reflect the good crystallinity. The influence of the incorporation of Cu, (Cu, V) and (Cu, Nb) ions on the 2θ angle of the (110) and (111) planes of the BaTiO3 structure is illustrated in Figure 2. It can be seen that both planes of BaTiO3 were shifted to a lower 2θ angle owing to doping and dual doping by Cu, (Cu, V) and (Cu, Nb) ions. The ionic radius of the Ba2+ site is 1.61 Å in the XII coordination, while that of the Ti4+ site is 0.605 Å In the VI coordination. The ionic radii of Cu2+, V4+ and Nb5+ ions in the VI coordination are 0.73 Å, 0.58 Å and 0.64 Å, respectively. The detected shifts to a lower 2θ angle suggested the replacement of the Ti4+ site (0.605 Å) by Cu2+, V4+ and Nb5+ ions of larger or close ionic radii. As shown in Figure 3 and Table 1, the calculations of the unit cell volume illustrate the expansion of the lattice parameters and unit cell volume of pure BaTiO3 after the incorporation of Cu2+, V4+ and Nb5+ ions, confirming the effective substitution process. These increases in the unit cell volume of BaTiO3 can be assigned to the larger ionic radii, especially Cu2+ and Nb5+ ions compared to Ti4+ site (0.605 Å). The average crystallite size of the BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 samples were estimated by using the Williamson Hall equation as shown below [27]:
βcosθ = (Kλ/D) + 4εsinθ
where β represents the full width at half maximum (radian), θ is the angle of the peaks in radians, K is a constant (0.9), λ is the wavelength of the X-ray radiation, D is the crystallite size and ε is the microstrain. By plotting 4sinθ (X-axis) with βcosθ (Y-axis), the average crystallite size can be estimated from the intercept of the fitting line, while the microstrain can be found from the slope. As illustrated in Figure 4, the average crystallite size of pure BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 powders were estimated to be 72, 62, 58 and 56 nm, respectively.

2.2. Scanning Electron Microscope (SEM) Analysis

The scanning electron microscope (SEM) images of the synthesized pure BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 powders and the effect of dopants on morphology are shown in Figure 5. The SEM image of the pure BaTiO3 sample reveals the synthesis of coagulated large grains, which possess an asymmetrical shape with an irregular edge. The Cu doping into the BaTiO3 lattice somewhat reduces the size of the grains, but they are still coagulated together. In contrast, the addition of the dual dopants of (Cu, V) and (Cu, Nb) ions powerfully changes the morphology of BaTiO3 and leads to the formation of uniform semispherical particles with reduced grain size. These results point out that (Cu, V) and (Cu, Nb) ions have a strong restriction effect on the growth of the BaTiO3 grains during the calcination process at high temperature. The energy-dispersive X-ray (EDX) investigation was carried out to analyze the elemental composition of the synthesized samples, as demonstrated in Figure 6. The patterns of the energy-dispersive X-ray spectra of BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 powders verify the presence of Ba, Ti, O, Cu, V and Nb elements and ruled out the existence of any other impurity elements. The wt. % of the dopants elements, including Cu, V and Nb ions, is close with the added values during the synthesis process.

2.3. Optical Properties and Band Gap Energy

The effective photocatalysis process is strongly related to the optical properties and the band gap energy of the used catalysts. The diffuse reflectance (DR) analysis was carried out to investigate the optical properties of the pure BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 catalysts within a wavelength of 200–2400 nm, as shown in Figure 7. The reflectance spectrum (%) of the pure BaTiO3 catalyst has a high intensity (80–90%) within 400–2400 nm, while below 400 nm, the reflectance (%) intensely decreased in a sharp manner (absorption edge), which matches with the band gap energy absorption of the pure BaTiO3 structure. The addition of Cu2+, V4+ and Nb5+ ions strongly reduces the intensity of the reflectance and also shifts the absorption edge to a longer wavelength (low energy). The diffuse reflectance spectra display that Cu, (Cu, V) and (Cu, Nb) ions enhance the visible light absorption capacity of the BaTiO3 catalyst. The Tauc and Kubelka–Munk equations were used to estimate the band gap energy of pure BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 catalysts as displayed below [28]:
α = A(hν − Eg)n/hν
F (R) = (1 − R)2/2R = α/S
In both equations, the α represents the absorption coefficient, hν is the photon energy, Eg is the band gap energy, F (R) is the Kubelka–Munk function, R is the reflectance and S is the scattering coefficient. As represented in Figure 8, the plot of [F(R) hν]2 with hν (photon energy) gives the band gap energy of the BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 samples. The measured value of the energy band gap of the undoped BaTiO3 catalyst was identified to be 3.2 eV, which is in agreement with the published studies [29,30]. The valence band maximum (VBM) of the BaTiO3 structure is principally composed of Ti 3d and O 2p states, while the conduction band minimum (CBM) basically consists of Ti 3d states [30]. Owing to the incorporation of Cu, (Cu, V) and (Cu, Nb) ions, the band gap energy of the BaTiO3 structure was reduced to 2.99 eV (λ = 414 nm), 2.84 eV (λ = 436 nm) and 2.72 eV (λ = 455 nm), respectively. The results of the optical properties prove that the dopants-treated BaTiO3 catalysts have high visible light absorption compared to the pure one. Owing to the insertion of Cu2+, V4+ and Nb5+ ions, the conduction band (CB) is possibly derived from Cu-3d, V-3d or Nb-4d and Ti-3d electrons, while the valence band (VB) mostly arises from Cu-3d, V-3d or Nb-4d plus O-2p and Ti-3d states. As a result, the intense reductions of the band gap energy of BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 catalysts can be assigned to the presence of impurity states below the conduction band (CB) and above the valence band (VB) due to Cu2+, V4+ and Nb5+ ions and the formation of oxygen vacancies for charge neutrality [18,30,31]. The sunlight spectrum has nearly 52% infrared red radiation (700–2500 nm), 43% visible light rays (400–700 nm) and 5% ultraviolet wavelengths (280–400 nm) [32]. The optical characteristics prove the suitability of the engineered band gap BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 catalysts for photocatalytic applications under visible light irradiation. In addition, the presence of long absorption states (as illustrated in Figure 8) of BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 catalysts support the wide-ranging visible light absorption of the treated samples.

2.4. Depollution Properties of the Photocatalysts

The photocatalytic depollution properties of BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 catalysts toward polluted water containing methyl green and methyl orange were studied under natural sunlight and Xenon lamp photoreactor irradiation. Figure 9 shows the change in absorption spectra of methyl green (20 mg/L, catalyst dose 40 mg/100 mL) in the presence of pure BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 catalysts during 60 min of natural sunlight irradiation. In a sequenced manner, the pure BaTiO3 catalyst displays low decreases in the absorbance with time with a measured efficiency of 33% during this period. The reductions of the absorbance curve are more noticeable for the BaTi0.96Cu0.04O3 catalyst, and the photodegradation activity value was estimated to be 67% in the same period. In case of BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 catalysts, the absorbance curve of methyl green dye was powerfully decreased with full degradation activity of 93% and 97%, respectively. The degradation performance of these catalysts was completely assigned to the photocatalytic process, since the adsorption capacity of all samples was nearly lower than 5% as illustrated in Figure 10. For kinetic study, the rate constant of the photodegradation reactions was estimated using the non-linear least squares analysis according to [1,2]:
A = X e K t + E
where A represents the polluted molecules calculated residual, X indicates the amplitude of the process, K symbolizes the pseudo-first-order rate constant and E signifies the endpoint. As shown in Figure 11, the measured photodegradation rate constants based on non-linear least squares analysis were 0.0065 (R2 = 0.98), 0.016 (R2 = 0.98), 0.03 (R2 = 0.96) and 0.035 min−1 (R2 = 0.96) for BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 catalysts, respectively.
Figure 12a–d show the measurements of the reusability performance, efficient reactive radicals, total organic carbon (TOC) and photodegradation efficiency of methyl green under Xenon lamp irradiation compared to natural sunlight irradiation for BaTi0.96Cu0.02Nb0.02O3 catalyst, respectively. In the reusability test, as shown in Figure 12a, during three repeated cycles, the BaTi0.96Cu0.02Nb0.02O3 catalyst revealed high efficiencies of 97, 95 and 94% toward 20 ppm methyl green solution. The high photodegradation activity of the BaTi0.96Cu0.02Nb0.02O3 catalyst reflects the high stability for multiple uses. To analyze the reactive radicals which have the main effects during the photodegradation reaction, four trapping agents [33] including silver nitrate (AgNO3), isopropyl alcohol (IPA), triethanolamine (TEOA) and benzoquinone (BQ) were used to capture the electrons (e), hydroxyl (·OH) radicals, holes (h+) and superoxide (O2•–) species, respectively, as clarified in Figure 12b. The photodegradation activity of the BaTi0.96Cu0.02Nb0.02O3 catalyst against 20 mg/L methyl green dye was obviously decreased after the addition of TEOA and BQ substances which verifies that the holes (h+) and superoxide (O2•–) species are the basic radicals in the photo-removal process. Figure 12c displays the measurement of the total organic carbon (TOC) analysis of 20 mg/L methyl green dye for BaTi0.96Cu0.02Nb0.02O3 catalyst under sunlight irradiation. The total organic carbon analysis was carried out to investigate the mineralization behavior of methyl green dye to CO2 and H2O through the photocatalysis process. The obtained results prove that the rate of the total organic carbon was decreased with the sunlight irradiation time, reaching a photo-removal value of 94% after 70 min, demonstrating the complete photodegradation of methyl green dye using the BaTi0.96Cu0.02Nb0.02O3 catalyst to CO2 and H2O (simple products). As a result, these findings point out that the BaTi0.96Cu0.02Nb0.02O3 catalyst has a high mineralization efficiency beside the decolorization reaction. Figure 12d shows the photodegradation activity of the BaTi0.96Cu0.02Nb0.02O3 catalyst toward 20 mg/L methyl green dye under direct sunlight and Xenon lamp irradiation. Under both irradiation sources, the photodegradation activity of the BaTi0.96Cu0.02Nb0.02O3 catalyst was 97 and 99%, indicating the robust photocatalytic characteristics under the free solar energy to remove the methyl green dye in 60 min.
The photodegradation of methyl orange pollutant (MO) as an anionic organic dye was investigated using the BaTi0.96Cu0.02Nb0.02O3 catalyst (best sample) under natural sunlight irradiation. Figure 13 illustrates the change in the absorption spectra of methyl orange dye (concentration = 20 mg/L, catalyst dose = 40 mg/100 mL) in the presence of the BaTi0.96Cu0.02Nb0.02O3 catalyst during 90 min of sunlight illumination. Gradual and notable decreases of the absorption curve of methyl orange dye with time were seen to reach the final photodegradation efficiency of 94%. As shown in Figure 14a, the plot of Ct/C0 with time for methyl orange dye degradation decreased with the sunlight irradiation. The kinetic plot of the BaTi0.96Cu0.02Nb0.02O3 catalyst for methyl orange dye reveals that the rate constant was 0.028 min−1, and the degradation reaction obeys the pseudo-first-order kinetic model with an R-square value of 0.92, as shown in Figure 14b. The stability or the reusability test of the BaTi0.96Cu0.02Nb0.02O3 catalyst for 20 mg/L of methyl orange dye during three cycles proves the high photodegradation properties of this composition with final efficiencies of 94, 91 and 88%, as depicted in Figure 14c. The photocatalytic results of both dyes (methyl green and methyl orange) prove the superior photodegradation features of the BaTi0.96Cu0.02Nb0.02O3 catalyst for practical uses under natural and free sunlight energy.
The photocatalytic reaction or mechanism for the degradation of the tested organic dyes (methyl green and methyl orange) of the Bi0.96La0.02Gd0.02FeO3 catalyst essentially depends on the photo-excitation of the electrons from the top of the valence band (VB) to the bottom of the conduction band (CB) by sunlight energy. The photogenerated electrons in the conduction band (CB) attack the oxygen molecules present in solution to form the superoxide (O2•–) species, whereas the positive holes (h+) linked to the water molecules to yield the hydroxyl (·OH) radicals, respectively. The strong interaction between ·OH and O2•– radicals and methyl green or methyl orange produces CO2 and H2O, as shown below [3,4,5]:
BaTi0.96Cu0.02Nb0.02O3 (best catalyst) → natural sunlight irradiation → e+ h+ pairs
Cu2+ + e → Cu+
Nb5+ + e → Nb4+
e + O2 → O2•–
h+ + H2O → ·OH + H+
(·OH, O2•–) → Methyl green and methyl orange → CO2 and H2O (non-harmful products)
The remarkable photodegradation efficiency of the BaTi0.96Cu0.02Nb0.02O3 catalyst can be related mainly to three factors including the high visible light absorption capacity, homogenous morphology of the grains and the efficient dual dopants separation of the charge carriers (electron–hole pairs).
Figure 15 depicts the X-ray diffraction patterns of the BaTi0.96Cu0.02Nb0.02O3 catalyst before and after the photodegradation test for methyl green dye. The X-ray diffraction peaks of both powders have a good relative intensity and indexed to the tetragonal BaTiO3 structure, confirming the stability of the BaTi0.96Cu0.02Nb0.02O3 catalyst after the photocatalytic degradation reaction. Ahamed et al. [34] reported that Zn-doped BaTiO3 exhibits a visible light photodegradation efficiency of 85% for 20 mg/L methylene blue in 80 min. Bhat et al. [22] showed that Rh-doped BaTiO3 has a photoactivity of 96% for 10 ppm methylene blue in 120 min. The Mn-doped BaTiO3 nanotube as a photocatalyst exhibited improved visible light photocatalytic activity for methylene blue dye during 300 min [23]. Ag-doped BaTiO3 [24] possesses a visible light photodegradation efficiency of 79% for 20 mg/L rhodamine B. Amaechi et al. [25] showed that Fe-doped BaTiO3 has a sunlight photodegradation efficiency of 75% for 20 mg/L methyl orange dye. In our study, the BaTi0.96Cu0.02Nb0.02O3 catalyst exhibited a photodegradation efficiency of 97% and 94% during 60 and 90 min under natural sunlight, respectively. This comparison illustrates the remarkable photocatalytic efficiency of the BaTi0.96Cu0.02Nb0.02O3 catalyst.

3. Preparation of Samples and Measurements

In this work, all materials were purchased from Sigma-Aldrich (Steinheim, Germany) and used as received. Barium nitrate (Ba(NO3)2, 99.6%), titanium dioxide (TiO2, 99.9%), copper(II) nitrate trihydrate (Cu(NO3)2·3H2O, 99.95%), ammonium monovanadate (NH₄VO₃, ≥99.0%) and niobium(V) chloride (NbCl5, 99%) substances were used to prepare the pure BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 powders by the solid-state reaction method. The used weights for each composition are summarized in Table 2. A pure BaTiO3 sample was prepared through good mixing of Ba(NO3)2 and TiO2 (molar ratio 1:1). Then, the obtained mixture was calcined at 600 °C for 4 h, and after that, the calcination temperature was elevated to 1250 °C for 5 h and left to cool to room temperature. The BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 powders were synthesized by the same method through mixing proper weights of the dopants materials (Table 2) with the Ba(NO3)2-TiO2 mixture. The X-ray diffraction analysis (PANalytical X-ray diffraction equipment model X′Pert PRO), scanning electron microscope (SEM, model Quanta 250 FEG) and double-beam spectrophotometer JASCO (JASCO, V-570 UV-Vis-NIR) were used to characterize the synthesized BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 compositions. The photocatalytic properties of pure BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 powders for the degradation of methyl green and methyl orange dyes were measured using a 100 mL solution of methyl green (20 mg/L) and methyl orange (20 mg/L) plus 40 mg of the catalyst. The effect of adsorption was firstly tested under dark conditions, and then, the mixed solution (catalyst + dye) was exposed to a Xenon lamp or direct sunlight illumination (June, 12–2 pm). Every ten minutes, 4 mL of the exposed solution was taken and centrifuged to eliminate the particles of the powder. The efficiency of the photodegradation process was calculated via measuring the change in absorbance of the maximum peak of methyl green (633 nm) and methyl orange (464 nm). The photodegradation efficiency (PE) is expressed as follows:
PE = At/A0 = Ct/C0,
where A0 and C0 signify the primary absorbance and concentration of methyl green and methyl orange, whereas At and Ct indicate the absorbance and concentration of methyl green and methyl orange during the different irradiation times.

4. Conclusions

In this study, a new BaTi0.96Cu0.02Nb0.02O3 photocatalyst exhibits a high depollution activity for 20 mg/L methyl green and methyl orange pollutants under natural sunlight and Xenon lamp photoreactor irradiations. The solid-state reaction method was used to synthesize high-purity BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 powders. The X-ray diffraction analysis of all samples proves the formation of a tetragonal perovskite BaTiO3 structure. The SEM images demonstrate that the addition of the dual dopants of (Cu, V) and (Cu, Nb) ions to BaTiO3 structure remarkably modifies the morphology of the particles and leads to the formation of uniform semispherical grains. The incorporation of Cu, (Cu, V) and (Cu, Nb) ions adjusts the band gap energy of the BaTiO3 catalyst to absorb the most visible light spectrum. BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 powders reveal a band gap energy of 3.2, 2.99, 2.84 and 2.72 eV with the formation of long absorption tails, respectively. The modification of BaTiO3 by (Cu, Nb) ions induced a superior photodegradation activity for 20 mg/L methyl green and methyl orange pollutants with a total efficiency of 97% and 94% during 60 and 90 min under natural sunlight, respectively. The total organic carbon analysis indicates that the tested dyes were effectively mineralized to CO2 and H2O in presence of BaTi0.96Cu0.02Nb0.02O3 catalyst. In addition, BaTi0.96Cu0.02Nb0.02O3 catalyst exhibits a high stability for both dyes during three cycles.

Funding

This research received no external funding.

Data Availability Statement

The datasets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.

Acknowledgments

The authors extend their appreciation to the Deanship of Scientific Research at King Faisal University for the logistic support of this work through Project No. (GRANT 4,195).

Conflicts of Interest

The author declares that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Belghiti, M.; Tanji, K.; El Mersly, L.; Lamsayety, I.; Ouzaouit, K.; Faqir, H.; Benzakour, I.; Rafqah, S.; Outzourhit, A. Fast and non-selective photodegradation of basic yellow 28, malachite green, tetracycline, and sulfamethazine using a nanosized ZnO synthesized from zinc ore. React. Kinet. Mech. Catal. 2022, 135, 2265–2278. [Google Scholar] [CrossRef]
  2. Dra, A.; Tanji, K.; Arrahli, A.; Iboustaten, E.; El Gaidoumi, A.; Kherchafi, A.; Benabdallah, A.C.; Kherbeche, A. Valorization of Oued Sebou natural sediments (Fez-Morocco Area) as adsorbent of methylene blue dye: Kinetic and thermodynamic study. Sci. World J. 2020, 2020, 2187129. [Google Scholar]
  3. Abid, M.Z.; Rafiq, K.; Rauf, A.; Shah, S.S.A.; Jin, R.; Hussain, E. Synergism of Co/Na in BiVO4 microstructures for visible-light driven degradation of toxic dyes in water. Nanoscale Adv. 2023, 5, 3247–3259. [Google Scholar] [CrossRef] [PubMed]
  4. Batra, V.; Kaur, I.; Pathania, D.; Sonu; Chaudhary, V. Efficient dye degradation strategies using green synthesized ZnO-based nanoplatforms: A review. Appl. Surf. Sci. Adv. 2022, 11, 100314. [Google Scholar] [CrossRef]
  5. Mzimela, N.; Tichapondwa, S.; Chirwa, E. Visible-light-activated photocatalytic degradation of rhodamine B using WO3 nanoparticles. RSC Adv. 2022, 12, 34652–34659. [Google Scholar] [CrossRef] [PubMed]
  6. Wu, Y.; Li, Y.; Hu, H.; Zeng, G.; Li, C. Recovering hydrogen energy from photocatalytic treatment of pharmaceutical-contaminated water using Co3O4 modified {001}/{101}-TiO2 nanosheets. ACS EST Eng. 2021, 1, 603–611. [Google Scholar] [CrossRef]
  7. Ihsan, A.; Irshad, A.; Warsi, M.F.; Din, M.I.; Zulfiqar, S. NiFe2O4/ZnO nanoparticles and its composite with flat 2D rGO sheets for efficient degradation of colored and colorless effluents photocatalytically. Opt. Mater. 2022, 134, 113213. [Google Scholar] [CrossRef]
  8. Ighnih, H.; Haounati, R.; Malekshah, R.E.; Ouachtak, H.; Jada, A.; Addi, A.A. Photocatalytic degradation of RhB dye using hybrid nanocomposite BiOCl@Kaol under sunlight irradiation. J. Water Process. Eng. 2023, 54, 103925. [Google Scholar] [CrossRef]
  9. Pandit, M.A.; Billakanti, S.; Muralidharan, K. A simplistic approach for the synthesis of CuS-CdS heterostructure: A novel photo catalyst for oxidative dye degradation. J. Environ. Chem. Eng. 2020, 8, 103542. [Google Scholar] [CrossRef]
  10. Rajan, K.D.; Srinivasan, D.; Gotipamul, P.P.; Khanna, S.; Chidambaram, S.; Rathinam, M. Design of a novel ZnBi2O4/Bi2O3 type-II photo-catalyst via short term hydrothermal for enhanced degradation of organic pollutants. Mater. Sci. Eng. B 2022, 285, 115929. [Google Scholar] [CrossRef]
  11. Darabdhara, J.; Roy, S.; Ahmaruzzaman, M. Efficient photocatalytic degradation of an organic dye by the fabrication of a novel ternary composite based on zeolitic imidazolate framework via a facile in-situ synthetic approach. Inorg. Chem. Commun. 2023, 152, 110694. [Google Scholar] [CrossRef]
  12. Beura, R.; Rajendran, S.; Pinilla, M.A.G.; Thangadurai, P. Enhanced photo-induced catalytic activity of Cu ion doped ZnO—Graphene ternary nanocomposite for degrading organic dyes. J. Water Process. Eng. 2019, 32, 100966. [Google Scholar] [CrossRef]
  13. Jouali, A.; Salhi, A.; Aguedach, A.; Aarfane, A.; Ghazzaf, H.; Lhadi, E.K.; El krati, M.; Tahiri, S. Photo-catalytic degradation of methylene blue and reactive blue 21 dyes in dynamic mode using TiO2 particles immobilized on cellulosic fibers. J. Photochem. Photobiol. A 2019, 383, 112013. [Google Scholar] [CrossRef]
  14. Borah, D.; Saikia, P.; Sarmah, P.; Gogoi, D.; Rout, J.; Ghosh, N.N.; Bhattacharjee, C.R. Composition controllable alga-mediated green synthesis of covellite CuS nanostructure: An efficient photocatalyst for degradation of toxic dye. Inorg. Chem. Commun. 2022, 142, 109608. [Google Scholar] [CrossRef]
  15. Samanta, D.; Basnet, P.; Jha, S.; Chatterjee, S. Proficient route in synthesis of glucose stabilized Ag modified ZnS nanospheres for mechanistic understandings of commercially used dyes degradation. Inorg. Chem. Commun. 2022, 141, 109498. [Google Scholar] [CrossRef]
  16. Araujo, R.N.; Nascimento, E.P.; Firmino, H.C.T.; Macedo, D.A.; Neves, G.A.; Morales, M.A.; Menezes, R.R. α-Fe2O3 fibers: An efficient photocatalyst for dye degradation under visible light. J. Alloys Compd. 2021, 882, 160683. [Google Scholar] [CrossRef]
  17. Ismail, A.A.; Faisal, M.; Al-Haddad, A. Mesoporous WO3-graphene photocatalyst for photocatalytic degradation of Methylene Blue dye under visible light illumination. J. Environ. Sci. 2018, 66, 328–337. [Google Scholar] [CrossRef]
  18. Uma, P.I.; Shenoy, U.S.; Bhat, D.K. Doped BaTiO3 cuboctahedral nanoparticles: Role of copper in photocatalytic degradation of dyes. Appl. Surf. Sci. Adv. 2023, 15, 100408. [Google Scholar] [CrossRef]
  19. Ray, S.K.; Cho, J.; Hur, J. A critical review on strategies for improving efficiency of BaTiO3-based photocatalysts for wastewater treatment. J. Environ. Manag. 2021, 290, 112679. [Google Scholar] [CrossRef]
  20. Masekela, D.; Hintsho-Mbita, N.C.; Sam, S.; Yusuf, T.L.; Mabuba, N. Application of BaTiO3-based catalysts for piezocatalytic, photocatalytic and piezophotocatalytic degradation of organic pollutants and bacterial disinfection in wastewater: A comprehensive review. Arab. J. Chem. 2023, 16, 104473. [Google Scholar] [CrossRef]
  21. Elmahgary, M.G.; Mahran, A.M.; Ganoub, M.; Abdellatif, S.O. Optical investigation and computational modelling of BaTiO3 for optoelectronic devices applications. Sci. Rep. 2023, 13, 4761. [Google Scholar] [CrossRef] [PubMed]
  22. Bhat, D.K.; Bantawala, H.; Shenoy, U.S. Rhodium doping augments photocatalytic activity of barium titanate: Effect of electronic structure engineering. Nanoscale Adv. 2020, 2, 5688–5698. [Google Scholar] [CrossRef] [PubMed]
  23. Nageri, M.; Kumar, V. Manganese-doped BaTiO3 nanotube arrays for enhanced visible light photocatalytic applications. Mater. Chem. Phys. 2018, 213, 400–405. [Google Scholar] [CrossRef]
  24. Khan, M.A.M.; Kumar, S.; Ahmed, J.; Ahamed, M.; Kumar, A. Influence of silver doping on the structure, optical and photocatalytic properties of Ag-doped BaTiO3 ceramics. Mater. Chem. Phys. 2021, 259, 124058. [Google Scholar] [CrossRef]
  25. Amaechi, I.C.; Youssef, A.H.; Kolhatkar, G.; Rawach, D.; Gomez-Yañez, C.; Claverie, J.P.; Sun, S.; Ruediger, A. Ultrafast microwave-assisted hydrothermal synthesis and photocatalytic behaviour of ferroelectric Fe3+-doped BaTiO3 nanoparticles under simulated sunlight. Catal. Today 2021, 360, 90–98. [Google Scholar] [CrossRef]
  26. Kim, J.-H.; Jung, W.-S.; Kim, H.-T.; Yoon, D.-H. Properties of BaTiO3 synthesized from barium titanyl oxalate. Ceram. Int. 2009, 35, 2337–2342. [Google Scholar] [CrossRef]
  27. Khan, M.; Mishra, A.; Shukla, J.; Sharma, P. X-ray analysis of BaTiO3 ceramics by Williamson-Hall and size strain plot methods. Nanoscale Res. Lett. 2021, 16, 115. [Google Scholar]
  28. Salem, B.B.; Essalah, G.; Ameur, S.B.; Duponchel, B.; Guermazi, H.; Guermazia, S.; Leroy, G. Synthesis and comparative study of the structural and optical properties of binary ZnO-based composites for environmental applications. RSC Adv. 2023, 13, 6287–6303. [Google Scholar] [CrossRef] [PubMed]
  29. Kappadan, S.; Thomas, S.; Kalarikkal, N. Enhanced photocatalytic performance of BaTiO3/g-C3N4 heterojunction for the degradation of organic pollutants. Chem. Phys. Lett. 2021, 771, 138513. [Google Scholar] [CrossRef]
  30. Yang, F.; Yang, L.; Ai, C.; Xie, P.; Lin, S.; Wang, C.-Z.; Lu, X. Tailoring bandgap of perovskite BaTiO3 by transition metals Co-doping for visible-light photoelectrical applications: A first-principles study. Nanomaterials 2018, 8, 455. [Google Scholar] [CrossRef] [PubMed]
  31. Patil, R.P.; Karanjekar, A.N.; Gaikwad, V.B.; Jain, G.H.; Shinde, S.D. Doping-induced diffused phase transition triggers gas-sensing performance of Sn-doped BaTiO3 nanostructures. Surf. Interface Anal. 2022, 54, 25–36. [Google Scholar] [CrossRef]
  32. Moriomoto, T.; Oka, R.; Minagawaa, K.; Masui, T. Novel near-infrared reflective black inorganic pigment based on cerium vanadate. RSC Adv. 2022, 12, 16570–16575. [Google Scholar] [CrossRef]
  33. Jiang, G.; Wei, Z.; Chen, H.; Du, X.; Li, L.; Liu, Y.; Huang, Q.; Chen, W. Preparation of novel carbon nanofibers with BiOBr and AgBr decoration for the photocatalytic degradation of rhodamine B. RSC Adv. 2015, 5, 30433–30437. [Google Scholar] [CrossRef]
  34. Ahamed, M.; Khan, M.A.M. Enhanced photocatalytic and anticancer activity of Zn-doped BaTiO3 nanoparticles prepared through a green approach using Banana Peel extract. Catalysts 2023, 13, 985. [Google Scholar] [CrossRef]
Figure 1. Crystal structure based on X-ray diffraction analysis of (a) BaTiO3, (b) BaTi0.96Cu0.04O3, (c) BaTi0.96Cu0.02V0.02O3 and (d) BaTi0.96Cu0.02Nb0.02O3 samples.
Figure 1. Crystal structure based on X-ray diffraction analysis of (a) BaTiO3, (b) BaTi0.96Cu0.04O3, (c) BaTi0.96Cu0.02V0.02O3 and (d) BaTi0.96Cu0.02Nb0.02O3 samples.
Catalysts 13 01365 g001
Figure 2. Impact of incorporation of Cu, (Cu, V) and (Cu, Nb) ions on 2θ angle of (110) and (111) crystallographic planes of BaTiO3 structure.
Figure 2. Impact of incorporation of Cu, (Cu, V) and (Cu, Nb) ions on 2θ angle of (110) and (111) crystallographic planes of BaTiO3 structure.
Catalysts 13 01365 g002
Figure 3. Impact of incorporation of Cu, (Cu, V) and (Cu, Nb) ions on unit cell volume of pure BaTiO3 sample.
Figure 3. Impact of incorporation of Cu, (Cu, V) and (Cu, Nb) ions on unit cell volume of pure BaTiO3 sample.
Catalysts 13 01365 g003
Figure 4. Williamson–Hall plots of BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 samples for crystallite size estimation.
Figure 4. Williamson–Hall plots of BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 samples for crystallite size estimation.
Catalysts 13 01365 g004
Figure 5. Scanning electron microscope images of (a) BaTiO3, (b) BaTi0.96Cu0.04O3, (c) BaTi0.96Cu0.02V0.02O3 and (d) BaTi0.96Cu0.02Nb0.02O3 powders.
Figure 5. Scanning electron microscope images of (a) BaTiO3, (b) BaTi0.96Cu0.04O3, (c) BaTi0.96Cu0.02V0.02O3 and (d) BaTi0.96Cu0.02Nb0.02O3 powders.
Catalysts 13 01365 g005
Figure 6. Energy-dispersive X-ray (EDX) spectra of (a) BaTi0.96Cu0.02V0.02O3 and (b) BaTi0.96Cu0.02Nb0.02O3 powders.
Figure 6. Energy-dispersive X-ray (EDX) spectra of (a) BaTi0.96Cu0.02V0.02O3 and (b) BaTi0.96Cu0.02Nb0.02O3 powders.
Catalysts 13 01365 g006
Figure 7. Diffuse reflectance (DR) spectra of BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 powders.
Figure 7. Diffuse reflectance (DR) spectra of BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 powders.
Catalysts 13 01365 g007
Figure 8. Estimation of band gap energy of BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 powders.
Figure 8. Estimation of band gap energy of BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 powders.
Catalysts 13 01365 g008
Figure 9. Absorbance curves of 20 ppm methyl green of (a) BaTiO3, (b) BaTi0.96Cu0.04O3, (c) BaTi0.96Cu0.02V0.02O3 and (d) BaTi0.96Cu0.02Nb0.02O3 powders during sunlight irradiation.
Figure 9. Absorbance curves of 20 ppm methyl green of (a) BaTiO3, (b) BaTi0.96Cu0.04O3, (c) BaTi0.96Cu0.02V0.02O3 and (d) BaTi0.96Cu0.02Nb0.02O3 powders during sunlight irradiation.
Catalysts 13 01365 g009
Figure 10. Plots of Ct/C0 with time (min) of BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 powders.
Figure 10. Plots of Ct/C0 with time (min) of BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 powders.
Catalysts 13 01365 g010
Figure 11. Displays the non-linear first-order fitting of all BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 samples.
Figure 11. Displays the non-linear first-order fitting of all BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 samples.
Catalysts 13 01365 g011
Figure 12. Shows the measurements of (a) reusability, (b) effective radicals, (c) total organic carbon and (d) the photodegradation efficiency under a Xenon lamp compared to sunlight irradiation for the BaTi0.96Cu0.02Nb0.02O3 catalyst.
Figure 12. Shows the measurements of (a) reusability, (b) effective radicals, (c) total organic carbon and (d) the photodegradation efficiency under a Xenon lamp compared to sunlight irradiation for the BaTi0.96Cu0.02Nb0.02O3 catalyst.
Catalysts 13 01365 g012
Figure 13. Absorbance curves of 20 mg/L methyl orange of BaTi0.96Cu0.02Nb0.02O3 powder during natural sunlight irradiation.
Figure 13. Absorbance curves of 20 mg/L methyl orange of BaTi0.96Cu0.02Nb0.02O3 powder during natural sunlight irradiation.
Catalysts 13 01365 g013
Figure 14. Illustrates (a) plot of Ct/C0 with time (min), (b) plot of ln (Ct/C0) with time (min) and (c) reusability test of 20 mg/L methyl orange for BaTi0.96Cu0.02Nb0.02O3 catalyst.
Figure 14. Illustrates (a) plot of Ct/C0 with time (min), (b) plot of ln (Ct/C0) with time (min) and (c) reusability test of 20 mg/L methyl orange for BaTi0.96Cu0.02Nb0.02O3 catalyst.
Catalysts 13 01365 g014
Figure 15. X-ray diffraction patterns of BaTi0.96Cu0.02Nb0.02O3 catalyst (a) before and (b) after the photocatalytic process.
Figure 15. X-ray diffraction patterns of BaTi0.96Cu0.02Nb0.02O3 catalyst (a) before and (b) after the photocatalytic process.
Catalysts 13 01365 g015
Table 1. Crystallite size D, lattice parameters (a, b, c), c/a ratio, unit cell volume (V) of pure BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 samples.
Table 1. Crystallite size D, lattice parameters (a, b, c), c/a ratio, unit cell volume (V) of pure BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 samples.
SamplesD
(nm)
a
Å
b
Å
c
Å
c/aV
Å3
BaTiO3713.99873.99874.01111.003164.1359
BaTi0.96Cu0.04O3614.00024.00024.01741.003864.2861
BaTi0.96Cu0.02V0.02O3584.00084.00084.01431.003364.2535
BaTi0.96Cu0.02Nb0.02O3564.00074.00074.01501.003564.2639
Table 2. Detailed weights of chemical substances used in the synthesis of BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 catalysts.
Table 2. Detailed weights of chemical substances used in the synthesis of BaTiO3, BaTi0.96Cu0.04O3, BaTi0.96Cu0.02V0.02O3 and BaTi0.96Cu0.02Nb0.02O3 catalysts.
CompositionBa(NO3)2
(g)
TiO2
(g)
Cu(NO3)2·3H2O
(g)
NH4VO3
(g)
NbCl5
(g)
BaTiO37.842.396---
BaTi0.96Cu0.04O37.842.300.29--
BaTi0.96Cu0.02V0.02O37.842.300.14490.07-
BaTi0.96Cu0.02Nb0.02O37.842.300.1449-0.162
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Alsulaim, G.M. New BaTi0.96Cu0.02X0.02O3 (X = V, Nb) Photocatalysts for Dyes Effluent Remediation: Broad Visible Light Response. Catalysts 2023, 13, 1365. https://doi.org/10.3390/catal13101365

AMA Style

Alsulaim GM. New BaTi0.96Cu0.02X0.02O3 (X = V, Nb) Photocatalysts for Dyes Effluent Remediation: Broad Visible Light Response. Catalysts. 2023; 13(10):1365. https://doi.org/10.3390/catal13101365

Chicago/Turabian Style

Alsulaim, Ghayah M. 2023. "New BaTi0.96Cu0.02X0.02O3 (X = V, Nb) Photocatalysts for Dyes Effluent Remediation: Broad Visible Light Response" Catalysts 13, no. 10: 1365. https://doi.org/10.3390/catal13101365

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