Efficacious elimination of crystal violet pollutant via photo

Blog

HomeHome / Blog / Efficacious elimination of crystal violet pollutant via photo

Aug 05, 2023

Efficacious elimination of crystal violet pollutant via photo

Scientific Reports volume 13,

Scientific Reports volume 13, Article number: 7723 (2023) Cite this article

290 Accesses

Metrics details

The photo-Fenton process is an appropriate method of the Advanced Oxidation Process that is used in the photocatalysis of organic dyes like crystal violet (CV). La3+ ion substituted gadolinium zirconium oxide Gd(2−x)La(x)Zr2O7 nanopowders (x = 0.1, 0.2, 0.3, and 0.5) have been successfully prepared by using sol–gel auto-combustion method to be used for the efficient photocatalysis of CV with photo-Fenton process. The well-crystallized defect-fluorite, structured with space group: Fm-3m, was detected using X-ray diffraction analysis. The lattice parameters were found to increase with the evaluated La3+ ion concentration. The grain size of the synthesized powders increased with the increase in La3+ ion content. The SAED patterns depicted fluorite structured fluorite. UV/Vis. spectrophotometer was used for the determination of band gap energy of Gd(2−x)La(x)Zr2O7 nanopowders which increased with increasing La3+ ion content. It was found to enhance from 4 to 3.6 eV. The visible spectrophotometer was used for determining unknown concentrations during the photocatalysis process to assure the effectiveness of the process. Overall, results illustrate that the photo-Fenton reaction on Gd(2−x)La(x)Zr2O7 performed excellently in removing crystal violet (CV). The photo-remediation ratio of CV reached 90% within 1 h.

The paper, textile, leather, plastic, electroplating, food processing, pharmaceutical, and agricultural sectors are just a few of the industries that have seen a rise in commerce in the contemporary period1,2. Because these sectors produce high-quality products, they are significant to society as a whole. The majority of industrial operations rely on organic dyes to color their goods, and then they discharge these colors into freshwater aquifers, such as streams and rivers, which eventually empty into the sea3,4,5.

These dyes have been there for a long time, and not only are they harming nearly every type of life seriously, but they are also upsetting the natural equilibrium6,7. About 20% of these colors are dumped into the environment as effluents, harming the ecology. These pollutants are exceedingly difficult to decompose due to their stable structure8,9. Additionally, these pollutants can lead to major issues in people, such as eye irritation, skin allergies, genetic mutations, and liver issues10,11,12.

One of these hazardous dyes, known as crystal violet (CV), is used in a variety of goods made by Gram, including fertilizers, detergents, bacteriostatic agents, colors for leather, and detergents. It is a potent carcinogen for marine creatures. Additionally, CV destroys chromosomes, which results in serious problems when the injured cells divide. Furthermore, CV dye is classified as triphenylmethane cationic dye which is utilized in textile and paper dye sectors. Various products, including fertilizers, antifreeze, detergents, and leather, are also colored using this technique. As a histology stain, CV is also used, most notably in Gram staining to classify the bacteria13,14.

Aquatic life disturbance and water contamination are caused by dye dumping in wastewater15. As a result, a suitable and efficient method for treating wastewater containing dyes, such as CV16, is required due to the dye's proven ability to cause cancer and other mutations in both people and animals17 and humans is essential. Traditional oxidants and coagulants, as well as biodegradation, coagulation, adsorption, and physical deposition, proved insufficient for treating CV18,19.

On the other hand, better oxidants, microwave catalysis, photocatalysis, membrane technology, and advanced oxidation processes (AOPs) seem promising for CV decolorization20,21,22. The fundamental flaw of physical treatment is that it merely moves dyes from a liquid to a solid form, which is difficult to clean. As a result, for degrading such pollutants, chemical treatment employing AOPs, particularly heterogeneous photocatalysis, has garnered interest 23.

According to Wang et al.24, metal oxides in heterogeneous AOPs produce powerful, non-selective hydroxyl radicals (HO•) that break down a range of organic contaminants into short-chain aliphatic acids, which can be more readily eliminated25. In UV–visible light, the electron–hole pair process is necessary to introduce intermediate organic molecules that may completely mineralize at the surface of metal oxide, producing green end products. Since visible light is less expensive than UV light from an economic perspective, it has been extensively studied how new nano-sized photocatalysts respond to it.

In previous works, lots of materials were introduced in the photocatalysis process of dyes contained in wastewater. Nanosphere of TiO2, Mn-doped and PVP-capped ZnO NPs, Ag-modified Ti-doped-Bi2O3, ZnS NPS, CeO2-TiO2 nanocomposite, AgBr-ZnO nanocomposite, and grafted sodium alginate/ZnO/graphene oxide have all been utilized to decolorize CV26. Liu et al.27 used PM-Fe/Ni/H2O2 in the degradation of crystal violet with a Fenton like reaction by Fenton-like reaction. Oladipo et al.28 prepared Octadecylamine capped cadmium nanoparticles CdS to be used in the CV photodegradation process. ZnO/CNA was also prepared for the degradation of CV by Messaoudi et al.29 and lots of materials are used for the degradation of CV as also other organic dyes.

The performance of the aforementioned oxides will next be evaluated, and the results of the present investigation will be contrasted. In this work, the incorporation of lanthanum with Gd2Zr2O7 was discussed to enhance the value of the energy gap value to be appropriate for the photocatalysis process, especially by the Fenton-like method. The reason why we choose pyrochlore is its good ability in the photodegradation of organic dyes due to its chemical composition that distinguishes it from other compounds. This feature qualifies it to be a good photocatalyst. Since many types of pyrochlore were used before in the analysis processes, we used pyrochlore with new elements Gd(2−x)La(x)Zr2O7 that have distinctive features making them compatible with the purpose of photocatalysis. Gd(2−x)La(x)Zr2O7 are also distinguished from others by more characteristics and that assures the novelty of this work.

To understand deeply the fascinating features of pyrochlore, we must know about its classification and composition. Pyrochlore-structured oxides offer a wide range of compositions that lead to outstanding characteristics and applications in superconductivity, thermal barrier coatings, luminescence, and ferromagnetism. Pyrochlore has the general formula A2B2O6O (A = Y or rare earth; B = Ti, Zr, Hf, Sn, Tc, or Pb)30. The pyrochlore structure is a superstructure variant of the simple fluorite structure (AO2 = A4O8, with the A and B cations arranged along the 110 direction). The extra anion vacancy is found in the tetrahedral interstice between two B-site cations. Geometrical frustration and new magnetic effects are particularly vulnerable in these systems. Electronic insulators (e.g., La2Zr2O7), ionic conductors (e.g., Gd1.9Ca0.1Ti2O6.9), mixed ionic and electronic conductors, spin ice systems (Dy2Ti2O7), spin glass systems (Y2Mo2O7), and Haldane chain systems (Tl2Ru2O7) are among the physical properties of the pyrochlore structure (Cd2Re2O7). More disordered structures, such as bismuth pyrochlores, have also been investigated due to their fascinating high-frequency dielectric properties.

Interestingly, herein, rare-earth zirconates with the formula Ln2Zr2O7 (Ln = La–Gd) correspond to the pyrochlore structural type, which in turn belongs to the fluorite homologous series. The pyrochlore structure, as it is known, is a variant of the fluorite structure in which 1/8 of the oxygen is removed and oxygen vacancies are somewhat ordered31,32,33. Consequently, the effect of La3+-substituted Gd2Zr2O7 on the structural parameters, crystallite size, microstructure, and optical properties was considered. Meanwhile, CV was selected as the target elicited organic contaminant model to test the photo-Fenton performance of La3+ -substituted Gd2Zr2O7 nanoparticles.

Materials used in the preparation process include gadolinium nitrate hexahydrate Gd(NO3)3. 6H2O that was supplied from Alfa Aesar Co., as well as zirconyl oxychloride octahydrate ZrOCl2.8H2O from Alpha Chemika Co. and tartaric acid COOH(CHOH)2COOH from Lanxess Co.

The sol–gel auto-combustion technique will be used to create Gd(2−x)La(x)Zr2O7 nanoparticles, where x equals (0.1, 02, 0.3, and 0.5). After mixing, ZrOCl2.8H2O and GdH12N3O15 were dissolved in 40 ml of distilled water in a 1:1 molar ratio. Tartaric acid, which functions as a fuel, will be added to the mixture on a heater set to 350 °C until sol–gel forms. Following that. Sol–gel will be dried in an oven at 100 °C for 24 h. Finally, the dried sample will be fired in a furnace at 1100 °C for two hours to generate Gd(2−x)La(x)Zr2O7 nanoparticles.

The phase composition, crystallinity, crystallite size, and structural parameters of the different samples with Cu Kα radiation were established based on the X-ray diffractometer diffraction patterns (Philips PW 1390). Thermo Scientific's Nicolet iS 10 spectrophotometer is frequently used to measure the Fourier transform infrared (FTIR) absorption spectra of prepared samples at room temperature, which range from 4000 to 400 cm−1. The tested sample was created using the compression procedure, which entailed mixing 200 mg of KBr and 2 mg of powdered Gd(2−x)La(x)Zr2O7 samples to create a transparent disc. UV/Vis/Near IR spectrophotometry (a Jasco V770, Japan) technique is the technology that is used to investigate the optical properties and optical energy gap. The microstructure of Gd(2−x)La(x)Zr2O7 particles was inspected using transmission electron microscopy (Jeol JEM-1011, Japan).

In the visible part of the electromagnetic spectrum, crystal violet has a large absorption band (591 nm). Despite its usage in the textile, culinary, and cosmetic sectors, CV dye can damage aquatic environments.

As shown in Fig. 1., using Gd(2−x)La(x)Zr2O7 nanoparticles as a model, it was chosen as the target organic pollutant to study how it degraded in the presence of UV radiation. The photocatalytic oxidation experiments were carried out in a 100 ml beaker mounted on a magnetic stirrer and exposed to two parallel, 36-Watt UV lamps above the beaker, which was housed in a lamina with silvered inside walls. To ensure uniform photocatalyst dispersion and facilitate the establishment of an adsorption–desorption equilibrium, the catalyst was introduced to a beaker containing 50 mL of an aqueous solution of CV (10 ppm). A solution of 33% H2O2 in 1 mL was added to the reaction container. The reaction was thought to have started with the addition of H2O2. At regular intervals during the procedure, liquid aliquots were taken out of the vessel (15 min). Centrifuging was done on the liquid before analysis. The liquid solutions were analyzed using a visible spectrophotometer after the reactions. Based on the organic dye's rate of degradation, kinetic investigations were carried out. The degradation processes of the dye molecules might be expressed as follows using the Langmuir–Hinshelwood model, which adopts that dye degradation follows pseudo-first-order kinetics:

Graphical representation of photocatalysis process of CV using Gd(2−x)La(x)Zr2O7 nanoparticles.

Integrating this equation yields the following relationship, with the restriction that C = Co at t = 0, where Co is the initial concentration in the bulk solution after dark adsorption and t is the reaction time.

In this equation, Ao stands for the dye's initial absorbance, and T stands for the dye's absorbance at time t. t is the amount of time the dye was exposed to light, and Kapp is the apparent reaction rate constant. The slope of the linear plots on a graph of − ln[At/A0] vs. time should be equal to the apparent-first-order rate constant (Kapp). The Kapp values show how quickly photocatalyst materials break down CV molecules when H2O2 and visible light are present. The CV removal efficiency (η) was determined in the following manner:

Beer-law Lambert states that the dye's absorbance is proportionate to its CV dye concentration.

The submitted work is original and it is not been published elsewhere in any form or language (partially or in full). All authors agreed with the content that all gave explicit consent to submit, and that they obtained consent from the responsible authorities at the institute/organization where the work has been carried out.

X-ray diffraction pattern of Gd(2−x)La(x)Zr2O7 fabricated based on a sol–gel auto-combustion route with tartaric acid as a fuel and annealed at 1100 °C for 2 h is depicted in Fig. 2. The broad diffraction peaks associated with (111), (200), (220), (311), and (222) reflections of the defect-fluorite-structured Gd2Zr2O7 with space group: Fm-3 m (PDF#80–0471) at 2θ = 29.41°, 34.12°, 48.90°, 58.26°, and 61.18° were detected and these results are in a good agreement with other researches34,35. However, the pyrochlore superstructures at 111 (14°), 311 (28°), 331 (37°), and 511 (45°) (space group: Fd-3 m) were not found. Due to the Gd2Zr2O7 structure's proximity to the pyrochlore–fluorite transition's geometric border (RGd/RZr = 1.46), 100% order is not recognized and is attributed to atom inversion. The movement of 10% of oxygen atoms from O1 sites to the initially unoccupied O3 sites is what causes the disordering in the oxygen sublattice.

XRD of Gd(2−x)La(x)Zr2O7 nanoparticles, Gd2Zr2O7 (JCPD01-080–0471(c)).

The prepared rare earth oxide zirconate-based material is mainly. It has a pyrochlore or fluorite crystal structure. Pyrochlores generated from (La1-xGdx)2Zr2O7, If the crystal structure is explained using La2Zr2O7 as an example, La3+ cation is in a specific crystallographic position, 16c (0, 0, 0) (A), Zr4+ The cation occupies 16d sites (1/2, 1/2, 1/2) (B sites), On the (001) plane, it alternately in the [110] direction and the [− 110] direction. An ordered lattice in which the cations of A and B are arranged in a row structure. For the oxygen ion, there are two 48f. positions.

where the wavelength of the Cu target is 1.5406 nm, dRX is the crystallite size, k = 0.9 is a correction factor to account for particle morphologies, FWHM is the full width at half maximum, and is the Bragg angle. As shown in Table 1, the average was discovered to be 27.75 nm at x = 0.1, 38.525 nm at x = 0.2, 40.76 nm at x = 0.3 and 30.85 at x = 0.5. Using Bragg's equation, a = dhkl \(\sqrt{{h}^{2}+{k}^{2}+{I}^{2}}\), the lattice parameter (a) of the synthesized Gd(2−x)La(x)Zr2O7 powders was estimated. d stands for the inter-planar spacing of the primary diffraction peak.

There are similar properties between La and Gd to be substituents for each other. The electronic structure for both of them respectively, [Xe] 5d1 6s2 and [Xe] 4f7 5d1 6s2. Atomic radii are respectively, 195 and 233 pm. Therefore, the results of crystallite size in nm and Lattice parameters for Gd(2−x)La(x)Zr2O7 are very reasonable.

The synthesized sample's FT-IR spectra are shown in Fig. 3. It should be observed that the absorption peak between wave numbers 1640 cm−1 and 3425 cm−1 is connected to the bending and vibrating of OH in water attributable to the water that has been absorbed from the atmosphere34. The absorption bands at 547 and 1425 cm−1 are caused by Gd–O vibrations. The recognizable peak at 715 cm−1 correlates to the Zr–O–Zr stretch-related vibrations. The stretching vibration of the M–O bond (M=Zr–Gd) produces a distinctive peak at 847 and 666 cm−1. Peaks at 1505, 1398, and 1080 cm−1 are associated with Zr–OH vibrations35.

FTIR of Gd(2−x)La(x)Zr2O7 nanoparticles.

The fracture surface of Gd(2−x)La(x)Zr2O7 where x = 0.1 and 0.5, is distinguished using TEM images presented in Fig. 4. The grain growth rate increased with evaluated La3+ ion concentration. The grain displays a cubic-like structure with a degree of agglomeration. The particle size was increased from 20 to 50 nm with a further increase in La3+ ions. Such results can be explained on the basis of the change of ionic radius between La3+ and Gd3+ ions. The SAED diffraction patterns indicated in Fig. 4 were in agreement with a defect-fluorite structured fluorite structure.

TEM micrographs of Gd(2−x)La(x)Zr2O7 nanoparticles.

Figure 5 indicates the Electron Diffraction pattern of Gd(2−x)La(x)Zr2O7 nanoparticles which proves the results of the X-ray and both prove the crystalline features of the prepared samples.

Electron Diffraction pattern of Gd(2−x)La(x)Zr2O7 nanoparticles.

Gd(2−x)La(x)Zr2O7 phosphors' optical band gap is identical to that of Gd(2−x)La(x)Zr2O7 itself based on the Kubelka–Munk theory and diffuse UV–Vis spectral reflectance Fig. 6 illustrates the sample's diffuse reflectance spectrum.

Diffuse reflectance spectrum of Gd(2−x)La(x)Zr2O7 nanoparticles.

From the measurement of UV/Vis. analysis, it is noticed that an increase in the absorbance spectra of the contaminated water (crystal violet solutions) with the increase in purification time. The Kubelka–Munk (K–M) function is used to transform reflectance spectra to comparable absorption spectra as shown in Fig. 7.

Kubelka–Munk plot of pure Gd(2−x)La(x)Zr2O7.

The descriptions of the following equations, which were obtained from the literature for the Eg calculation, indicate the sort of transition that was taken into consideration.

n = 2 for an indirect permitted transition (plotted as h)1/2 vs. E), n = 3 for an indirect prohibited transition (plotted as h)1/3 versus E), n = 3/2 for a direct forbidden transition, and n = 1/2 for a direct allowed transition. (Eg signifies the band gap (eV), h the Planck's constant (J.s), B the absorption constant, v the light frequency (s−1), and (α) the extinction coefficient, which is proportional to F. (R). The best linear fit in the absorption spectra can be used to experimentally calculate the n value for a particular transition using several formulas.

Below the conduction band, which reached up to the valence band maximum, the Gd d states were seen. La2Zr2O7's La d states are connected to Gd d states that extend in the direction of the band gap. In the Gd16Zr16O56 system, the valence band is mostly contributed by O p states, whereas the conduction band is produced by Zr d states. The Gd d states have a partial coupling to the valence and conduction bands and a partial band gap appearance. Thermal conductivity variations of La2Zr2O7 and Gd2Zr2O7 with varying La/Gd concentrations36.

Figure 8 illustrates the final calibration curve used to calculate the concentration of crystal violet solutions during the purification process using the information on the solutions' wavelength-dependent absorbance (591 nm).

Calibration curve of crystal violet solutions.

The concluded equation is:

According to the examination of the visible spectrophotometer measurements, the absorbance spectra of the contaminated water (crystal violet solutions) increase as the purification time increases. From the Crystal Violet solution, known concentrations of 1, 2, 3,…, and 10 ppm were generated. By using a visible spectrophotometer to measure their absorbance spectrum, a calibration curve for those concentrations was created, and an equation was then developed to use their absorbance to compute the unknown amounts that would follow from the purification procedure, as shown in Fig. 9.

Absorbance of 10 ppm solution of crystal violet after treatment with (a)H2O2 only, 0.05 gm of Gd(2−x)La(x)Zr2O7 without H2O2 where (c) x = 0.1, (g) x = 0.2, (k) x = 0.3, (o) = 0.5, Also with H2O2 where (e) = 0.1, (i) = 0.2, (m) = 0.3, (q) = 0.5 and (b–r) are concentrations conversion C/C0 versus time for degradation process where (s) is the figure of combined data.

It was put through a photocatalyst test using crystal violet solution (10 ppm) with intervals of 15 min between 0 and 90 min. Gd(2−x)La(x)Zr2O7 nanoparticles were utilized as a catalyst for the Fenton-like reaction to purify by being exposed to hydrogen peroxide solution (32 watts). Different researches have been carried out with lots of materials for decolorization of crystal violet by photodegradation or even Fenton-like reaction as shown in Table 3 and Gd(2−x)La(x)Zr2O7 nanoparticles proved a great efficiency compared with these materials.

It is worth noting that Gd(2−x)La(x)Zr2O7 has never been used with the same composition in the photocatalysis process, according to our consideration. According to the literature survey, there are two types of researches or more that prepared the same elements used in our research, but with a different composition, the basis of which is La2Zr2O7, with substitutions of gadolinium, unlike our research in which G2Zr2O7 is the base. In addition, this composition was not used in the process of water purification, and this indicates that our research is one of the first research that used Gd(2−x)La(x)Zr2O7 with the same method in the photocatalysis process37,38.

The purification procedure was performed under identical conditions using gadolinium zirconium oxide alone, hydrogen peroxide alone again, and then both of them combined with 1 ml of hydrogen peroxide and 0.025 g of pyrochlore in 50 ml of crystal violet solution. It was observed that pyrochlore alone demonstrated a respectable efficiency and that this efficiency was increased by combining it with hydrogen peroxide. They showed rapid degradation of crystal violet, going from a concentration of 10 ppm to 0 ppm in no more than 90 min. By comparing with other researches, most of them used the same percent of photocatalyst in wastewater treatment process39.

Table 2 shows the values of change in concentration or relative concentration (C/C0) for crystal violet solutions (50 ml) during the purification process. The photocatalysis process has been carried out in three parallel categories. The first of them included the photocatalysis of crystal violet solution (50 ml) by using 1 ml of hydrogen peroxide alone (H2O2). The second has proceeded with Gd(2−x)La(x)Zr2O7 pyrochlore in addition to hydrogen peroxide. The last one was carried out by using Gd(2−x)La(x)Zr2O7 pyrochlore alone (Gd(2−x)La(x)Zr2O7 nanopowder). Moreover, all of these categories have proceeded under the influence of ultraviolet radiation in which the sample was taken for analysis every 15 min.

Photocatalysis efficiency is a vital parameter that should be taken into consideration. In the case of photocatalysis with Gd(2−x)La(x)Zr2O7 nanoparticles alone, it proved an average efficiency according to values of relative concentration (C/C0) in Table 2.

For the catalysis with pyrochlore in addition to hydrogen peroxide, the interpretation is slightly different, as it is done with a process called Fenton-Like Reaction, and in this case, hydrogen peroxide is indispensable. Similar to the situation of purification with hydrogen peroxide alone, this case is characterized by its high efficacy and a rate of removal that approaches zero in a period time that is likewise close to 90 min. Although it may appear superficially that Gd(2−x)La(x)Zr2O7 pyrochlore has no function as long as we get the same percentage with or without it, the reality is rather different. The efficiency of the purification process in the case of pyrochlore with hydrogen peroxide is better than hydrogen peroxide alone, even if we get the same efficiency approximately at the end of the 90 min purification period according to the values of the relative concentrations5. Thus, It was found that the efficiency of the Fenton-Like approach, which comprises pyrochlore and hydrogen peroxide, is greater than pyrochlore alone at the start of the process. Furthermore, we employed a very low concentration of pyrochlore, up to 0.025 g per 50 ml of pollutant. If this percentage is increased, the purification process will be accelerated. and thus proves the efficiency of pyrochlore in the degradation process by employing a Fenton-like reaction (Table 3).

In this context, the photocatalytic method is characterized by the separation of electron/hole pairs (e−/h+) and the separation of photogenerated carriers can be connected to the crystalline phase, the aspect ratio of nanopowders, electronic structure of the facets, and defects. The reaction Eqs. (6)–(19) can be used to represent the mechanism of photocatalytic degradation of crystal violet dye40.

The photogenerated VB (h+) nanoparticles combine with either H2O or OH- to form OH· via the following reaction with assessed activity and indirectly decay the dye molecule41.

As a result of the presence of hydroxyl groups, water, and oxygen at the surface of (Gd(2−x)La(x)Zr2O7 nanoparticles, the electrons (e−) in the conduction band react with species absorbed on the (Gd(2−x)La(x)Zr2O7 surface, free radicals to generate O2·−, and following the subsequent steps, it results in the generation of OH· radicals.

Finally, the produced agents degraded the dye pollution.

It is known that the use of H2O2 in combination with the photocatalyst was shown to be effective in eliminating hazardous organic species from water. H2O2 can hinder the surface recombination of electron–hole pairs and stimulate the production of hydroxyl radicals5.

Due to its reaction with the valence band holes and conduction band electrons in photocatalytic systems, H2O2 raises the concentration of the ·OH and ·O2 radicals. H2O2 can react with ·OH, HO2·, or ·O2 radicals to produce further radicals (Eqs. 20–25). The oxidative species interact with one another or with the holes they create to produce O2 (Eqs. 26, 27). Additionally, UV light exposure to H2O2 results in the formation of hydroxyl radicals (Eq. 28).

The great differences are mainly derived from photo-induced thermal energy, which in-situ heating active sites to lower the H2O2 activation barrier44.

In summary, Gd(2−x)La(x)Zr2O7 with a variable Gd/La ratio was successfully prepared using a sol–gel auto-combustion approach based on tartaric acid as fuel. The La-substituted Gd2Zr2O7 nanopowders existed in a defective fluorite phase. The crystallite size of the synthesized nanoparticles was found to be from 28.5 to 39.2 nm. Unit cell volume was found to be 1.2 × 10–21 cm3. The grain size of Gd(2−x)La(x)Zr2O7 increased with increasing La3+ ion content. Most of the synthesized particles demonstrate a cubic-like structure. The diffuse reflectance spectrum confirmed the synthesized particles are highly transmissive. The band gap energy of the synthesized Gd(2−x)La(x)Zr2O7 samples was found to be (4–3.6 eV). Finally, the synergistic effect of Gd(2−x)La(x)Zr2O7 and H2O2 possessed an advanced oxidation reaction with high efficiency towards organic pollutants degradation, such as crystal violet dye, in a cost-effective way and a short time.

All data generated or analyzed during this study are included in this article.

Ali, N. et al. Selenide-chitosan as high-performance nanophotocatalyst for accelerated degradation of pollutants. Chem. Asian J. 15, 2660–2673 (2020).

Article CAS PubMed Google Scholar

Ahmad, I., Khan, S. B., Kamal, T. & Asiri, A. M. Visible light activated degradation of organic pollutants using zinc–iron selenide. J. Mol. Liq. 229, 429–435 (2017).

Article CAS Google Scholar

Ali, N. et al. Enhanced photodegradation of methylene blue with alkaline and transition-metal ferrite nanophotocatalysts under direct sun light irradiation. J. Chin. Chem. Soc. 66, 402–408 (2019).

Article CAS Google Scholar

Ghalab, S., Sebak, M. A., Abdallah, E. M., & Abdelghany, A. M. Chitosan/waste glass composite as new material for copper removal from contaminated water. Mater. Chem. Phys. 290, 126613 (2022).

Article CAS Google Scholar

Chong, M. N., Jin, B., Chow, C. W. & Saint, C. Recent developments in photocatalytic water treatment technology: A review. Water Res. 44, 2997–3027 (2010).

Article CAS PubMed Google Scholar

Ribeiro, A. R., Nunes, O. C., Pereira, M. F. & Silva, A. M. An overview on the advanced oxidation processes applied for the treatment of water pollutants defined in the recently launched Directive 2013/39/EU. Environ. Int. 75, 33–51 (2015).

Article CAS PubMed Google Scholar

Abdelghany, A., Oraby, A. & Abdelbaky, M. The effect of borate bioactive glass ceramics containing silver nanoparticles on removal physiognomies of methylene blue. Optik 267, 169694 (2022).

Article ADS CAS Google Scholar

Tabelini, C. H. B., Lima, J. P. P. & Aguiar, A. Gallic acid influence on azo dyes oxidation by Fenton processes. Environ. Technol. 43(22), 3390–3400 (2021).

Article PubMed Google Scholar

Ali, N. et al. Thermochemical and electrochemical aspects of carbon dioxide methanation: A sustainable approach to generate fuel via waste to energy theme. Sci. Total Environ. 712, 136482 (2020).

Article ADS CAS PubMed Google Scholar

Nezamzadeh-Ejhieh, A. & Banan, Z. Sunlight assisted photodecolorization of crystal violet catalyzed by CdS nanoparticles embedded on zeolite A. Desalination 284, 157–166 (2012).

Article CAS Google Scholar

Nezamzadeh-Ejhieh, A. & Karimi-Shamsabadi, M. Comparison of photocatalytic efficiency of supported CuO onto micro and nano particles of zeolite X in photodecolorization of Methylene blue and Methyl orange aqueous mixture. Appl. Catal. A 477, 83–92 (2014).

Article CAS Google Scholar

Mittal, A., Mittal, J., Malviya, A., Kaur, D. & Gupta, V. Adsorption of hazardous dye crystal violet from wastewater by waste materials. J. Colloid Interface Sci. 343, 463–473 (2010).

Article ADS CAS PubMed Google Scholar

Zahoor, M. Removal of crystal violet from water by adsorbent prepared from Turkish coffee residue. Tenside, Surfactants, Deterg. 49, 107–113 (2012).

Article CAS Google Scholar

Thairu, Y., Nasir, I. A. & Usman, Y. Laboratory perspective of gram staining and its significance in investigations of infectious diseases. Sub-Sahar. Afr. J. Med. 1, 168 (2014).

Article Google Scholar

Edokpayi, J. N., Odiyo, J. O. & Durowoju, O. S. Impact of wastewater on surface water quality in developing countries: A case study of South Africa. Water Qual. 10, 66561 (2017).

Google Scholar

Mashkoor, F., Nasar, A. & Asiri, A. M. Exploring the reusability of synthetically contaminated wastewater containing crystal violet dye using tectona grandis sawdust as a very low-cost adsorbent. Sci. Rep. 8, 1–16 (2018).

Article CAS Google Scholar

Mani, S. & Bharagava, R. N. Exposure to crystal violet, its toxic, genotoxic and carcinogenic effects on environment and its degradation and detoxification for environmental safety. Rev. Environ. Contam. Toxicol. 237, 71–104 (2016).

CAS PubMed Google Scholar

Li, L., Wei, D., Wei, G. & Du, Y. Transformation of cefazolin during chlorination process: Products, mechanism and genotoxicity assessment. J. Hazard. Mater. 262, 48–54 (2013).

Article CAS PubMed Google Scholar

Yang, Y. et al. Magnetic ion exchange resin for effective removal of perfluorooctanoate from water: Study of a response surface methodology and adsorption performances. Environ. Sci. Pollut. Res. 25, 29267–29278 (2018).

Article CAS Google Scholar

Wang, M., Tan, G., Ren, H., Xia, A. & Liu, Y. Direct double Z-scheme Og-C3N4/Zn2SnO4N/ZnO ternary heterojunction photocatalyst with enhanced visible photocatalytic activity. Appl. Surf. Sci. 492, 690–702 (2019).

Article ADS CAS Google Scholar

Zhang, X. et al. g-C3N4/UiO-66 nanohybrids with enhanced photocatalytic activities for the oxidation of dye under visible light irradiation. Mater. Res. Bull. 99, 349–358 (2018).

Article CAS Google Scholar

Costa, E. P. et al. New trend on open solar photoreactors to treat micropollutants by photo-Fenton at circumneutral pH: Increasing optical pathway. Chem. Eng. J. 385, 123982 (2020).

Article CAS Google Scholar

Chestnut, L. G. & Mills, D. M. A fresh look at the benefits and costs of the US acid rain program. J. Environ. Manage. 77, 252–266 (2005).

Article CAS PubMed Google Scholar

Wang, B. et al. Application of heterogeneous catalytic ozonation for refractory organics in wastewater. Catalysts 9, 241 (2019).

Article Google Scholar

Matavos-Aramyan, S. & Moussavi, M. Advances in Fenton and Fenton based oxidation processes for industrial effluent contaminants control-a review. Int. J. Environ. Sci. Nat. Resour 2, 1–18 (2017).

Google Scholar

Abdel-Khalek, A. A., Mahmoud, S. & Zaki, A. Visible light assisted photocatalytic degradation of crystal violet, bromophenol blue and eosin Y dyes using AgBr-ZnO nanocomposite. Environ. Nanotechnol. Monit. Manag. 9, 164–173 (2018).

Google Scholar

Liu, J., Du, Y., Sun, W., Chang, Q. & Peng, C. Preparation of new adsorbent-supported Fe/Ni particles for the removal of crystal violet and methylene blue by a heterogeneous Fenton-like reaction. RSC Adv. 9, 22513–22522 (2019).

Article ADS CAS PubMed PubMed Central Google Scholar

Oladipo, S. D. & Omondi, B. Photodegradation of crystal violet dye in water using octadecylamine-capped CdS nanoparticles synthesized from Cd (II) N, N′-diarylformamidine dithiocarbamates and their 2, 2-bipyridine adducts. J. Coord. Chem. 75, 2170–2188 (2022).

Article CAS Google Scholar

Messaoudi, Z. A. et al. Adsorption and photocatalytic degradation of crystal violet dye under sunlight irradiation using natural and modified clays by zinc oxide. Chem. Methodol. 6, 661–676 (2022).

CAS Google Scholar

Subramanian, M. G. Aravamudan and GV Subba Rao. Prog. Solid State Chem. 15, 124 (1983).

Google Scholar

Li, Y. et al. Defect formation energies in A2B2O7 pyrochlores. Scripta Mater. 107, 18–21 (2015).

Article CAS Google Scholar

Stanek, C. et al. Predicted structure and stability of A4B3O12 δ-phase compositions. Phys. Rev. B 80, 174101 (2009).

Article ADS Google Scholar

Popov, V. et al. Short-and long-range order balance in nanocrystalline Gd2Zr2O7 powders with a fluorite-pyrochlore structure. Russ. J. Inorg. Chem. 59, 279–285 (2014).

Article CAS Google Scholar

Luo, F. et al. Insight into the effect of Nd2O3 and CeO2 co-doped Gd2Zr2O7 ceramics without structural design: Phase evolution and chemical durability. Vacuum 203, 111256 (2022).

Article ADS CAS Google Scholar

Bahamirian, M., Hadavi, S., Farvizi, M., Rahimipour, M. & Keyvani, A. Enhancement of hot corrosion resistance of thermal barrier coatings by using nanostructured Gd2Zr2O7 coating. Surf. Coat. Technol. 360, 1–12 (2019).

Article CAS Google Scholar

Khan, H., Iqbal, Y., Khan, M. & Zeng, Y. Variations in the thermal conductivity of La2Zr2O7 and Gd2Zr2O7 with variable La/Gd concentrations. Phys. B 614, 413018 (2021).

Article CAS Google Scholar

Li, L. et al. Preparation of epitaxial La2−xGdxZr2O7 buffer layer for coated conductors using sol-gel method. Mater. Chem. Phys. 250, 123147 (2020).

Article ADS CAS Google Scholar

Li, L. et al. Highly epitaxial YBa2Cu3O7− δ films grown on gradient La2−xGdxZr2O7-buffered NiW-RABiTS using all sol–gel process. Supercond. Sci. Technol. 34, 045004 (2021).

Article ADS Google Scholar

Kaviyarasu, K. et al. High performance of pyrochlore like Sm2Ti2O7 heterojunction photocatalyst for efficient degradation of rhodamine-B dye with waste water under visible light irradiation. J. King Saud Univ. Sci. 32, 1516–1522 (2020).

Article Google Scholar

Mahmoud, S. A. & Fouad, O. A. Synthesis and application of zinc/tin oxide nanostructures in photocatalysis and dye sensitized solar cells. Sol. Energy Mater. Sol. Cells 136, 38–43 (2015).

Article CAS Google Scholar

El-Shazly, A. et al. Nanostructured ZnO photocatalysts prepared via surfactant assisted co-precipitation method achieving enhanced photocatalytic activity for the degradation of methylene blue dyes. J. Environ. Chem. Eng. 4, 3177–3184 (2016).

Article CAS Google Scholar

Saeed, M. et al. Synthesis of pn NiO-ZnO heterojunction for photodegradation of crystal violet dye. Alex. Eng. J. 65, 561–574 (2023).

Article Google Scholar

Alwan, D. A. & Hatem, O. A. Preparation and characterization of ternary composite rGO/Fe3O4/CdS and evaluating its efficiency in photodegradation of crystal violet dye. J. Chin. Chem. Soc. 69, 1845–1854 (2022).

Article CAS Google Scholar

Yang, L. et al. Photo-thermal synergy for boosting photo-Fenton activity with rGO-ZnFe2O4: Novel photo-activation process and mechanism toward environment remediation. Appl. Catal. B 292, 120198 (2021).

Article CAS Google Scholar

Download references

Open access funding provided by The Science, Technology & Innovation Funding Authority (STDF) in cooperation with The Egyptian Knowledge Bank (EKB).

Faculty of Science, New Mansoura University, International Coastal Rd, New Mansoura City, Dakahlia, Egypt

M. Abdelbaky

Spectroscopy Department, Physics Research Institute, National Research Centre, 33 ElBehouth St., DokkiGiza, 12311, Egypt

A. M. Abdelghany

Physics Department, Faculty of Science, Mansoura University, Mansoura, 35516, Egypt

A. H. Oraby & E. M. Abdelrazek

Electronic and Magnetic Materials Department, Advanced Materials Institute, Central Metallurgical Research and Development Institute (CMRDI), P.O. Box 87, Helwan, 11421, Cairo, Egypt

M. M. Rashad

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

M.A. designed the research and conducted the experiments. A.M.A. revised the research and the design. A.H.O. and E.M.A. revised the research. M.R. discussed the results from the beginning and revised the research in its final form. All authors agreed to publish this manuscript.

Correspondence to M. Abdelbaky.

The authors declare no competing interests.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and Permissions

Abdelbaky, M., Abdelghany, A.M., Oraby, A.H. et al. Efficacious elimination of crystal violet pollutant via photo-Fenton process based on Gd(2−x)La(x)Zr2O7 nanoparticles. Sci Rep 13, 7723 (2023). https://doi.org/10.1038/s41598-023-34838-w

Download citation

Received: 05 February 2023

Accepted: 09 May 2023

Published: 12 May 2023

DOI: https://doi.org/10.1038/s41598-023-34838-w

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.