Next Article in Journal
Exploring Preliminary Biocompatibility Testing in Coating Development
Previous Article in Journal
The Use of Polyurethane Composites with Sensing Polymers as New Coating Materials for Surface Acoustic Wave-Based Chemical Sensors—Part III: Ultrasonic Analyses and Optical Microscopy Characterization of the Coating Results
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Eco-Friendly Synthesis of Al2O3 Nanoparticles: Comprehensive Characterization Properties, Mechanics, and Photocatalytic Dye Adsorption Study

by
Ahlam Hacine Gharbi
1,2,
Salah Eddine Laouini
1,2,*,
Hadia Hemmami
1,2,3,
Abderrhmane Bouafia
1,2,
Mohammed Taher Gherbi
4,
Ilham Ben Amor
1,3,
Gamil Gamal Hasan
1,
Mahmood M. S. Abdullah
5,
Tomasz Trzepieciński
6 and
Johar Amin Ahmed Abdullah
7
1
Department of Process Engineering and Petrochemical, Faculty of Technology, University of El Oued, P.O. Box 789, El Oued 39000, Algeria
2
Laboratory of Biotechnology Biomaterial and Condensed Matter, Faculty of Technology, University of El Oued, El Oued 39000, Algeria
3
Renewable Energy Development Unit in Arid Zones (UDERZA), University of El Oued, P.O. Box 789, El Oued 39000, Algeria
4
Department of Mechanical Engineering, University of El Oued, El Oued 39000, Algeria
5
Department of Chemistry, College of Science, King Saud University, P.O. Box 2455, Riyadh 11451, Saudi Arabia
6
Department of Manufacturing Processes and Production Engineering, Rzeszow University of Technology, Al. Powstańców Warszawy 8, 35-959 Rzeszów, Poland
7
Department of Chemical Engineering, Higher Technical School, University of Seville, 41011 Seville, Spain
*
Author to whom correspondence should be addressed.
Coatings 2024, 14(7), 848; https://doi.org/10.3390/coatings14070848
Submission received: 14 May 2024 / Revised: 20 June 2024 / Accepted: 4 July 2024 / Published: 6 July 2024

Abstract

:
Metal and metal oxide nanoparticles are gaining traction in inorganic catalysis and photocatalysis, driving the development of eco-friendly methods. This study introduces an eco-friendly and cost-effective approach for synthesizing Al2O3 nanoparticles (NPs) using extracts derived from the leaves of Calligonum comosum L. The primary objective of this investigation is to assess the photocatalytic efficacy of the synthesized catalyst in addressing organic pollutants. The Al2O3 NPs exhibit a spherical morphology with crystalline arrangements, as evidenced by an average crystallite size of 25.1 nm in the XRD analysis. The band gap energy of the Al2O3 NPs is determined to be 2.86 eV. In terms of mechanical properties, the Al2O3 NPs show significant potential in enhancing both flexural and compressive properties, thereby making them a viable choice for improving the mechanical performance of composites. Notably, the Young’s modulus of the hybrid composite (comprising plant material and Al2O3 NPs) exhibits a remarkable increase of 34.4% in flexion and 78.3% in compression compared to the plant material alone. The catalytic performance of the Al2O3 NPs is evaluated using methylene blue (MB) as a cationic dye and Rose Bengal (RB) as an anionic dye. Impressively, the Al2O3 NPs demonstrate degradation efficiencies of 98.2% for MB and 90.5% for RB. The degradation processes occur under solar light irradiation, with a contact time of 120 m, a maintained pH of 7, and a temperature of 25 °C. This study found that Al2O3 nanoparticles are a promising, cost-effective, and environmentally friendly option for water treatment.

1. Introduction

Water, often referred to as the lifeblood of all living organisms, comprises a substantial fraction of their body mass [1]. In the contemporary era, however, industrial activities have significantly compromised this vital resource, particularly in the context of potable water and aquatic ecosystems. The extensive use of a variety of chemicals, notably dye contaminants from industries such as leather, textiles, pharmaceuticals, and cosmetics, has transformed water bodies into reservoirs of pollution. Effluents from textile industries, which are replete with a combination of organic, inorganic, polymeric, and elemental pollutants, pose a severe threat to water quality, aquatic biodiversity, soil fertility, and ecosystems at large [2,3,4,5]. In the face of these daunting challenges, nanotechnology has emerged as a promising solution, offering advanced catalytic processes to effectively mitigate water pollution. Although several wastewater treatment methods are currently in use, including flocculation, oxidation, and reverse osmosis, many of these techniques are hindered by lengthy procedures, high operational costs, and limited efficiency [6]. In stark contrast, photodegradation, especially when facilitated by photocatalysis, offers a rapid, cost-effective, and efficient alternative for water treatment [7]. Al2O3 NPs have demonstrated robust adsorption activity for heavy metals, thereby showing promise for water purification applications [8]. The escalating global challenge of providing clean and fresh water, particularly in the context of a growing population, is a pressing issue, with over 0.78 billion people lacking access to safe water resources [9]. Elevated concentrations of heavy metals such as lead and cadmium in water can lead to severe health issues [10,11]. Al2O3 NPs, with their impressive mechanical strength, including flexural, compressive, and tensile strengths, emerge as robust and versatile materials [12]; their large specific surface area and high adsorption capacity make them particularly advantageous for adsorption processes, especially at low temperatures [13]. Recent research indicates that plant extracts could potentially serve as a viable, eco-friendly option for the synthesis of nanomaterials. The active constituents present in plant extracts, including enzymes, polyphenols, phenolics, carbohydrates, and proteins, have demonstrated the ability to efficiently reduce, stabilize, and encapsulate metal ions, leading to the creation of metal nanoparticles [14]. Notably, nanoparticles produced from plant sources exhibit greater stability and uniformity in both shape and size compared to those produced by conventional chemical approaches [15]. Several studies have investigated the effect of Al2O3 NPs on the adsorption of cationic dyes, further highlighting their potential in addressing water pollution, for instance, Ali et al.’s work [16]. The synthesis of γ-alumina (Al2O3) nanoparticles involves the use of formamide and Tween-80. Al2O3 NPs proved effective in adsorbing methylene blue dye from water, with capacity increasing from 490 to 2210 mg g−1 as initial concentration rose from 50 to 400 mg L−1 under pH 9, 10 min reaction time, and 60 °C conditions. Manikandan et al. [17] synthesized green-treated (Al2O3 NPs) using a bio-reduction method, assessing pH’s impact on particle size. A Prunus × yedoensis leaf extract (PYLE) was used for green synthesis, exhibiting Al2O3 NPs’ nitrate removal (94%). The authors suggest optimizing pH and nanoparticle size for enhanced catalytic activity.
This research tackles an urgent environmental issue by synthesizing non-toxic Al2O3 NPs using extracts from the readily accessible and non-hazardous plant species Calligonum comosum L. The detailed analysis of these nanoparticles emphasizes their importance in fulfilling the needs of industrial applications, which demand a fine equilibrium between flexural strength, stiffness, high compressive strength, and resistance to deformation. The outcomes not only demonstrate the adaptability of this approach but also underscore its effectiveness in removing organic pollutants from water-based solutions. Additionally, the evaluation of their photocatalytic efficiency against standard organic dyes under sunlight is intended to provide useful information. The expected results suggest that the nanocomposite could serve as a sustainable and efficient solution in light industries and for tackling water pollution.

2. Materials and Methods

2.1. Materials

Leaves of Calligonum comosum L. and Arachis hypogaea L. were collected from El Oued, Algeria (coordinates: 33°22′06″ N, 6°52′03″ E). Aluminum nitrate nonahydrate (Al(NO3)3·9H2O, 99%), sodium hydroxide (NaOH, 97%), Methyl blue (MB, C16H18ClN3S) and Rose Bengal (RB, C20H44I4N4O13S) dyes, both with purity greater than 99%. All chemicals were procured from Biochem Chemopharama, Karlsruhe, Germany. Deionized water, which is free of minerals, was used as the solvent for all experiments.

2.2. Preparation of Plant Extract

First, fresh Calligonum comosum L. leaves were meticulously washed with distilled water to remove any contaminants. These leaves were then left to air-dry in a shaded area at room temperature, ensuring that all moisture evaporated naturally.
To prepare the extract, 10 g of the dried leaves was combined with 150 mL of deionized water. This mixture was heated to 65 °C and allowed to simmer for two hours. As the process progressed, the solution turned a light brown hue, indicating the extraction of essential compounds.
The next step involved filtering this brown solution through Whatman N°. 1 filter paper to remove any solid particles. The filtered extract was then carefully stored in a refrigerator, preserving its properties for future experiments [18].

2.3. Synthesis of Al2O3 NPs

To synthesize Al2O3 nanoparticles using a Calligonum comosum L. leaf extract, a specific protocol was followed. First, 50 mL of the leaf extract was combined with 100 mL of a 0.1 M Al(NO3)3·9H2O solution. This mixture was continuously stirred at 70 °C for 2 h. During the stirring process, a 2 M NaOH solution was gradually added drop by drop. The appearance of a light yellow color in the solution indicated the formation of Al2O3 nanoparticles.
The resulting precipitate was separated from the reaction mixture through centrifugation at 3000 rpm for 10 min, which helped remove impurities. After centrifugation, the precipitate was dried at 80 °C for 24 h. Finally, it was calcined at 900 °C for 6 h. The end product of this synthesis was white Al2O3 nanoparticles [19].

2.4. The Prepared Al2O3 NPs for Mechanical Properties

In the synthesis of the prepared Al2O3 NPs for mechanical properties, we combine 10 g (70%) of the plant leaves of Calligonum comosum L. with 4.3 g (30%) of Al2O3 NPs, followed by the addition of a specific quantity of resin. The mixture is thoroughly blended; placed in a mold having dimensions of 30 cm by 30 cm and a thickness of 4 mm, using a vacuum method; and left to dry for 48 h. This process results in the formation of a composite material, Al2O3 NPs, suitable for flexural and compression tests (Figure 1).

2.5. Characterization

To determine the morphology of the Al2O3 nanoparticles (NPs), we employed a scanning electron microscope (SEM, Leo Supra 55, Zeiss Inc., Oberkochen, Germany), operating at 10 kV. This technique allowed us to visualize the shape and surface structure of the nanoparticles. For crystal structure verification, we performed an X-ray diffraction (XRD) analysis using a Rigaku Miniflex 600 diffractometer (Rigaku, Tokyo, Japan) with Cu-Kα radiation (λ = 1.5406 Å). The analysis covered 2θ angles ranging from 10° to 80°, providing detailed insights into the structural characteristics and grain size of the nanoparticles. To identify the functional groups present in the prepared compounds, Fourier transform infrared (FTIR) spectroscopy was conducted. We utilized a Perkin-Elmer Corporation device (Series 1725X, Norwalk, CT, USA) for this purpose. The band gap energy (Eg) of the Al2O3 NPs was determined using the Tauc relationship: (hv) = A (h − Eg)n. Absorption spectra were recorded with a Shimadzu UV-Vis apparatus (UV-2450, Shimadzu, Duisburg, Germany). For this measurement, samples were dispersed in distilled water, maintaining a concentration of 0.1 mg of nanoparticles in 2 mL of water.

2.6. Mechanical Properties of the Prepared Al2O3 NPs

Following the fabrication of the Al2O3 NP material, mechanical tests, specifically flexural and compression tests, were conducted on both the plant material alone and the combination of plant material with Al2O3 NPs. Flexural examinations utilized the Zwick Z 2.5 machine, featuring a 2 kN sensor. Conversely, compression tests were carried out using the Zwick/Roell Z10 machine (Ulm, Germany), controlled by Test Xpert software version 12.0 and equipped with a precise 10 kN force sensor. To enable a comprehensive comparison of the mechanical performance of each sample, these assessments were executed under consistent conditions, including ambient temperature and a testing speed of 5 mm/min.

2.7. Photocatalytic Degradation of RB and MB

The use of Al2O3 nanoparticles (NPs) in photocatalysis has proven to be highly effective in degrading hazardous AZO dyes, like methylene blue (MB) and Rose Bengal (RB), present in wastewater. This method leverages sunlight to activate the nanoparticles, leading to the formation of reactive species that can decompose the dyes. As a result, this process not only breaks down the dyes but also fully mineralizes them, presenting an eco-friendly and efficient solution for wastewater treatment [20].
To evaluate the catalytic activity of the samples, the photodegradation of MB and RB dyes in aqueous solutions under sunlight was observed. The dye solutions were prepared at a concentration of 2.5 × 10−5 M, with 5.0 mg of a dye catalyst per 5.0 mg of Al2O3 nanoparticles. After thoroughly mixing the samples, they were exposed to direct sunlight. The reaction’s progress was tracked using a UV-Vis spectrometer at specified intervals of 5, 15, 30, 45, 60, 75, 90, 105, and 120 min. To stop the degradation process, the solution was centrifuged at 3000 rpm for 5 min. The absorbance of MB and RB dyes was measured at wavelengths of 663 nm and 542 nm, respectively, using a UV-Vis spectrophotometer.
To evaluate the extent of adsorption, the dye concentrations (MB and RB) were monitored by comparing the initial aqueous solution with the solution post-photodegradation test [21].
q e = C 0 C e v m
The quantity of dye absorbed by the adsorbent at equilibrium is symbolized as qe (mg·g−1). The dye solution’s initial concentration is noted as C0 (mg·L−1), and its concentration at equilibrium is indicated as Ct (mg·L−1). In this context, “m” refers to the mass of Al2O3 nanoparticles utilized (g), while “V” stands for the volume of the dye solution (L). The reaction’s progress is tracked through UV-Vis spectroscopy. When Al2O3 NPs are added to the reaction mixture, they catalyze the reduction in the dye. The efficiency of the photodegradation of the dye (MB and RB) is determined using Equation (2) [22].
D e g r a d a t i o n   r a t i o   ( % ) = ( C 0 C t ) C 0 × 100

3. Result and Discussion

3.1. UV–Vis Spectroscopy

Figure 2a shows UV-Vis spectra of the Calligonum comosum L. extract, revealing peaks at 211 nm (condensed tannins) [23], 232 nm (flavonoids) [24], and 276 nm (potential polyphenols) known for antioxidant properties [25,26]. In green synthesis, a solution with metal salts and a plant extract rich in tannins, flavonoids, or polyphenols is used. These compounds reduce Al3+ ions, leading to nanoparticle formation. The stabilizing properties of the compounds prevent nanoparticle agglomeration [27].
Figure 2b depicts the UV-Vis spectra of Al2O3 NPs, revealing a peak at 359 nm, and the absorption spectrum of Al2O3 NPs. These findings were compared to prior reports indicating an absorption peak for the biogenesis of Al2O3 NPs at 322 nm [28], 345 nm [29], and 382 nm [30]. Figure 2c illustrates the band gap energies of Al2O3 NPs, determined to be 2.86 eV. The distinct UV bands exhibited by Al2O3 NPs in the UV range indicate a narrow nanoparticle distribution and suggest potential applications in photocatalysis [31]. Previous studies reported varying optical band gaps for Al2O3 NPs produced in different solvent systems [17].

3.2. FTIR Spectroscopy Analysis

To determine the functional groups present in the leaf extract and Al2O3 nanoparticles (NPs), Fourier transform infrared (FTIR) spectroscopy was employed. Various peaks were observed in the FTIR spectrum of Al2O3 NPs at wavenumbers 3430, 2341, 1676, 1488, 1070, 828, and 586 cm−1 (Figure 3). The broad peak at 3430 cm−1 is indicative of the O–H vibration, typically from water molecules [32]. At 2341 cm−1, a weak band appears, which corresponds to the C-H bond stretching in methyl groups [32].
In the 1676 cm−1 region, bands signifying the presence of the O=C bond were detected [33]. The peak at 1488 cm−1 is associated with the Al-OH bonding mode [29]. The peaks at 1070 cm−1 and 828 cm−1 are attributed to the symmetric bending of Al–O–H and the Al–O stretching mode in the tetrahedral structure, respectively [34]. Finally, a strong absorption band at 586 cm−1 confirms the presence of the Al-O bond stretching vibration, thereby validating the formation of Al2O3 NPs [35].

3.3. X-ray Diffraction

Figure 4 illustrates the XRD analysis of Al2O3 nanoparticles, revealing the presence of gamma (γ) and delta (δ) phases. The diffraction pattern matches the standard JCPDS patterns for γ-Al2O3 (JCPDS 10-0425) and δ-Al2O3 (JCPDS 00-016-0394). The crystal characteristics are as follows: γ-Al2O3 adopts a cubic structure with dimensions a = 7.9000 Å, b = 7.9000 Å, c = 7.9000 Å, α = 90°, β = 90°, and γ = 90°, while δ-Al2O3 exhibits a tetragonal structure with dimensions a = 7.9430 Å, b = 7.9430 Å, c = 23.5000 Å, α = 90°, β = 90°, and γ = 90°. Notably, the XRD peaks used for crystallite size calculation correspond specifically to γ-Al2O3 nanoparticles.
Distinct diffraction peaks are observed at specific 2θ values for γ-Al2O3 nanoparticles, 18.14°, 31.61°, 37.71°, and 45.08°, corresponding to crystal planes (111), (220), (311), and (400), respectively. Similarly, for δ-Al2O3 nanoparticles, peaks appear at 16.72°, 20.40°, 29.19°, 32.34°, 33.60°, 36.42°, 40.55°, 42.96°, 52.74°, 57.00°, and 62.24°, corresponding to crystal planes (103), (114), (107), (222), (118), (312), (226), (1110), (2210), (2212), and (4010), respectively. The absence of impurity peaks in the XRD spectra underscores the high purity of the Al2O3 nanoparticle samples [36].
Using the Debye–Scherrer equation, the average particle size was determined.
d = kλ/βcosθ
The standalone pure Al2O3 nanoparticles were analyzed for crystallite size using X-ray diffraction. The crystallite size (d) was determined to be 25.1 nm, where θ represents Bragg’s angle, λ denotes the wavelength of X-ray radiation (1.5406 Å), and β represents the full width at half maximum (FWHM) of the diffraction peak. In the equation d = kλ/β, the constant k is 0.94.

3.4. Morphological and Elemental Analysis

SEM images revealed that the Al2O3 NPs exhibit an almost spherical shape with irregularities, displaying agglomeration and density, as shown in Figure 5a [29]. The nanoparticles also demonstrate a uniform size distribution, ranging from 30 to 40 nm, with an average size of 33 nm.
The EDX spectra of the Al2O3 NPs validate their purity. Both aluminum (Al) and oxygen (O) are present in the sample, constituting weight percentages of 41.36% and 58.64%, respectively. The atomic percentages indicate 32.96% for aluminum and 67.04% for oxygen, affirming that the prepared Al2O3 NPs are essentially free from impurities.

3.5. Mechanical Properties of the Prepared Al2O3 NPs

3.5.1. Flexion Test

In Figure 6a,b, the flexural test, the plant sample derived from Calligonum comosum L. leaves exhibits notable stiffness with a Young’s modulus of 347.87 MPa. The composite sample (plant + Al2O3 NPs) outperforms this with an impressive Young’s modulus of 530.32 MPa (Table 1). The Young’s modulus of the hybrid composite (plant and Al2O3 NPs) showed a remarkable 34.4% increase in flexion. This highlights the enhanced stiffness achieved through the incorporation of Al2O3 NPs, offering valuable insights for applications prioritizing flexural strength and rigidity.

3.5.2. Compression Test

In Figure 6c,d, the compression testing, the plant sample shows strong inherent strength with a Young’s modulus of 1139.23 MPa. However, the composite sample (plant + Al2O3 NPs), treated with aluminum oxide, exhibits remarkable improvement with a Young’s modulus of 5242.88 MPa (Table 2). The Young’s modulus of the hybrid composite (plant and Al2O3 NPs) showed a remarkable 78.3% increase in compression. This underscores the significant enhancement in compressive properties achieved through the strategic incorporation of aluminum oxide, valuable for applications requiring high compressive strength and resistance to deformation.

3.6. Photocatalytic Degradation of Methylene Blue and Rose Bengal

The degradation mechanism of methylene blue (MB) and Rose Bengal (RB) via pure Al2O3 nanoparticles under solar irradiation is driven by a sophisticated interplay of photochemical reactions. These reactions, triggered by the intrinsic semiconductor properties of Al2O3 NPs upon exposure to solar light, initiate a sequence of events crucial for breaking down organic dye molecules. Initially, solar photons are absorbed by the Al2O3 NPs, generating electron–hole pairs within the semiconductor material.
These photoinduced electron–hole pairs migrate to the surface of the photocatalyst, where they participate in redox reactions with the adsorbed dye molecules. Specifically, electrons in the conduction band (CB) of Al2O3 NPs reduce the dye molecules, while holes in the valence band (VB) are involved in water oxidation. This process yields reactive oxygen species (ROS) such as hydroxyl radicals (OH) and superoxide radicals (O2•−), which act as potent oxidizing agents. These ROS initiate degradation by attacking the adsorbed MB and RB molecules, thereby initiating the breakdown process (see Equations (4)–(9)) [21,22,23,26] (Figure 7).
The degradation itself involves the cleavage of chemical bonds and chromophoric groups within the dye molecules, transforming them into smaller, less chromatic fragments. The presence of Al2O3 NPs significantly influences the overall efficiency of the photocatalytic process under solar light irradiation.
In summary, the photocatalytic degradation of MB and RB using pure Al2O3 NPs under solar light is governed by the absorption of photons, leading to the generation of electron–hole pairs, which in turn drive redox reactions with the dyes on the nanoparticle surface. This cascade of events culminates in the breakdown of the dye molecules into less complex fragments, highlighting the pivotal role of Al2O3 NPs in enhancing the efficiency of this photocatalytic procedure.
Al2O3 + hv → e(CB) + h+(VB)
h+(VB) + H2O → H+→ OH
e(CB) + O2→ O.∙−2
h+(CB) + OH → OH∙
OH + MB or RB (dye) → CO2 + H2O + other degradation products
O∙−2 + MB or RB (dye) → CO2 + H2O + other degradation products
The degradation of methylene blue (MB) and Rose Bengal (RB) under solar light using pure Al2O3 nanoparticles is visually depicted in Figure 8a–d. Initially, distinctive absorption peaks are observed at approximately 663 nm for MB and 542 nm for RB. These peaks gradually diminish as the reaction proceeds, indicating a continuous degradation process over an extended period. Figure 8 illustrates the absorption spectra, demonstrating a visible decrease in peak intensities for both dyes, accompanied by a fading of their vibrant colors over time.
The evaluation of photocatalytic activity shows that Al2O3 nanoparticles achieve degradation efficiencies of 98.2% and 90.5% for MB and RB, respectively, within 120 min. Notably, RB exhibits enhanced decomposition efficiency over time, consistent with findings reported in previous studies [28,31,37]. These results underscore the effective dye removal capability of Al2O3 nanoparticles (Table 3).

3.7. Recyclability and Stability

Researchers conducted experiments to evaluate the recyclability of δ-Al2O3 and γ-Al2O3 nanoparticles as photocatalysts for water purification, focusing on their ability to be separated and reused effectively. Each cycle involved drying the used photocatalysts for subsequent reuse under identical conditions as the initial cycle [39,40,41].
The results of five consecutive cycles of photocatalyst recycling are illustrated in Figure 9a,c. These findings demonstrate that δ-Al2O3 and γ-Al2O3 nanoparticles maintained high effectiveness and reusability in degrading Rose Bengal (RB) and methylene blue (MB) dyes through photodegradation. Figure 9b,d indicate a slight decrease in photocatalyst efficiency, reducing from 98.2% to 87.36% after five cycles for MB degradation, and from 90.5% to 80.32% after five cycles for RB degradation.
This decline is attributed to the inevitable loss of photocatalyst material during recycling processes such as the washing, centrifugation, or adsorption of intermediate species generated during photocatalysis [42,43]. X-ray diffraction (XRD) data in Figure 9e show consistent diffraction peaks of the Al2O3 nanoparticles’ photocatalyst before and after the photodegradation process across the five cycles. This stability underscores the suitability of Al2O3 nanoparticles for repeated use in water remediation applications [44,45].
It is worth noting that the slight decrease in decomposition efficiency over the cycles could potentially be mitigated with improved methods of catalyst recovery and reuse. Furthermore, the stability of the Al2O3 NPs’ photocatalyst, as evidenced by the consistent XRD diffraction peaks, is a promising indicator of their long-term viability for water remediation applications. However, further research may be needed to fully understand the long-term effects of repeated use and potential strategies for maintaining optimal performance [46,47]. It would also be interesting to explore the performance of these photocatalysts under different conditions and with different types of pollutants. This could provide a more comprehensive understanding of their potential applications in water remediation [48,49,50].

4. Conclusions

In this research, the successful synthesis of Al2O3 NPs using an extract from Calligonum comosum L. yielded spherical-shaped particles with a crystal structure, boasting a crystallite size of 25.1 nm. The determined band gap energy for these Al2O3 NPs stood at 2.86 eV. These findings underscore the importance of integrating Al2O3 NPs for targeted treatments of plant-derived materials. This study provides valuable insights applicable across diverse scenarios, particularly in applications requiring a delicate balance between flexural strength, rigidity, high compressive strength, and resistance to deformation. Notably, the Young’s modulus of the hybrid composite (plant and Al2O3 NPs) exhibited a significant 34.4% increase in flexion and a remarkable 78.3% increase in compression compared to the plant material alone. Such results not only emphasize the versatility of this innovative methodology but also highlight its potential impact on the materials industry. Furthermore, the incorporation of Al2O3 NPs emerges as a promising avenue for enhancing the performance of composites, with a specific focus on lightweight and high-performance materials. Notably, these biosynthesized Al2O3 NPs demonstrated exceptional efficacy in the removal of organic dyes. Under specific conditions, namely, a contact time of 120 min, pH = 7, and a solution temperature of 25 °C, the decomposition coefficients reached 98.2% for methylene blue (MB) and 90.5% for Rose Bengal (RB). These findings also suggest that Al2O3 NPs exhibit considerable potential as effective agents for wastewater treatment, particularly in the degradation of dyes. The successful synthesis and remarkable catalytic performance of these nanoparticles contribute valuable insights to the advancement of efficient and sustainable approaches for water treatment and pollution control.

Author Contributions

Conceptualization: A.H.G., H.H., S.E.L., A.B., I.B.A., M.T.G. and M.M.S.A.; Data Curation: S.E.L., A.B., T.T. and M.T.G.; Formal analysis: A.B., S.E.L., M.M.S.A. and M.T.G.; Investigation: S.E.L., A.B., M.T.G., M.M.S.A. and G.G.H.; Methodology: S.E.L., A.B. and M.T.G.; Project administration: S.E.L., J.A.A.A. and M.T.G.; Resources: S.E.L., A.B., M.T.G., G.G.H., J.A.A.A. and T.T.; Software: A.B. and M.T.G. Supervision: S.E.L.; Validation: A.B., S.E.L. and M.T.G.; Visualization: A.B. and S.E.L.; Writing—original draft: A.B., S.E.L., G.G.H. and M.T.G.; Writing—review and editing: A.B., S.E.L., G.G.H. and M.T.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Researchers Supporting Project number RSPD2024R688, King Saud University, Riyadh, Saudi Arabia.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Acknowledgments

The authors acknowledge the financial support through Researchers Supporting Project number RSPD2024R688, King Saud University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. del Carmen Gómez-Regalado, M.; Martín, J.; Santos, J.L.; Aparicio, I.; Alonso, E.; Zafra-Gómez, A. Bioaccumulation/bioconcentration of pharmaceutical active compounds in aquatic organisms: Assessment and factors database. Sci. Total Environ. 2023, 861, 160638. [Google Scholar] [CrossRef] [PubMed]
  2. Abba, E.; Shehu, Z.; Haruna, R.M. Green synthesis and characterization of CuO@SiO2 nanocomposite using gum Arabic (Acacia senegalensis) (L) against malaria vectors. Trends Sci. 2021, 18, 28. [Google Scholar] [CrossRef]
  3. Mishra, R.K.; Mentha, S.S.; Misra, Y.; Dwivedi, N. Emerging pollutants of severe environmental concern in water and wastewater: A comprehensive review on current developments and future research. Water-Energy Nexus 2023, 6, 74–95. [Google Scholar] [CrossRef]
  4. Bao, J.; Guo, S.; Fan, D.; Cheng, J.; Zhang, Y.; Pang, X. Sonoactivated Nanomaterials: A potent armament for wastewater treatment. Ultrason. Sonochemistry 2023, 99, 106569. [Google Scholar] [CrossRef] [PubMed]
  5. Zouari Ahmed, R.; Laouini, S.E.; Salmi, C.; Bouafia, A.; Meneceur, S.; Mohammed, H.A.; Chihi, S.; Alharthi, F.; Abdullah, J.A.A. Green synthesis of α-Fe2O3 and α-Fe2O3@Ag NC for degradation of rose Bengal and antimicrobial activity. Biomass Convers. Biorefinery 2023. [CrossRef]
  6. Koul, B.; Yadav, D.; Singh, S.; Kumar, M.; Song, M. Insights into the domestic wastewater treatment (DWWT) regimes: A review. Water 2022, 14, 3542. [Google Scholar] [CrossRef]
  7. Jiang, Z.; Cheng, B.; Zhang, L.; Zhang, Z.; Bie, C. A review on ZnO-based S-scheme heterojunction photocatalysts. Chin. J. Catal. 2023, 52, 32–49. [Google Scholar] [CrossRef]
  8. Sangor, F.I.; Al-Ghouti, M.A. Waste-to-value: Synthesis of nano-aluminum oxide (nano-γ- Al2O3) from waste aluminum foils for efficient adsorption of methylene blue dye. Case Stud. Chem. Environ. Eng. 2023, 8, 100394. [Google Scholar] [CrossRef]
  9. Marsili, V.; Mazzoni, F.; Marzola, I.; Alvisi, S.; Franchini, M. Intermittent water supply system rehabilitation through a multiphase methodology based on network analysis and hydraulic modeling. J. Water Resour. Plan. Manag. 2023, 149, 04023048. [Google Scholar] [CrossRef]
  10. Rehman, K.; Fatima, F.; Waheed, I.; Akash, M.S.H. Prevalence of exposure of heavy metals and their impact on health consequences. J. Cell. Biochem. 2018, 119, 157–184. [Google Scholar] [CrossRef]
  11. Pradhan, B.; Chand, S.; Chand, S.; Rout, P.R.; Naik, S.K. Emerging groundwater contaminants: A comprehensive review on their health hazards and remediation technologies. Groundw. Sustain. Dev. 2023, 20, 100868. [Google Scholar] [CrossRef]
  12. Miri, Z.; Haugen, H.J.; Loca, D.; Rossi, F.; Perale, G.; Moghanian, A.; Ma, Q. Review on the strategies to improve the mechanical strength of highly porous bone bioceramic scaffolds. J. Eur. Ceram. Soc. 2023, 44, 23–42. [Google Scholar] [CrossRef]
  13. Kam, O.R.; Bakouan, C.; Zongo, I.; Guel, B. Removal of Thallium from Aqueous Solutions by Adsorption onto Alumina Nanoparticles. Processes 2022, 10, 1826. [Google Scholar] [CrossRef]
  14. Gligor, O.; Mocan, A.; Moldovan, C.; Locatelli, M.; Crișan, G.; Ferreira, I.C. Enzyme-assisted extractions of polyphenols—A comprehensive review. Trends Food Sci. Technol. 2019, 88, 302–315. [Google Scholar] [CrossRef]
  15. Peralta-Videa, J.R.; Huang, Y.; Parsons, J.G.; Zhao, L.; Lopez-Moreno, L.; Hernandez-Viezcas, J.A.; Gardea-Torresdey, J.L. Plant-based green synthesis of metallic nanoparticles: Scientific curiosity or a realistic alternative to chemical synthesis? Nan-otechnol. Environ. Eng. 2016, 1, 4. [Google Scholar] [CrossRef]
  16. Rangasamy, P.; Hansiya, V.S.; Maheswari, P.U.; Suman, T.; Geetha, N. Phytochemical Analysis and Evaluation of In vitro Antioxidant and Anti-urolithiatic Potential of various fractions of Clitoria ternatea L. Blue Flowered Leaves. Asian J. Pharm. Anal. 2019, 9, 67–76. [Google Scholar] [CrossRef]
  17. Manikandan, V.; Jayanthi, P.; Priyadharsan, A.; Vijayaprathap, E.; Anbarasan, P.; Velmurugan, P. Green synthesis of pH-responsive Al2O3 nanoparticles: Application to rapid removal of nitrate ions with enhanced antibacterial activity. J. Photochem. Photobiol. A Chem. 2019, 371, 205–215. [Google Scholar] [CrossRef]
  18. Hassaan, M.A.; Meky, A.I.; Fetouh, H.A.; Ismail, A.M.; El Nemr, A. Central Composite Design and Mechanism of Antibiotic Ciprofloxacin Photodegradation under Visible Light by Green Hydrothermal Synthesized Cobalt-Doped Zinc Oxide Nanoparticles. Sci. Rep. 2024, 14, 9144. [Google Scholar] [CrossRef] [PubMed]
  19. Sharma, J.K.; Akhtar, M.S.; Ameen, S.; Srivastava, P.; Singh, G. Green synthesis of CuO nanoparticles with leaf extract of Calotropis gigantea and its dye-sensitized solar cells applications. J. Alloys Compd. 2015, 632, 321–325. [Google Scholar] [CrossRef]
  20. Rao, V.N.; Kwon, H.; Lee, Y.; Ravi, P.; Ahn, C.W.; Kim, K.; Yang, J.-M. Synergistic integration of MXene nanosheets with CdS@ TiO2 core@ shell S-scheme photocatalyst for augmented hydrogen generation. Chem. Eng. J. 2023, 471, 144490. [Google Scholar]
  21. Amor, I.B.; Hemmami, H.; Laouini, S.E.; Ahmed, S.; Mohammed, H.A.; Abdullah, J.A.A.; Azooz, E.A.; Al-Mulla, E.A.J.; Alharthi, F. Enhancing oxidant and dye scavenging through MgO-based chitosan nanoparticles for potential antioxidant coatings and efficient photocatalysts. Biomass Convers. Biorefinery 2023, 1–15. [Google Scholar] [CrossRef]
  22. Zidane, Y.; Laouini, S.E.; Bouafia, A.; Meneceur, S.; Tedjani, M.L.; Alshareef, S.A.; Almukhlifi, H.A.; Al-Essa, K.; Al-Essa, E.M.; Rahman, M.M.; et al. Green synthesis of multifunctional MgO@AgO/Ag2O nanocomposite for photocatalytic degradation of methylene blue and toluidine blue. Front. Chem. 2022, 10, 1083596. [Google Scholar] [CrossRef] [PubMed]
  23. Amarowicz, R.; Weidner, S.; Wójtowicz, I.; Karmac, M.; Kosinska, A.; Rybarczyk, A. Influence of low-temperature stress on changes in the composition of grapevine leaf phenolic compounds and their antioxidant properties. Funct. Plant Sci. Biot 2010, 4, 90–96. [Google Scholar]
  24. Kaur, P.; Arora, S.; Singh, R. Isolation, characterization and biological activities of betulin from Acacia nilotica bark. Sci. Rep. 2022, 12, 9370. [Google Scholar] [CrossRef] [PubMed]
  25. Vilas-Boas, Â.; Valderrama, P.; Fontes, N.; Geraldo, D.; Bento, F. Evaluation of total polyphenol content of wines by means of voltammetric techniques: Cyclic voltammetry vs differential pulse voltammetry. Food Chem. 2019, 276, 719–725. [Google Scholar] [CrossRef] [PubMed]
  26. Gherbi, B.; Laouini, S.E.; Meneceur, S.; Bouafia, A.; Hemmami, H.; Tedjani, M.L.; Thiripuranathar, G.; Barhoum, A.; Menaa, F. Effect of pH Value on the Bandgap Energy and Particles Size for Biosynthesis of ZnO Nanoparticles: Efficiency for Photocatalytic Adsorption of Methyl Orange. Sustainability 2022, 14, 11300. [Google Scholar] [CrossRef]
  27. Zlatić, G.; Martinović, I.; Pilić, Z.; Paut, A.; Mitar, I.; Prkić, A.; Čulum, D. Green inhibition of corrosion of aluminium alloy 5083 by Artemisia annua L. extract in artificial seawater. Molecules 2023, 28, 2898. [Google Scholar] [CrossRef] [PubMed]
  28. Nagarajan, P.; Subramaniyan, V.; Elavarasan, V.; Mohandoss, N.; Subramaniyan, P.; Vijayakumar, S. Biofabricated Aluminium Oxide Nanoparticles Derived from Citrus aurantium L.: Antimicrobial, Anti-Proliferation, and Photocatalytic Efficiencies. Sustainability 2023, 15, 1743. [Google Scholar]
  29. Natrayan, L.; Rao, Y.S.; Prasad, P.R.; Bhaskar, K.; Patil, P.P.; Abdeta, D.B. Biosynthesis-Based Al2O3 Nanofiller from Cymbopogon citratus Leaf/Jute/Hemp/Epoxy-Based Hybrid Composites with Superior Mechanical Properties. Bioinorg. Chem. Appl. 2023, 2023, 9299658. [Google Scholar] [CrossRef] [PubMed]
  30. Manogar, P.; Morvinyabesh, J.E.; Ramesh, P.; Jeyaleela, G.D.; Amalan, V.; Ajarem, J.S.; Allam, A.A.; Khim, J.S.; Vijayakumar, N. Biosynthesis and antimicrobial activity of aluminium oxide nanoparticles using Lyngbya majuscula extract. Mater. Lett. 2022, 311, 131569. [Google Scholar] [CrossRef]
  31. Songmene, V.; Khettabi, R.; Viens, M.; Kouam, J.; Hallé, S.; Morency, F.; Masounave, J.; Djebara, A. Mesure, Contrôle et Caractérisation des Nanoparticules: Procédure Appliquée à L’usinage et au Frottement Mécanique; Institut de Recherche Robert-Sauvé en Santé et en Sécurité du Travail (IRSST): Montréal, QC, Canada, 2014; 93p, ISBN 978-2-89631-716-5. [Google Scholar]
  32. Ghelani, D.; Faisal, S. Synthesis and characterization of Aluminium Oxide nanoparticles. Authorea 2022. [Google Scholar] [CrossRef]
  33. Roque-Ruiz, J.; Reyes-López, S. Synthesis of α- Al2O3 nanopowders at low temperature from aluminum formate by combustion process. J Mater. Sci Eng 2016, 6, 1–8. [Google Scholar]
  34. Djebaili, K.; Mekhalif, Z.; Boumaza, A.; Djelloul, A. XPS, FTIR, EDX, and XRD analysis of Al2O3 scales grown on PM2000 alloy. J. Spectrosc. 2015, 2015, 868109. [Google Scholar] [CrossRef]
  35. Prashanth, P.; Raveendra, R.; Krishna, R.H.; Ananda, S.; Bhagya, N.; Nagabhushana, B.; Lingaraju, K.; Naika, H.R. Synthesis, characterizations, antibacterial and photoluminescence studies of solution combustion-derived α- Al2O3 nanoparticles. J. Asian Ceram. Soc. 2015, 3, 345–351. [Google Scholar] [CrossRef]
  36. Karunakaran, C.; Anilkumar, P.; Gomathisankar, P. Photoproduction of iodine with nanoparticulate semiconductors and insulators. Chem. Cent. J. 2011, 5, 31. [Google Scholar] [CrossRef] [PubMed]
  37. Anna, K.K.; Bogireddy, N.K.R.; Agarwal, V.; Bon, R.R. Synthesis of α and γ phase of aluminium oxide nanoparticles for the photocatalytic degradation of methylene blue under sunlight: A comparative study. Mater. Lett. 2022, 317, 132085. [Google Scholar] [CrossRef]
  38. Ezeibe Anderson, U.; Nleonu Emmanuel, C.; Onyemenonu Christopher, C.; Arrousse Nadia, N.C.C. Photocatalytic activity of aluminium oxide nanoparticles on degradation of ciprofloxacin. Saudi J. Eng. Technol. 2022, 7, 132–136. [Google Scholar]
  39. Makofane, A.; Motaung, D.E.; Hintsho-Mbita, N.C. Green Synthesis of Silver Deposited on Copper Ferrite Nanoparticles for the Photodegradation of Dye and Antibiotics. Appl. Surf. Sci. Adv. 2024, 21, 100601. [Google Scholar] [CrossRef]
  40. Taghavi Fardood, S.; Moradnia, F.; Ganjkhanlu, S.; Ouni, L.; Ramazani, A.; Sillanpää, M. Green Synthesis and Characterization of Spinel CoAl2O4 Nanoparticles: Efficient Photocatalytic Degradation of Organic Dyes. Inorg. Chem. Commun. 2024, 167, 112719. [Google Scholar] [CrossRef]
  41. Chihi, S.; Bouafia, A.; Meneceur, S.; Laouini, S.E.; Ahmed, R.Z. Effect of precursor concentration on the bandgap energy and particles size for green synthesis of hematite α-Fe2O3 nanoparticles by the aqueous extract of Moltkia ciliata and evaluation of the antibacterial activity. Biomass Convers. Biorefinery 2023. [Google Scholar] [CrossRef]
  42. Thakur, A.; Kaur, H.; Aguedal, H.; Singh, V.; Singh, V.; Goel, G. Biogenic Synthesis of Fe-Doped Al2O3 Nanoparticles Using Eichhornia Crassipes for the Remediation of Toxicant Malachite Green Dye: Kinetic and Thermodynamic Studies. Inorg. Chem. Commun. 2024, 163, 112340. [Google Scholar] [CrossRef]
  43. Guo, J.; Van Bui, H.; Valdesueiro, D.; Yuan, S.; Liang, B.; Van Ommen, J. Suppressing the Photocatalytic Activity of TiO2 Nanoparticles by Extremely Thin Al2O3 Films Grown by Gas-Phase Deposition at Ambient Conditions. Nanomaterials 2018, 8, 61. [Google Scholar] [CrossRef] [PubMed]
  44. Tai, L.T.; Dat, N.M.; Nam, N.T.H.; An, H.; Huong, L.M.; Cong, C.Q.; Hai, N.D.; Phong, M.T.; Hieu, N.H. Green Synthesis of Copper Oxide Nanoparticles for Photodegradation of Malachite Green and Antibacterial Properties under Visible Light. Opt. Mater. 2023, 136, 113489. [Google Scholar] [CrossRef]
  45. Jansanthea, P.; Inyai, N.; Chomkitichai, W.; Ketwaraporn, J.; Ubolsook, P.; Wansao, C.; Wanaek, A.; Wannawek, A.; Kuimalee, S.; Pookmanee, P. Green Synthesis of CuO/Fe2O3/ZnO Ternary Composite Photocatalyst Using Grape Extract for Enhanced Photodegradation of Environmental Organic Pollutant. Chemosphere 2024, 351, 141212. [Google Scholar] [CrossRef] [PubMed]
  46. Zheng, A.L.T.; Abdullah, C.A.C.; Chung, E.L.T.; Andou, Y. Recent progress in visible light-doped ZnO photocatalyst for pollution control. Int. J. Environ. Sci. Technol. 2023, 20, 5753–5772. [Google Scholar] [CrossRef]
  47. Lau, G.E.; Che Abdullah, C.A.; Wan Ahmad, W.A.; Assaw, S.; Zheng, A.L. Eco-Friendly Photocatalysts for Degradation of Dyes. Catalysts 2020, 10, 1129. [Google Scholar] [CrossRef]
  48. Goutam, S.P.; Saxena, G.; Roy, D.; Yadav, A.K.; Bharagava, R.N. Green Synthesis of Nanoparticles and Their Applications in Water and Wastewater Treatment. In Bioremediation of Industrial Waste for Environmental Safety; Springer Singapore: Singapore, 2020; pp. 349–379. [Google Scholar]
  49. Hussain, R.T.; Hossain, M.S.; Shariffuddin, J.H. Green Synthesis and Photocatalytic Insights: A Review of Zinc Oxide Nanoparticles in Wastewater Treatment. Mater. Today Sustain. 2024, 26, 100764. [Google Scholar] [CrossRef]
  50. Roy, S.; Tran, T.A.; Chowdhury, Z.Z.; Prathibha, B.S. Wastewater Treatment Using Green Synthesis; CRC Press: Boca Raton, FL, USA, 2023; ISBN 9781003342830. [Google Scholar]
Figure 1. A schematic illustrating the process of synthesizing Al2O3 NPs using an extract from the Calligonum comosum L. plant.
Figure 1. A schematic illustrating the process of synthesizing Al2O3 NPs using an extract from the Calligonum comosum L. plant.
Coatings 14 00848 g001
Figure 2. UV-Vis spectra of synthesis materials: (a) Calligonum comosum L. extract, (b) Al2O3 nanoparticles, and (c) optical energy gap of Al2O3 nanoparticles.
Figure 2. UV-Vis spectra of synthesis materials: (a) Calligonum comosum L. extract, (b) Al2O3 nanoparticles, and (c) optical energy gap of Al2O3 nanoparticles.
Coatings 14 00848 g002
Figure 3. FTIR spectrum of prepared samples (Calligonum comosum L. extract and Al2O3 NPs).
Figure 3. FTIR spectrum of prepared samples (Calligonum comosum L. extract and Al2O3 NPs).
Coatings 14 00848 g003
Figure 4. XRD pattern of Al2O3 NPs.
Figure 4. XRD pattern of Al2O3 NPs.
Coatings 14 00848 g004
Figure 5. (a,b) SEM images of Al2O3 NPs, (c) particle size distributions of Al2O3 NPs, and (d) EDX of Al2O3 NPs.
Figure 5. (a,b) SEM images of Al2O3 NPs, (c) particle size distributions of Al2O3 NPs, and (d) EDX of Al2O3 NPs.
Coatings 14 00848 g005
Figure 6. Young’s modulus flexural test stress (a,b) and compressive test (c,d) for samples (plant and plant + Al2O3 NPs).
Figure 6. Young’s modulus flexural test stress (a,b) and compressive test (c,d) for samples (plant and plant + Al2O3 NPs).
Coatings 14 00848 g006
Figure 7. General photodegradation process mechanism.
Figure 7. General photodegradation process mechanism.
Coatings 14 00848 g007
Figure 8. Photocatalytic behavior of MB and RB using (ad) Al2O3 NPs at different irradiation times under UV-VIS irradiation.
Figure 8. Photocatalytic behavior of MB and RB using (ad) Al2O3 NPs at different irradiation times under UV-VIS irradiation.
Coatings 14 00848 g008
Figure 9. Recyclability of δ-Al2O3 and γ-Al2O3 NPs’ photocatalysts for degradation of dye. (a) MB dye, (c) RB dye, and (b,d) reusability (degradation) efficiency vs. number of cycles of photodegradation of MB and RB dye by δ-Al2O3 and γ-Al2O3 NPs, respectively. (e) XRD analysis of δ-Al2O3 and γ-Al2O3 NPs, pure and reused.
Figure 9. Recyclability of δ-Al2O3 and γ-Al2O3 NPs’ photocatalysts for degradation of dye. (a) MB dye, (c) RB dye, and (b,d) reusability (degradation) efficiency vs. number of cycles of photodegradation of MB and RB dye by δ-Al2O3 and γ-Al2O3 NPs, respectively. (e) XRD analysis of δ-Al2O3 and γ-Al2O3 NPs, pure and reused.
Coatings 14 00848 g009
Table 1. Young’s Modulus from Flexural Test for Plant-based Samples with and without Al2O3 NPs.
Table 1. Young’s Modulus from Flexural Test for Plant-based Samples with and without Al2O3 NPs.
SamplesPlant
Young’s modulus (MPa)347.87
SamplesPlant + Al2O3 NPs
Young’s modulus (MPa)530.32
Table 2. Young’s Modulus from Compressive Test for Plant-based Samples with and without Al2O3 NPs.
Table 2. Young’s Modulus from Compressive Test for Plant-based Samples with and without Al2O3 NPs.
SamplePlant
Young’s modulus (MPa)1139.23
SamplePlant + Al2O3 NPs
Young’s modulus (MPa)5242.88
Table 3. Comparison of Photocatalytic Degradation efficiency results for Al2O3 NPs compared to this work.
Table 3. Comparison of Photocatalytic Degradation efficiency results for Al2O3 NPs compared to this work.
Metal NPsMethod of SynthesisThe Particle Size (nm)Organic PollutantsTime (min)Dye Removal (%)Ref.
Al2O3green method28MB15089.1[28]
α-Al2O3sol–gel42MB4 h85[37]
γ-Al2O3precipitation36MB4 h91.6[37]
Al2O3sol–gel/MB3095.9[31]
Al2O3green method50–100nitrate in aqueous solution10594[17]
Al2O3chemical precipitation/ciprofloxacin1–5 h90[38]
Al2O3green method25.1MB12098.2This work
Rose Bengal90.5
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

Gharbi, A.H.; Laouini, S.E.; Hemmami, H.; Bouafia, A.; Gherbi, M.T.; Ben Amor, I.; Hasan, G.G.; Abdullah, M.M.S.; Trzepieciński, T.; Abdullah, J.A.A. Eco-Friendly Synthesis of Al2O3 Nanoparticles: Comprehensive Characterization Properties, Mechanics, and Photocatalytic Dye Adsorption Study. Coatings 2024, 14, 848. https://doi.org/10.3390/coatings14070848

AMA Style

Gharbi AH, Laouini SE, Hemmami H, Bouafia A, Gherbi MT, Ben Amor I, Hasan GG, Abdullah MMS, Trzepieciński T, Abdullah JAA. Eco-Friendly Synthesis of Al2O3 Nanoparticles: Comprehensive Characterization Properties, Mechanics, and Photocatalytic Dye Adsorption Study. Coatings. 2024; 14(7):848. https://doi.org/10.3390/coatings14070848

Chicago/Turabian Style

Gharbi, Ahlam Hacine, Salah Eddine Laouini, Hadia Hemmami, Abderrhmane Bouafia, Mohammed Taher Gherbi, Ilham Ben Amor, Gamil Gamal Hasan, Mahmood M. S. Abdullah, Tomasz Trzepieciński, and Johar Amin Ahmed Abdullah. 2024. "Eco-Friendly Synthesis of Al2O3 Nanoparticles: Comprehensive Characterization Properties, Mechanics, and Photocatalytic Dye Adsorption Study" Coatings 14, no. 7: 848. https://doi.org/10.3390/coatings14070848

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