Next Article in Journal
Effect of pH, COD, and HRT on the Performance of Microbial Fuel Cell Using Synthetic Dairy Wastewater
Previous Article in Journal
Monitoring Discharge in Vegetated Floodplains: A Case Study of the Piave River
Previous Article in Special Issue
Spatio-Temporal Model of a Product–Sum Simulation on Stream Network Based on Hydrologic Distance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Green Development of Titanium Dioxide Using Astragalus boeticus for the Degradation of Cationic and Anionic Dyes in an Aqueous Environment

1
Laboratory of Materials, Nanotechnology and Environment (LMNE), Faculty of Science, Mohammed V University, Rabat BP 1014, Morocco
2
Laboratory of Physical Chemistry of Inorganic and Organic Materials (LPCMIO), Ecole Normale Supérieure (ENS), Centre des Sciences des Matériaux (CSM), Mohammed V University, Rabat BP 4014, Morocco
3
Laboratory of Spectroscopy, Molecular Modeling, Materials, Nanomaterials, Water and Environment, CERNE2D, Faculty of Science, Mohammed V University, Rabat BP 1014, Morocco
*
Author to whom correspondence should be addressed.
Water 2023, 15(19), 3471; https://doi.org/10.3390/w15193471
Submission received: 4 September 2023 / Revised: 20 September 2023 / Accepted: 28 September 2023 / Published: 30 September 2023
(This article belongs to the Special Issue Advanced Technology for Smart Environment and Water Treatment)

Abstract

:
Wastewater discharge from the textile industry poses significant health problems for humans. As a result, the effluent waters are often rich in dyes, whose low natural decomposition capacity makes their treatment complex, thus contributing to environmental degradation. It becomes imperative to implement effective solutions for treating these contaminated waters, with a primary goal: to make them fit for human consumption. The present study focuses on the development of green TiO2 nanoparticles (TiO2-NP) using titanium (IV) isopropoxide as a precursor, along with the extract of Astragalus boeticus (A.B). These green TiO2 nanoparticles have been developed for use as highly efficient photocatalysts for the degradation of two types of dyes: Reactive Yellow 161 (RY161), an anionic dye, and Crystal Violet (CV), a cationic dye. The structural, microstructural, and optical properties of the synthesized material were characterized using XRD, FTIR, SEM, EDX, and UV-Vis methods. The results of these analyses revealed that the nanoparticles have a size of approximately 68 nm, possess an anatase structure, exhibit a spherical surface morphology, and have a band gap of 3.22 eV. The photocatalytic activity of the synthesized material demonstrated a 94.06% degradation of CV dye in a basic environment (pH = 10) within 30 min, with an initial CV concentration of 10 mg/L and a catalyst mass of 1 g/L. Additionally, it achieved a 100% degradation of RY161 dye in an acidic environment (pH = 4) within 90 min, with an initial RY161 concentration of 30 mg/L and a catalyst mass of 1 g/L. Furthermore, the recycling study indicated that the green TiO2 NPs catalyst could be effectively reused for up to five cycles. These experimental findings suggest that the developed TiO2 catalyst holds significant potential as an eco-friendly solution for remediating aqueous media polluted by both anionic and cationic dyes.

1. Introduction

In recent years, it has been observed that the Earth’s atmosphere has begun to deteriorate due to the increase in and rapid expansion of industrialization, urbanization, and population growth. These factors have led to the massive release of dangerous and undesirable substances, such as dyes, pesticides, heavy metals, and antibiotics [1,2]. The textile sector, which is a significant consumer of water, generates a high level of pollution in the aquatic environment, with discharges heavily contaminated by dyes. Among these dyes, Crystal Violet (CV), a cationic dye, and Reactive Yellow 161 (RY161), an anionic dye, exhibit persistent effects in aqueous media. CV was chosen for its widespread use in various industries, including the textile, paint, chemical manufacturing, and biotechnology sectors [3,4]. RY161, on the other hand, was selected for its commercial azo character, commonly used for dyeing cotton textiles [5].
In recent years, the treatment of contaminated waters has become a significant challenge due to the complexity of their chemical structure, particularly the presence of recalcitrant aromatic rings [6]. Conventional methods struggle to effectively degrade these pollutants. Therefore, urgent action is required to treat and ensure the safety of these contaminated waters before discharge.
Advanced oxidation processes (AOPs) have gained popularity for their ability to transform recalcitrant pollutants into biodegradable products, mineralize most pollutants into harmless compounds like CO2 and H2O, and consume less energy compared to methods like incineration. Over the past few decades, various techniques for treating organic contaminants in wastewater and air have been explored [7,8].
One promising approach is photodegradation using metal nanoparticle photocatalysts [9,10,11]. Nanoparticles (NPs) are tiny particles with dimensions between 1 and 100 nm, and metal oxide NPs exhibit remarkable chemical, physical, magnetic, electronic, and optical properties [10,12]. Several metal oxide nanoparticles have been synthesized, including ZnO [13,14], ZnAl2O4@ZnO [15], Ag2O-Ag/ZnAl-oxide [16], Fe [17], ZnO/AC [18], CoMnCrO4 [19], MgFeCrO4 [20], and TiO2 [10,12,21,22,23].
Among these, TiO2 stands out as a semiconducting photocatalyst with a wide bandgap, making it an excellent absorber in the ultraviolet range [24,25,26,27]. It is highly efficient and possesses key properties for wastewater treatment and medicinal purposes, including high chemical and thermal stability, cost-effectiveness, non-toxicity, and biocompatibility [10,28,29].
While chemical synthesis of nanoparticles can be toxic and environmentally harmful, green synthesis methods avoid the use of harmful and toxic agents [10]. Plant extracts offer a convenient and environmentally friendly approach to biosynthesis. They prevent nanoparticle agglomeration, enhance stability, and are cost-effective. Plant substances act as reducing and capping agents, making this method natural and safe [10,30,31].
Previous research has successfully prepared TiO2 nanoparticles using various plant extracts, such as Jatropha curcas L., Citrus Limetta, Cinnamon, Syzygium Cumini, Citrus Limon juice, and costus speciosus. These nanoparticles have found applications in areas like antioxidant and antimicrobial studies [18,32,33,34,35].
Based on the foregoing, this study focuses on the green synthesis of TiO2 NPs using titanium isopropoxide (TTIP) as a titanium source and Astragalus boeticus leaf extract as a reducing and protecting agent. The choice of this plant is justified by its numerous environmental and natural qualities, as well as the fact that it has never been used in wastewater treatment. Furthermore, it is worth noting that this plant is environmentally friendly, as evidenced by its historical use in various fields, including medicine [36,37], and its seeds have been used as a coffee substitute [38]. This plant offers the advantage of providing a safe and environmentally friendly extract for the reduction in and coating of nanoparticles, a crucial step in the preparation of nanoscale materials. Quantitative tests have shown that extracts from this plant are rich in natural antioxidants, such as polyphenols and flavonoids, which possess well-known antioxidant properties. These compounds play a crucial role in protecting materials against damage caused by free radicals [39] generated during the photodegradation process. The synergy between environmental benefits and the antioxidant properties of this plant makes it an ideal choice for our research.
The synthesized material was characterized and used as a catalyst for the photocatalytic degradation of both cationic (CV) and anionic (RY161) textile dyes in an aqueous medium under UV irradiation. These two dyes were selected due to their persistent effects. The effect of various parameters, such as the catalytic dose, solution pH, and initial dye concentration, was studied. Detailed information on the reuse of the TiO2 nanoparticles prepared after the photodegradation process was also provided.

2. Materials and Methods

2.1. Materials

The leaves of Astragalus boeticus (A.B) were meticulously gathered from the local fields situated in the Ain Aouda region of Rabat, Morocco. These leaves served as a vital source of natural materials for the subsequent experiments and studies. For the chemical components and substances involved in the experiments, Titanium (IV) isopropylate, also known as titanium (IV) isopropoxide (C12H28O4Ti), was procured from Sigma-Aldrich with a stated purity level of 97%. This compound played a central role in the synthesis or processes being conducted. Sodium hydroxide (NaOH) and hydrochloric acid (HCl, 37%) were obtained from Panreac Appli-chem. The Reactive Yellow dye utilized in the experiments was sourced from Dystar Textilfarben, GmbH, and Co. Deutschland KG D-6007 Frankfurt. Additionally, Crystal Violet dye was purchased from Sigma-Aldrich. To prepare the necessary solutions and reagents for the experiments, double-distilled water was used. This ensured the purity of the solvent, reducing the likelihood of any impurities affecting the results. It is worth noting that all the reagents were used as obtained without undergoing any additional purification processes.

2.2. Preparation of Astragalus boeticus Extract

The fresh leaves underwent a thorough washing process, first with tap water to remove surface impurities, followed by rinsing with double-distilled water to ensure cleanliness. Subsequently, they were left to air dry at room temperature for a period ranging from 4 to 6 days until they reached an appropriate level of dryness. Once dried, the leaves were finely ground into a powder by crushing them in a mortar using a porcelain pestle. The preparation of the leaf extract then involved taking 30 g of the obtained powder and mixing it with 150 mL of distilled water. This mixture was then heated to 80 °C for a duration of 80 min. The heat facilitated the extraction of the desired compounds from the leaves into the water, creating a solution of the leaf extract. Following the heating process, the obtained extract was carefully filtered through Whatman filter paper with a pore size of 0.45 µm.

2.3. Synthesis of TiO2 NPs by the Extract of Astragalus boeticus

TiO2 nanoparticles were synthesized through the hydrolysis of titanium isopropoxide (TTIP), which is the primary method for producing TiO2 NPs [40].
In this process, 75 mL of Astragalus boeticus (A.B) extract was slowly added drop by drop to a solution containing 75 mL of TTIP. The resulting mixture was stirred for 9 h at room temperature. After the stirring period, the mixture was subjected to centrifugation at 8000 rpm for 8 min to separate the two phases. The resulting nanoparticles (NPs) were then dried at 110 °C for 24 h and subsequently subjected to calcination at 570 °C in a muffle furnace (Nabertherm GmbH) for 3 h (see Figure 1).

2.4. Characterization Techniques

The TiO2 NPs underwent comprehensive characterization, including structural, morphological, and optical assessments. The crystalline phase of the synthesized titanium nanoparticles (TiO2 NPs) was determined using X-ray powder diffraction (XRD) with a Shimadzu XDR-6100 instrument, employing Cu K radiation (λ = 1.5406 Å). To identify the functional groups present in TiO2, Fourier-transform infrared spectroscopy (FTIR) analysis was performed in the range of 4000–400 cm−1 using a Bruker Alpha Platinum-ATR spectrometer with a KBr disk. Scanning electron microscopy and energy-dispersive analysis (EDX) were utilized to visualize the morphology of TiO2 NPs, employing a Quanta 200 instrument. UV-visible transmission spectra were acquired using a Jasco V-673 spectrophotometer.
Photocatalytic experiments with the nanoparticles involved the assessment of two dyes: one anionic, Reactive Yellow 161, and the other cationic, Crystal Violet. These experiments were conducted using a UV-5800 spectrophotometer from Metash.

2.5. Photodegradation Test of RY161 and CV Dyes by Elaborated Green TiO2 NPs

The photolysis and adsorption phenomena were assessed by monitoring the degradation efficiency of the organic compounds Reactive Yellow (RY61) and Crystal Violet (CV) in aqueous solutions using green-synthesized NPs.
The photocatalytic process for both dyes was conducted in a 500 mL photochemical reactor under visible light irradiation. Prior to the photodegradation, various amounts of TiO2 NPs were brought into contact with separately prepared 100 mL solutions of each dye. The pH of the solutions was adjusted to 4 and then to 10 using 0.1M HCl and NaOH solutions. The solutions were stirred magnetically in the dark for 30 min. This adsorption step was essential to immobilize the pollutants on the support surface.
Subsequently, the solution of RY161 with the semiconductor was exposed to visible light (125 W) and stirred for 3 h. Samples of 5 mL each were collected at 15 min intervals during the first hour and at 30 min intervals for the remaining two hours. The solutions were then centrifuged using a SIGMA laborzentrifugen 2–15. The concentration changes in both dyes during the photocatalytic process were monitored using a UV-Visible spectrophotometer, with maximum wavelengths of 435 nm and 580 nm for the RY161 and CV dyes, respectively.
Ultraviolet light was emitted by a Radium HRL 125W/230/E27 mercury vapor lamp with a wavelength between 254 and 380 nm.
The yield of photocatalytic degradation was calculated from Equation (1):
R % = C 0 C t C 0 100
where C0 represents the initial concentration of the aqueous solutions of RY161 and CV, and Ct is the concentration of the solutions (RY161 and CV) during irradiation [41].
The photocatalytic degradation kinetics of cationic and anionic dye was evaluated by the Langmuir–Hinshelwood kinetics equation [42], as presented below (Equation (2)):
L n   ( C 0 C t ) = k a p p   t + C o n s t a n t
Kapp is the apparent first-order constant (min−1).

2.6. Reusability Test

The reusability test involved the recovery and regeneration of the green TiO2 NPs’ catalyst following the photocatalytic degradation experiment. After the completion of the photocatalytic reaction, the synthesized photocatalyst was subjected to a thorough washing process using double-distilled water for a duration of 1 h. This washing step was essential to remove any residual contaminants, reaction by-products, or adsorbed species from the surface of the catalyst, ensuring its cleanliness and readiness for subsequent use. Once the washing process was completed, the TiO2 NPs catalyst was carefully collected and prepared for reuse. To do so, it was subjected to an oven-drying procedure, allowing it to dry thoroughly overnight at a temperature of 100 °C. This drying step aimed to remove any remaining moisture and ensure that the catalyst was in a completely dry state, ready to be employed in further photocatalytic degradation experiments.

3. Results and Discussion

3.1. Sample Characterization

3.1.1. XRD Analysis

The XRD spectrum of synthesized TiO2 NPs is shown in Figure 2. Notable peaks in a 2θ range of 10° < 2θ < 70° were observed at 25.36°, 36.79°, ° 37.86°, 38.22°, 48.08°, 53.93°, 54.97°, and 62.67°, corresponding to the following values of Miller indices (hkl): (101), (103), (004), (112), (200), (105), (211), and (213), respectively [43]. All diffraction peaks are well indexed to the pure anatase phase according to the standard JCPDS map No. 99-101-0679. This has been validated by similar results from several published works [7,44].
The Debye–Scherrer relationship is used to calculate the average grain size of NPs [10,45]. The measured grain size of the synthesized TiO2 nanoparticles was about 68 nm.
D = k λ β   c o s   c o s   θ    
where
D: Crystal size of the NPs;
λ: Wavelength of the X-ray source;
K: Constant crystal form factor;
Θ: Bragg diffraction angle;
β: Total angular width at half maximum (FWHM) of the XRD peaks recorded at diffraction angle 2θ.

3.1.2. FTIR Analysis

The FTIR spectra of TiO2 nanoparticles are depicted in Figure 3. The band observed around 500 cm−1 to 600 cm−1 can be attributed to the stretching vibration of Ti-O and Ti-O-Ti bonds within the TiO2 structure [46]. This band is indicative of the presence of chemical bonds within the material and is crucial for characterizing the TiO2 structure. Another significant peak is observed at 1645 cm−1, corresponding to the characteristic bending vibration of the—OH group. The presence of this hydroxyl group is important as it suggests an interaction with water or other compounds containing hydroxyl groups. This can have significant implications in various applications of TiO2, including its reactivity with water or its ability to adsorb compounds containing hydroxyl groups. Lastly, a broad band in the range of 3650–3100 cm−1 is seen, corresponding to the intermolecular interaction between the TiO2 surface and the hydroxyl group of water molecules [47]. This interaction is crucial in many TiO2 applications, particularly in heterogeneous catalysis and photocatalysis, where water is often present as a reactant or reagent. Analyzing this band can provide insights into the nature of the interaction between TiO2 and water, which is essential for understanding the reaction mechanisms involved in these processes.

3.1.3. SEM with EDX Analysis

The structure, morphology, and apparent particle size of the synthesized nanoparticles were analyzed by scanning electron microscopy (SEM). The SEM image of TiO2 NPs is presented in Figure 4a, which shows that the particles are uniformly distributed and spherical in shape.
The EDX spectrum of the NPs is presented in Figure 4. The inset in Figure 4b shows the weight and atomic percentages of Ti (63.55 wt% and 36.32 atomic %) and O (32.41 wt% and 54.59 atomic %). The presence of the titanium and oxygen peaks alone confirms the formation of pure anatase TiO2. The carbon (C) peak is due to the carbon support used during this analysis. No other impurity peaks were observed.

3.1.4. UV-Vis Analysis

The optical properties are essential for the photocatalytic study because they allow one to describe the number of photons absorbed during the photocatalytic treatment [12]. The absorption spectrum and Tauc graph of prepared samples that were analyzed by UV-vis spectroscopy are shown in Figure 5. The UV-vis absorption pattern shows that the highest absorption is in the range of 250–400 nm, as shown in Figure 5a. The band gap of the synthesized nanoparticles was calculated by the Tauc Equation (4) [48].
( α h ν ) 2 / n = A   ( h ν E g )
where:
α: Absorption coefficient,
A: Constant proportionality,
: Energy of the incident photon,
Eg: Energy of the band gap.
Figure 5. (a) UV-Vis absorbance spectrum and (b) Tauc’s plot (inset) of TiO2 nanoparticles.
Figure 5. (a) UV-Vis absorbance spectrum and (b) Tauc’s plot (inset) of TiO2 nanoparticles.
Water 15 03471 g005
The value of n is related to the kind of optical transition of the semiconductor (for the direct transition n = 1 and for the indirect transition n = 4) [10,49]. The calculated direct band gap of the synthesized TiO2 NPs is 3.22 eV, as shown in Figure 4b, which is also consistent with the literature [12,50].

3.2. Photocatalytic Activity of TiO2 NPs for Degradation of Dyes

In the context of this research theme, the study of the degradation of RY161 and CV and the impact of specific key parameters on the photocatalytic activity of TiO2 nanoparticles (TiO2 NPs) in dye degradation are explored. The mechanism of heterogeneous photocatalysis begins with the careful selection of an appropriate semiconductor, possessing the required electronic structure characterized by an empty conduction band and a filled valence band. This configuration facilitates the light-induced redox process, which plays a pivotal role in dye degradation [51]. The investigation starts by analyzing the influence of dye concentration in the solution and conducting a kinetic study to clarify how this variable affects the degradation process. Simultaneously, exploration extends to the effect of the mass of TiO2 NPs on the photocatalytic reaction, involving a kinetic study to evaluate how the quantity of photocatalyst impacts degradation efficiency. Additionally, examination delves into the impact of solution pH on the degradation process, encompassing another kinetic study to gain a deeper understanding of how variations in pH can modulate the photocatalytic activity of TiO2 NPs. Finally, the proposed mechanism is comprehensively addressed to explain how TiO2 NPs act as catalysts in dye degradation. This thorough investigation contributes to a profound understanding of the effectiveness of this dye degradation technology and the optimization of its operational parameters for a more efficient and environmentally friendly application.

3.2.1. Effect of Photolysis and Adsorption

A preliminary study of the degradation efficiency of two dyes, RY161 (30 mg/L) and CV (10 mg/L), was conducted. The study was divided into two parts: the first part involved direct photolysis of the two dyes (without the addition of a catalyst) under UV irradiation, and the second part focused on the adsorption process with the addition of 1 g/L of TiO2 NPs based on A.B extract, both in the absence and presence of UV irradiation (referred to as adsorption and photocatalysis, respectively). The initial pH of each solution was maintained without adjustment (pH RY161 = 6.1 and pH CV = 5.9).
The results, presented in Figure 6, indicate that after 3 h of direct exposure of the two dyes to UV radiation without the addition of catalytic particles, the degradation rate remained negligible. This highlights the crucial role of the catalyst in the reaction medium (Figure 6a,b). However, an adsorption removal rate of 26.4% for RY161 and 20.65% for CV was achieved after 3 h using the synthesized catalyst. Furthermore, significantly higher removal efficiency, reaching 99.2% for RY161 and 96.34% for CV, was obtained after 3 h of contact with the synthesized TiO2 NPs under UV irradiation.
Comparing the experimental results, it is evident that the degradation efficiency of the two target dyes is greatly improved when in contact with TiO2 NPs under UV irradiation (photocatalysis) compared to adsorption alone. Direct photolysis, on the other hand, did not exhibit any advantage in the degradation process of the two pollutants.
These findings confirm that TiO2 NPs based on Astragalus boeticus extract, in combination with UV irradiation, exhibit superior photocatalytic properties for RY161 and CV molecules compared to TiO2 NPs without UV.
Based on these results, the decision was made to further investigate the photocatalytic efficiency of TiO2 NPs using UV radiation and to identify the key factors that positively or negatively impact their efficiency.

3.2.2. Effect of TiO2 NPs Mass and Kinetic Study

It is well known that the rate of photocatalytic degradation is closely related to the appropriate catalyst dose [52].
In order to determine the effect of the added catalyst dose on the degradation of 30 mg/L of RY161 dye and 10 mg/L of CV, a variation of the catalyst mass from 0.5 to 2 g/L was tested while keeping the pH of the two solutions fixed (pHS(RY161) = 6.1 and pHS(CV) = 5.9).
The results presented in Figure 7a,c demonstrate that the degradation efficiency of both dyes increases with the augmentation of the catalyst’s mass. In fact, an increase in the quantity of synthesized TiO2 NPs from 0.5 to 1 g/L significantly enhances the degradation rate. This increase can be attributed to the multiplication of active reaction sites on the TiO2 NPs [53] and the heightened capture of photons at the catalyst’s surface. A substantial number of high-energy dynamic sites on the catalyst’s surface could overcome the resistance to the mass transfer of pollutants between the solid and aqueous phases, leading to an enhancement in degradation [54]. Above this value (1 g/L), the degradation rate decreases, which is likely due to an excess of catalyst, and the agglomeration of the grains between them, which masks a good part of the photosensitive surface [24].
This was confirmed by the determined value of the apparent rate constants. In fact, the Kapp values obtained for a dose of 2 g/L (0.00872 min−1) were significantly lower than that obtained at 0.5 and 1 g/L (0.02187 min−1, 0.02617 min−1 respectively) for CV (Figure 7d). Meanwhile, for the RY161 dye, a Kapp value of 0.5 g/L (0.01146 min−1) was lower than 1 g/L (0.04865 min−1) and 2 g/L (0.01636 min−1), which is quite normal because the amount chosen (0.5 g/L) is lower compared to the initial concentration (30 mg/L) Figure 7b.

3.2.3. Effect of Concentration and Kinetic Study

Generally, the concentration of pollutants is one of the key parameters affecting the rate of photocatalytic degradation as a high concentration of pollutants limits the penetration of photons on the catalyst surface [24].
The effect of the initial pollutant concentration on the photocatalytic activity of the two dyes was evaluated by varying the initial concentration from 30 to 60 mg/L for RY161 and from 5 to 15 mg/L for CV. The results are presented in Figure 8, which illustrates the degradation kinetics of the dyes as a function of radiation time in the presence of supported TiO2 NPs (1 g/L). It is observed that the higher the initial concentration of the dyes, the longer the time required for their degradation.
Furthermore, increasing the concentration of RY161 from 30 to 60 mg/L leads to a decrease in the rate of the photocatalytic reaction from 0.04865 to 0.01292 min−1 (Figure 8b). Similar behavior was observed for the cationic dye, with a decrease in the rate constant from 0.02563 to 0.02172 min−1 for concentrations ranging from 5 to 15 mg/L, respectively (Figure 8d).
This decrease in efficiency at high concentrations can be attributed to the reduction in OH radicals generated on the surface of the catalyst as the active sites become occupied by dye ions. Additionally, at higher concentrations, a significant amount of UV radiation can be absorbed by the dye molecules, reducing the transmittance of light. This is unfavorable for the absorption of light energy by the catalysts [42].

3.2.4. Effect of Solution pH and Kinetic Study

pH is an important factor that influences the surface properties of solids, surface charge, and particle aggregate sizes in water [55,56].
Its impact on the photocatalytic activity is important to allow the evaluation of the efficiency of the technique in the case of a pollutant that is partially ionized or carrying charged functions, and it is necessary to determine the electrostatic interactions that occur between the adsorbent and the adsorbate. Indeed, depending on the point of zero charge (PZC) of the material, the surface charge of the latter depends on the pH. Thus, for TiO2 where PZC = 6.5, the surface is positively and negatively charged as shown in Equations (5) and (6), respectively [57,58]:
pH < 6.5   TiOH 2 + TiOH + H +
pH > 6.5   TiOH TiO + H +
The photocatalytic degradation of RY161 (30 mg/L) and CV (10 mg/L) in the presence of synthesized TiO2 NPs at different pH levels (4, initial pH of the solution, and 10) was investigated.
Figure 9 illustrates the removal of RY161 and CV as a function of irradiation time for various pH values. The results indicate that photodegradation is more significant at acidic pH (pH = 4) for RY161 and at basic pH (pH = 10) for CV.
This can be explained by the charge nature of these two pollutants. In an acidic medium, a substantial adsorption of the anionic dye RY161 onto TiO2 nanoparticles is observed, likely due to the electrostatic attraction between the positive charge of the synthesized TiO2 NPs and the negative charge of the dye [55]. However, the rate of photocatalytic degradation decreases with increasing pH, transitioning from 0.0507 to 0.009954 min−1 (Figure 9b).
Conversely, the best adsorption and degradation were observed in a basic medium where TiO2 carries a negative charge, favoring the adsorption of the cationic dye CV. Increasing the pH from 4 to 5.9 resulted in an increase in the rate of photocatalytic reactions from 0.00972 to 0.02015 min−1, with the degradation process following the Langmuir–Hinshelwood (L-H) model (Figure 9d). However, when the CV solution was highly basic (pH = 10), 94.06% of the CV molecules were removed within the first 30 min, confirming the high degradation efficiency of the catalyst. Therefore, its degradation performance did not conform to the L-H model throughout the entire 180 min process.

3.2.5. Mechanism Proposed for the Photocatalytic Activity of TiO2 NPs

Under the effect of visible light and after the excitation of the semiconductor, the electrons of the valence band (BV) pass towards the conduction band (BC) whose energy of the photons is higher or equal to the difference of the energy between these two bands of TiO2 NPs, which involves the production of an electron-hole pair (e−,h+), as shown in Equation (7) (Figure 10) [59].
These charge carriers generate and perform redox reactions, and oxygen molecules are oxidized to give superoxide ions ( O 2 . )   Equation (8); the latter can react with hydrogen peroxide to form elements like ( OH . ) , ( OH )   and   ( O 2 )   Equation (9), while water molecules and hydroxyl anions are reduced to hydroxyl radicals ( OH . ) according to Equation (10) and Euation (11), respectively. These free radicals are highly reactive and non-selective species, capable of oxidizing persistent organic pollutants to form intermediates that will themselves react to full mineralization, forming degradation products such as (CO2) and (H2O) Equation (12) [8,44].
The schematic mechanism of the photocatalytic activity is presented in Figure 10, and the corresponding reactions are also presented below.
TiO 2 + h ϑ   TiO 2   ( h + , e )
O 2 + e O 2 .
O 2 . + H 2 O 2 OH . + OH + O 2
TiO 2 ( h + ) + H 2 O   TiO 2 + H + + OH .
TiO 2 ( h + ) + OH TiO 2 + OH .  
Organic   Matter + OH . intermediate   products + CO 2 + H 2 O  

3.3. Reusability of the catalyst TiO2 NPs

The reuse of TiO2 NPs synthesized from Astragalus boeticus extract has been verified. Indeed, five recycling attempts operating under optimal operating conditions (pHRY161 = 4, CRY161 = 30 ppm, m = 1 g/L and pHCV = 10, CCV = 10 ppm, m = 1 g/L) were tested to prove the stability of the synthesized catalyst (Figure 11).
The obtained results show that after the fifth cycle, the photocatalytic efficiency decreased by only 7% and 9% for RY161 (Figure 10a) and CV (Figure 10b), respectively.
Indeed, during the first two cycles for both target pollutants, the decomposition capacity remained stable at 100%, then from the third and fourth cycle, this photocatalytic oxidation rate decreased very slightly and reached 2% for RY161 and 4% for CV.
Moreover, a degradation capacity of 93% and 91% was obtained for RY161 and CV, respectively, after the fifth cycle. Therefore, TiO2 NPs are stable, possess excellent reusability, and have the potential to be used multiple times without a considerable change in activity.

3.4. Comparative Study of the Photocatalytic Degradation of Organic Pollutants by Nanomaterials Synthesized Using a Green Way

Photocatalytic degradation of various organic pollutants using catalytic TiO2 nanoparticles synthesized by the green route using plant extracts was examined by comparing the results obtained from each study (Table 1). The latter clearly shows that TiO2 NPs synthesized from Astragalus boeticus extract are one of the most efficient heterogeneous catalysts, with a percentage of degradation of 100% and 94.06%, respectively, for Reactive Yellow 161 and Crystal Violet for the degradation of persistent pollutants due to the presence of substances that function as a reducing and capping agent.

4. Conclusions

The green synthesis of TiO2 nanoparticles was carried out using an extract from the leaves of Astragalus boeticus, a plant widely available in recycling contexts, making it an even more environmentally friendly choice. This green synthesis method is cost-effective, non-toxic, natural, and environmentally friendly. The crystalline phase and morphology of TiO2 NPs were characterized through XRD, FTIR, UV-Vis, and SEM/EDX analyses. The characterization of the synthesized TiO2 NPs revealed that the nanoparticles have a size of approximately 68 nm, an anatase crystal structure with a spherical surface morphology, and a calculated band gap of 3.22 eV, which is in line with the literature. Photocatalytic activity demonstrated that the cationic pollutant rapidly degrades in a basic medium (pH = 10) with a yield of 94.06% within 30 min, while the anionic pollutant achieves a yield of 100% after 90 min in an acidic medium (pH = 4). A recyclability study revealed the effective reuse of the green TiO2 NPs catalyst for up to five cycles. This excellent reusability and photocatalytic activity demonstrate that the synthesized photocatalyst is not only environmentally friendly but also holds potential applications in water purification.
While the green synthesis of TiO2 nanoparticles using Astragalus boeticus leaf extract offers numerous advantages, it is essential to acknowledge certain limitations of this approach. Firstly, the long-term stability of TiO2 NPs in complex environments needs to be further investigated. Additionally, the present study focused on the degradation of specific dyes, and the efficiency of this photocatalyst may vary depending on the pollutants present in the water. Further research is required to assess its applicability to other pollutants. Lastly, the scale of TiO2 NP production should be considered for potential large-scale applications.

Author Contributions

Conceptualization, F.M. and A.E.Y.; methodology, M.E.A.; software, J.M. and H.S.; validation, F.M., M.E.A. and A.E.Y.; formal analysis, J.M.; investigation, H.S.; resources, F.M.; data curation, M.E.A.; writing—original draft preparation, F.M.; writing—review and editing, J.M.; visualization, A.E.Y. and M.T.; supervision, M.T.; project administration, M.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data presented in this article are available upon request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

TiO2 Titanium Dioxide
TiO2-NP Titanium Dioxide Nanoparticle
TTIP Titanium (IV) Isopropoxide
A.B Astragalus boeticus
RY161 Reactive Yellow 161
CV Crystal Violet
XRDX-Ray Diffraction
FTIRFourier-Transmission Infrared
SEMScanning Electron Microscopy
EDX Energy-Dispersive Analysis
UV-VisUltraviolet Visible
eVElectron Volt
pH Hydrogen Potential
NaOH Sodium Hydroxide
HCl Hydrochloric Acid
PZC Point of Zero Charge
OH. Hydroxyl Radical
AOPs Advanced Oxidation Processes

References

  1. Hasan, I.; Rana, A. A review on in SITU green synthesis of titanium dioxide nanoparticles and their photocatalytic activities. Mater. Today Proc. 2023, 81, 916–918. [Google Scholar] [CrossRef]
  2. Saravanan, A.; Kumar, P.S.; Vo, D.V.N.; Yaashikaa, P.R.; Karishma, S.; Jeevanantham, S.; Gayathri, B.; Bharathi, V.D. Photocatalysis for removal of environmental pollutants and fuel production: A review. Environ. Chem. Lett. 2021, 19, 441–463. [Google Scholar] [CrossRef]
  3. Al-Shahrani, S. Phenomena of Removal of Crystal Violet from Wastewater Using Khulays Natural Bentonite. J. Chem. 2020, 2020, 4607657. [Google Scholar] [CrossRef]
  4. Kulkarni, M.R.; Revanth, T.; Acharya, A.; Bhat, P. Removal of Crystal Violet dye from aqueous solution using water hyacinth: Equilibrium, kinetics and thermodynamics study. Resour. Technol. 2017, 3, 71–77. [Google Scholar] [CrossRef]
  5. Xie, W.; Zhang, M.; Liu, D.; Lei, W.; Sun, L.; Wang, X. Reactive yellow 161 decolorization by TiO2/porous boron nitridenanosheet composites in cotton dyeing effluent. ACS Sustain. Chem. Eng. 2017, 5, 1392–1399. [Google Scholar] [CrossRef]
  6. Ammari, Y.; Elatmani, K.; Qourzal, S.; Bakas, I.; Ejakouk, E.; Ait-Ichou, Y. Etude cinétique de la dégradation photocatalytique du colorant bleu de méthylène en présence de dioxyde de titane (TiO2), en suspension aqueuse. J. Mater. Environ. Sci. 2016, 7, 671–678. [Google Scholar]
  7. Srinivasan, M.; Venkatesan, M.; Arumugam, V.; Natesan, G. Green synthesis and characterization of titanium dioxide nanoparticles (TiO2 NPs) using Sesbania grandiflora and evaluation of toxicity in zebrafish embryos. Process Biochem. 2019, 80, 197–202. [Google Scholar] [CrossRef]
  8. El Yadini, A.; Marouane, B.; Ahmido, A.; Dunlop, P.; Byrne, J.A.; El Azzouzi, M.; El Hajjaji, S. Photolysis and photodegradation of fenamiphos insecticide by using slurry and supported TiO2. J. Mater. Environ. Sci. 2013, 4, 973–980. [Google Scholar]
  9. Goutam, S.P.; Saxena, G.; Singh, V.; Yadav, A.K.; Bharagava, R.N.; Thapa, K.B. Green synthesis of TiO2 nanoparticles using leaf extract of Jatropha curcas L. for photocatalytic degradation of tannery wastewater. Chem. Eng. J. 2017, 336, 386–396. [Google Scholar] [CrossRef]
  10. Nabi, G.; Majid, A.; Riaz, A.; Alharbi, T.; Arshad Kamran, M.; Al-Habardi, M. Green synthesis of spherical TiO2 nanoparticles using Citrus Limetta extract: Excellent photocatalytic water decontamination agent for RhB dye. Inorg. Chem. Commun. 2021, 129, 108618. [Google Scholar] [CrossRef]
  11. Iqbal, T.; Masood, A.; Khalid, N.R.; Tahir, M.B.; Asiri, A.M.; Alrobei, H. Green synthesis of novel lanthanum doped copper oxide nanoparticles for photocatalytic application: Correlation between experiment and COMSOL simulation. Ceram. Int. 2022, 48, 13420–13430. [Google Scholar] [CrossRef]
  12. Nabi, G.; Raza, W.; Tahir, M.B. Green Synthesis of TiO2 Nanoparticle Using Cinnamon Powder Extract and the Study of Optical Properties. J. Inorg. Organomet. Polym. Mater. 2020, 30, 1425–1429. [Google Scholar] [CrossRef]
  13. Iqbal, T.; Farman, S.; Afsheen, S.; Riaz, K.N. Novel study to correlate efficient photocatalytic activity of WO3 and Cr doped TiO2 leading to enhance the shelf-life of the apple. Appl. Nanosci. 2022, 12, 87–99. [Google Scholar] [CrossRef]
  14. Umar, A.; Chauhan, M.S.; Chauhan, S.; Kumar, R.; Kumar, G.; Al-Sayari, S.A.; Hwang, S.W.; Al-Hajry, A. Large-scale synthesis of ZnO balls made of fluffy thin nanosheets by simple solution process: Structural, optical and photocatalytic properties. J. Colloid Interface Sci. 2011, 363, 521–528. [Google Scholar] [CrossRef] [PubMed]
  15. Eskandari Azar, B.; Ramazani, A.; Taghavi Fardood, S.; Morsali, A. Green synthesis and characterization of ZnAl2O4@ZnO nanocomposite and its environmental applications in rapid dye degradation. Optik 2020, 208, 164129. [Google Scholar] [CrossRef]
  16. Shen, J.; Zeng, H.; Chen, C.; Xu, S. Applied Clay Science A facile fabrication of Ag 2 O-Ag/ZnAl-oxides with enhanced visible-light photocatalytic performance for tetracycline degradation. Appl. Clay Sci. 2020, 185, 105413. [Google Scholar] [CrossRef]
  17. Wang, T.; Jin, X.; Chen, Z.; Megharaj, M.; Naidu, R. Green synthesis of Fe nanoparticles using eucalyptus leaf extracts for treatment of eutrophic wastewater, Sci. Total Environ. 2014, 466–467, 210–213. [Google Scholar] [CrossRef]
  18. Mydeen, S.S.; Kumar, R.R.; Sambathkumar, S.; Kottaisamy, M.; Vasantha, V.S. Facile Synthesis of ZnO/AC Nanocomposites using Prosopis Juliflora for Enhanced Photocatalytic Degradation of Methylene Blue and Antibacterial Activity. Optik 2020, 224, 165426. [Google Scholar] [CrossRef]
  19. Moradnia, F.; Fardood, S.T.; Ramazani, A.; Osali, S.; Abdolmaleki, I. Green sol—gel synthesis of CoMnCrO 4 spinel nanoparticles and their photocatalytic application. Micro Nano Lett. 2020, 15, 674–677. [Google Scholar] [CrossRef]
  20. Moradnia, F.; Taghavi, S.; Ramazani, A.; Kumar, V. Green synthesis of recyclable MgFeCrO4 spinel nanoparticles for rapid photodegradation of direct black 122 dye. J. Photochem. Photobiol. A Chem. 2020, 392, 112433. [Google Scholar] [CrossRef]
  21. Muniandy, S.S.; Mohd Kaus, N.H.; Jiang, Z.T.; Altarawneh, M.; Lee, H.L. Green synthesis of mesoporous anatase TiO2 nanoparticles and their photocatalytic activities. RSC Adv. 2017, 7, 48083–48094. [Google Scholar] [CrossRef]
  22. Ahmad, W.; Jaiswal, K.K.; Soni, S. Green synthesis of titanium dioxide (TiO2) nanoparticles by using Mentha arvensis leaves extract and its antimicrobial properties. Inorg. Nano-Metal Chem. 2020, 50, 1032–1038. [Google Scholar] [CrossRef]
  23. Bhullar, S.; Goyal, N.; Gupta, S. Rapid green-synthesis of TiO2 nanoparticles for therapeutic applications. RSC Adv. 2021, 11, 30343–30352. [Google Scholar] [CrossRef]
  24. Halim, W. Synthèse Contrôlée par Additifs Latex de Nanoparticules Mésoporeuses à Base de TiO2: M (M=Pd, Ag, Cu, Ni…): Caractérisation et Applications en Photocatalyse. 01 Février 2021. Available online: https://tel.archives-ouvertes.fr/tel-03279447%0Ahttps://tel.archives-ouvertes.fr/tel-03279447/document (accessed on 27 September 2023).
  25. Saravanan, A.; Kumar, P.S.; Jeevanantham, S.; Karishma, S.; Kiruthika, A.R. Photocatalytic disinfection of micro-organisms: Mechanisms and applications. Environ. Technol. Innov. 2021, 24, 101909. [Google Scholar] [CrossRef]
  26. Liu, A.; Li, Z.; Wu, Z.; Xia, X. Talanta Study on the photocatalytic reaction kinetics in a TiO2 nanoparticles coated microreactor integrated micro fluidics device. Talanta 2018, 182, 544–548. [Google Scholar] [CrossRef]
  27. El Yadini, A.; Saufi, H.; Dunlop, P.S.M.; Byrne, J.A.; El Azzouzi, M.; El Hajjaji, S. Supported TiO2 on Borosilicate Glass Plates for Efficient Photocatalytic Degradation of Fenamiphos. J. Catal. 2014, 2014, 413693. [Google Scholar] [CrossRef]
  28. Alamelu, K.; Ja, B.M. TiO2-Pt composite photocatalyst for photodegradation and chemical reduction of recalcitrant organic pollutants. J. Environ. Chem. Eng. 2018, 6, 5720–5731. [Google Scholar] [CrossRef]
  29. Sethy, N.K.; Arif, Z.; Mishra, P.K.; Kumar, P. Green synthesis of TiO2 nanoparticles from Syzygium cumini extract for photo-catalytic removal of lead (Pb) in explosive industrial wastewater. Green Process. Synth. 2020, 9, 171–181. [Google Scholar] [CrossRef]
  30. Nasrollahzadeh, M.; Sajadi, S.M. Green synthesis, characterization and catalytic activity of the Pd/TiO2 nanoparticles for the ligand-free Suzuki—Miyaura coupling reaction. J. Colloid Interface Sci. 2016, 465, 121–127. [Google Scholar] [CrossRef] [PubMed]
  31. Sankar, R.; Manikandan, P.; Malarvizhi, V.; Fathima, T. Green synthesis of colloidal copper oxide nanoparticles using Carica papaya and its application in photocatalytic dye degradation. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2014, 121, 746–750. [Google Scholar] [CrossRef]
  32. Singh, S.; Maurya, I.C.; Tiwari, A.; Srivastava, P.; Bahadur, L. Green synthesis of TiO2 nanoparticles using Citrus limon juice extract as a bio-capping agent for enhanced performance of dye-sensitized solar cells. Surf. Interfaces 2022, 28, 101652. [Google Scholar] [CrossRef]
  33. Surya, C.; Arul, N.A.; Pandiyan, V.; Ravikumar, S.; Amutha, P.; Sobral, A.J.F.N.; Krishnakumar, B. Costus speciosus leaf extract assisted CS-Pt-TiO 2 composites: Synthesis, characterization and their bio and photocatalytic applications. J. Mol. Struct. 2019, 1195, 787–795. [Google Scholar] [CrossRef]
  34. Santhoshkumar, T.; Rahuman, A.A.; Jayaseelan, C.; Rajakumar, G.; Marimuthu, S.; Kirthi, A.V.; Velayutham, K.; Thomas, J.; Venkatesan, J.; Kim, S.K. Green synthesis of titanium dioxide nanoparticles using Psidium guajava extract and its antibacterial and antioxidant properties. Asian Pac. J. Trop. Med. 2014, 7, 968–976. [Google Scholar] [CrossRef] [PubMed]
  35. Varma, K.S.; Shukla, A.D.; Tayade, R.J.; Mishra, M.K.; Nguyen, V.H.; Gandhi, V. Interaction of levofloxacin with reverse micelle sol-gel synthesized TiO2 nanoparticles: Revealing ligand-to-metal charge transfer (LMCT) mechanism enhances photodegradation of antibiotics under visible light. Mater. Lett. 2022, 309, 131304. [Google Scholar] [CrossRef]
  36. Graziani, V.; Esposito, A.; Scognamiglio, M.; Chambery, A.; Russo, R.; Ciardiello, F.; Troiani, T.; Potenza, N.; Fiorentino, A.; D’Abrosca, B. Spectroscopic characterization and cytotoxicity assessment towards human colon cancer cell lines of acylated cycloartane glycosides from Astragalus boeticus L. Molecules 2019, 24, 1725. [Google Scholar] [CrossRef] [PubMed]
  37. Jaradat, N.A.; Zaid, A.N.; Abuzant, A.; Khalaf, S.; Abu-Hassan, N. Phytochemical and biological properties of four Astragalus species commonly used in traditional Palestinian medicine. Eur. J. Integr. Med. 2017, 9, 1–8. [Google Scholar] [CrossRef]
  38. Prohens, J.; Andújar, I.; Vilanova, S.; Plazas, M.; Gramazio, P.; Prohens, R.; Herraiz, F.J.; De Ron, A.M. Swedish coffee (Astragalus boeticus L.), a neglected coffee substitute with a past and a potential future. Genet. Resour. Crop Evol. 2014, 61, 287–297. [Google Scholar] [CrossRef]
  39. Yu, Z.C.; Zheng, X.T.; Lin, W.; He, W.; Shao, L.; Peng, C.L. Photoprotection of Arabidopsis leaves under short-term high light treatment: The antioxidant capacity is more important than the anthocyanin shielding effect. Plant Physiol. Biochem. 2021, 166, 258–269. [Google Scholar] [CrossRef]
  40. Drhimer, F.; Rahmani, M.; Regraguy, B.; El Hajjaji, S.; Mabrouki, J.; Amrane, A.; Fourcade, F.; Assadi, A.A. Treatment of a Food Industry Dye, Brilliant Blue, at Low Concentration Using a New Photocatalytic Configuration. Sustainability 2023, 15, 5788. [Google Scholar] [CrossRef]
  41. Kannan, K.; Radhika, D.; Vijayalakshmi, S.; Sadasivuni, K.K.; Ojiaku, A.A.; Verma, U. Facile fabrication of CuO nanoparticles via microwave-assisted method: Photocatalytic, antimicrobial and anticancer enhancing performance. Int. J. Environ. Anal. Chem. 2022, 102, 1095–1108. [Google Scholar] [CrossRef]
  42. Li, S.; Lin, Q.; Liu, X.; Yang, L.; Ding, J.; Dong, F.; Li, Y.; Irfan, M.; Zhang, P. Fast photocatalytic degradation of dyes using low-power laser-fabricated Cu2O-Cu nanocomposites. RSC Adv. 2018, 8, 20277–20286. [Google Scholar] [CrossRef] [PubMed]
  43. Al-taweel, S.S.; Saud, H.R. New route for synthesis of pure anatase TiO2 nanoparticles via utrasound- assisted sol-gel method. J. Chem. Pharm. Res. 2016, 8, 620–626. [Google Scholar]
  44. Ajmal, N.; Saraswat, K.; Bakht, M.A.; Riadi, Y.; Ahsan, M.J.; Noushad, M. Cost-effective and eco-friendly synthesis of titanium dioxide (TiO2) nanoparticles using fruit’s peel agro-waste extracts: Characterization, in vitro antibacterial, antioxidant activities. Green Chem. Lett. Rev. 2019, 12, 244–254. [Google Scholar] [CrossRef]
  45. Tesfaye, L.; Ramaswamy, K.; Bekele, B.; Saka, A.; Nagaprasad, N. Experimental investigation on the impacts of annealing temperatures on titanium dioxide nanoparticles structure, size and optical properties synthesized through sol-gel methods. Mater. Today Proc. 2021, 45, 5752–5758. [Google Scholar] [CrossRef]
  46. Bibi, S.; Ahmad, A.; Ali, M.; Anjum, R.; Haleem, A.; Siddiq, M.; Sakhawat, S.; Al, A. Photocatalytic degradation of malachite green and methylene blue over reduced graphene oxide (rGO) based metal oxides (rGO-Fe3O4/TiO2) nanocomposite under UV-visible light irradiation. J. Environ. Chem. Eng. 2021, 9, 105580. [Google Scholar] [CrossRef]
  47. Vetrivel, V.; Rajendran, K.; Kalaiselvi, V. Synthesis and characterization of Pure Titanium dioxide nanoparticles by Sol-gel method. Int. J. ChemTech Res. 2015, 7, 1090–1097. [Google Scholar]
  48. Moradnia, F.; Taghavi, S.; Ramazani, A.; Min, B.; Woo, S.; Varma, R.S. Magnetic Mg0.5Zn0.5FeMnO4 nanoparticles: Green sol-gel synthesis, characterization, and photocatalytic applications. J. Clean. Prod. 2021, 288, 125632. [Google Scholar] [CrossRef]
  49. Jiang, Z.; Liu, D.; Jiang, D.; Wei, W.; Qian, K.; Chen, M.; Xie, J. Bamboo leaf-assisted formation of carbon/nitrogen co-doped anatase TiO2 modified with silver and graphitic carbon nitride: Novel and green synthesis and cooperative photocatalytic activity. Dalton Trans. 2014, 43, 13792–13802. [Google Scholar] [CrossRef]
  50. Chandra, I.; Singh, S.; Senapati, S.; Srivastava, P.; Bahadur, L. Green synthesis of TiO2 nanoparticles using Bixa orellana seed extract and its application for solar cells. Sol. Energy 2019, 194, 952–958. [Google Scholar] [CrossRef]
  51. Saravanan, A.; Kumar, P.S.; Jeevanantham, S.; Anubha, M.; Jayashree, S. Degradation of toxic agrochemicals and pharmaceutical pollutants: Effective and alternative approaches toward photocatalysis. Environ. Pollut. 2022, 298, 118844. [Google Scholar] [CrossRef]
  52. Helali, S. Application de la photocatalyse pour la dégradation des polluants chimiques et bactériologiques dans l’eau en utilisant des catalyseurs irradiés par des photons de lumière naturelle ou artificielle (UV-A/UV-B). Doctoral dissertation, Université Claude Bernard-Lyon I, Villeurbanne, France, 2012. Français. NNT: 2012LYO10316.. [Google Scholar]
  53. Atrak, K.; Ramazani, A.; Taghavi, S. Eco-friendly synthesis of Mg0.5Ni0.5AlxFe2-xO4 magnetic nanoparticles and study of their photocatalytic activity for degradation of direct blue 129 dye. J. Photochem. Photobiol. A Chem. 2019, 382, 111942. [Google Scholar] [CrossRef]
  54. Saravanan, A.; Karishma, S.; Kumar, P.S.; Varjani, S.; Yaashikaa, P.R.; Jeevanantham, S.; Ramamurthy, R.; Reshma, B. Simultaneous removal of Cu(II) and reactive green 6 dye from wastewater using immobilized mixed fungal biomass and its recovery. Chemosphere 2021, 271, 129519. [Google Scholar] [CrossRef] [PubMed]
  55. Alahiane, S.; Qourzal, S.; El Ouardi, M.; Belmouden, M.; Assabbane, A. Adsorption et photodégradation du colorant indigo carmine en milieu aqueux en présence de TiO2/UV/O2 (Adsorption and photocatalytic degradation of indigo carmine dye in aqueous solutions using TiO2/UV/O2). J. Mater. Environ. Sci. 2013, 4, 239–250. [Google Scholar]
  56. Elsellami, L.; Vocanson, F.; Dappozze, F.; Puzenat, E.; Païsse, O.; Houas, A.; Guillard, C. Kinetic of adsorption and of photocatalytic degradation of phenylalanine effect of pH and light intensity. Appl. Catal. A Gen. 2010, 380, 142–148. [Google Scholar] [CrossRef]
  57. Preo, T.; Kallay, N. Point of Zero Charge and Surface Charge Density of TiO 2 in Aqueous Electrolyte Solution as Obtained by Potentiometric Mass Titration. Croat. Chem. Acta 2006, 79, 95–106. [Google Scholar]
  58. Azeez, F.; Al-hetlani, E.; Arafa, M.; Abdelmonem, Y.; Nazeer, A.A.; Amin, M.O.; Madkour, M. The effect of surface charge on photocatalytic degradation of methylene blue dye using chargeable titania nanoparticles. Sci. Rep. 2018, 8, 7104. [Google Scholar] [CrossRef]
  59. Bessekhouad, Y. Propriétés photocatalytiques de TiO2 nanocristallins dopés par des cations (Li+, Na+ et K+) et des hétérojonctions à base de sulfures et d’oxydes métalliques/TiO2; Autre; Université Paul Verlaine: Metz, France, 2003; Français. NNT: 2003METZ017S. tel-01749893. [Google Scholar]
  60. Pushpamalini, T.; Keerthana, M.; Sangavi, R.; Nagaraj, A.; Kamaraj, P. Comparative analysis of green synthesis of TiO2 nanoparticles using four different leaf extract. Mater. Today Proc. 2020, 40, S180–S184. [Google Scholar] [CrossRef]
Figure 1. Synthesis of TiO2 NPs from Astragalus boeticus leaf extract.
Figure 1. Synthesis of TiO2 NPs from Astragalus boeticus leaf extract.
Water 15 03471 g001
Figure 2. X-ray diffraction (XRD) pattern of synthesized TiO2 NPs.
Figure 2. X-ray diffraction (XRD) pattern of synthesized TiO2 NPs.
Water 15 03471 g002
Figure 3. FTIR spectrum of TiO2 NPs.
Figure 3. FTIR spectrum of TiO2 NPs.
Water 15 03471 g003
Figure 4. Structure and morphology analysis of TiO2 NPs synthesized using Astragalus boeticus leaf extract by (a) SEM analysis and (b) EDX profile determination.
Figure 4. Structure and morphology analysis of TiO2 NPs synthesized using Astragalus boeticus leaf extract by (a) SEM analysis and (b) EDX profile determination.
Water 15 03471 g004
Figure 6. Representation of adsorption, photolysis and photocatalysis during the degradation process of RY161 (a) and CV (b).
Figure 6. Representation of adsorption, photolysis and photocatalysis during the degradation process of RY161 (a) and CV (b).
Water 15 03471 g006
Figure 7. Effect of TiO2 NPs mass (a,c) and kinetic (b,d) on the photocatalytic degradation efficiency of RY161 (a,b) and CV (c,d) dyes.
Figure 7. Effect of TiO2 NPs mass (a,c) and kinetic (b,d) on the photocatalytic degradation efficiency of RY161 (a,b) and CV (c,d) dyes.
Water 15 03471 g007
Figure 8. Effect of different concentrations and kinetic of RY161 (a,b) and CV (c,d).
Figure 8. Effect of different concentrations and kinetic of RY161 (a,b) and CV (c,d).
Water 15 03471 g008
Figure 9. Effect of solution pH and kinetic study of RY161 (a,b) and CV (c,d) dyes.
Figure 9. Effect of solution pH and kinetic study of RY161 (a,b) and CV (c,d) dyes.
Water 15 03471 g009
Figure 10. Photocatalytic mechanism of the dye using TiO2 NPs.
Figure 10. Photocatalytic mechanism of the dye using TiO2 NPs.
Water 15 03471 g010
Figure 11. The reusability graph of RY161 (a) and CV (b) degraded.
Figure 11. The reusability graph of RY161 (a) and CV (b) degraded.
Water 15 03471 g011
Table 1. Comparative study of previous and current data on photocatalytic degradation of pollutants with green-synthesized TiO2 nanomaterials.
Table 1. Comparative study of previous and current data on photocatalytic degradation of pollutants with green-synthesized TiO2 nanomaterials.
Plant Used for the Synthesis of TiO2 NPsPollutant
Conc.
Catalyst
Conc.
Time (min)
Percentage
MorphologySize (nm)Ref
Jatropha curcasTannery wastewater
(Cr)
6.88 mg/L
5 g/5 L5 h
(76.48%)
Spherical75 nm[9]
Citrus limettaRhB dye
10 mg/L
0.7 g/50 mL80 min
(90%)
Spherical80–100 nm[10]
Syzygium cuminiLead (Pb)
8.621 mg/L
0.3 g/500 mL17 h
(82.53%)
Spherical10 nm[29]
Citrus limon juiceRG-19
6.7 Mm
0.03 g/100 mL60 min
(99.88%)
Spherical10–21 nm[32]
Piper betelVert Malachite
100 ppm
100 mg/50 mL50 min
(100%)
-6.6 nm[60]
Ocimum tenuiflorm50 min
(100%)
7.0 nm
Moringa oleifera30 min
(100%)
6.6 nm
Coriandrum sativum50 min
(100%)
6.8 nm
Astragalus boeticusRY161
30 mg/L
CV
10 mg/L
0.1 g/100 mL90 min
(100%)
30 min
(94.06%)
Spherical68 nmThis work
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

Maanane, F.; El Yadini, A.; El Alouani, M.; Mabrouki, J.; Saufi, H.; Tabyaoui, M. Green Development of Titanium Dioxide Using Astragalus boeticus for the Degradation of Cationic and Anionic Dyes in an Aqueous Environment. Water 2023, 15, 3471. https://doi.org/10.3390/w15193471

AMA Style

Maanane F, El Yadini A, El Alouani M, Mabrouki J, Saufi H, Tabyaoui M. Green Development of Titanium Dioxide Using Astragalus boeticus for the Degradation of Cationic and Anionic Dyes in an Aqueous Environment. Water. 2023; 15(19):3471. https://doi.org/10.3390/w15193471

Chicago/Turabian Style

Maanane, Fadwa, Adil El Yadini, Marouane El Alouani, Jamal Mabrouki, Hamid Saufi, and Mohamed Tabyaoui. 2023. "Green Development of Titanium Dioxide Using Astragalus boeticus for the Degradation of Cationic and Anionic Dyes in an Aqueous Environment" Water 15, no. 19: 3471. https://doi.org/10.3390/w15193471

APA Style

Maanane, F., El Yadini, A., El Alouani, M., Mabrouki, J., Saufi, H., & Tabyaoui, M. (2023). Green Development of Titanium Dioxide Using Astragalus boeticus for the Degradation of Cationic and Anionic Dyes in an Aqueous Environment. Water, 15(19), 3471. https://doi.org/10.3390/w15193471

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