Next Article in Journal
Implications of Water Quality Index and Multivariate Statistics for Improved Environmental Regulation in the Irtysh River Basin (Kazakhstan)
Previous Article in Journal
Flood Extent Delineation and Exposure Assessment in Senegal Using the Google Earth Engine: The 2022 Event
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Toxic Congo Red Dye Photodegradation Employing Green Synthesis of Zinc Oxide Nanoparticles Using Gum Arabic

by
Huda S. Alhasan
1,*,
Alaa R. Omran
1,
Abdullah Al Mahmud
2,
Amr Hussein Mady
2,3 and
Mohammad R. Thalji
4,*
1
Environmental Research and Studies Center, University of Babylon, Hilla 51002, Iraq
2
School of Chemical Engineering, Yeungnam University, 280 Daehak-ro, Gyeongsan 38541, Gyeongbuk, Republic of Korea
3
Petrochemical Department, Egyptian Petroleum Research Institute, Nasr City 11727, Cairo, Egypt
4
Korea Institute of Energy Technology (KENTECH), 200 Hyeokshin-ro, Naju 58330, Jeollanam-do, Republic of Korea
*
Authors to whom correspondence should be addressed.
Water 2024, 16(15), 2202; https://doi.org/10.3390/w16152202
Submission received: 2 July 2024 / Revised: 27 July 2024 / Accepted: 29 July 2024 / Published: 2 August 2024

Abstract

:
A green synthesis method for producing zinc oxide nanoparticles (ZnO NPs) was presented using natural Gum Arabic (GA) as a natural stabilizing agent. For the first time, the as-synthesized ZnO NPs were employed to photodegrade the toxic Congo Red (CR) dye in an aqueous solution. The structural and morphological characterizations confirmed the successful synthesis of ZnO NPs. The ZnO NPs possessed an average crystallite size of 42.7 nm. In addition, it was found that a concentration of 20 mg L−1 of CR dye yielded the most favorable photodegradation results, and 4 mg mL−1 of the photocatalyst was the optimal amount. The results showed a maximum degradation percentage of 99.5% at pH 8 after 30 min of irradiation. This indicates that the as-synthesized ZnO NPs have remarkable photocatalytic properties. Moreover, the study demonstrated the suitability of the pseudo-first-order kinetic model for representing the photodegradation process through kinetic studies of the photocatalyst process of CR dye by ZnO NPs using the Langmuir-Hinshelwood (L-H) model.

1. Introduction

Industrial developments in several sectors have significantly impacted environmental safety. One such sector is the production of dyes, which contaminate the aquatic environment when released into various water sources [1,2]. They are widely used in textiles, food, medicine, cosmetics, household products, and agricultural items [3,4,5]. Even in low quantities, they can cause severe diseases due to their hazardous properties, complex structure, and challenges associated with their decomposition [6,7]. Therefore, purifying the dye wastewater is a necessity and an urgent priority [8].
The complexity and resistance of dye molecules to standard treatment technologies have posed a significant challenge to their removal from wastewater [9]. Adsorption, coagulation, biodegradation, photocatalysis, and membrane filtering are among the techniques considered [10]. However, most of them often have limitations [11,12,13,14,15]. They can be costly due to the requirement for specialized equipment, energy consumption, or expensive materials [16]. Additionally, some processes might have slow kinetics, resulting in longer treatment times or requiring higher energy inputs to speed up the reactions [17]. Therefore, researchers are persistently exploring strategies to overcome these limitations. Currently, novel technologies and adaptations for several techniques are being developed to boost efficacy, minimize expenses, accelerate response times, and address challenges. Given the significant impact of dye wastewater on water quality, ongoing research is dedicated to developing more efficient, cost-effective, and long-term treatment solutions to address the shortcomings of previous procedures.
The photocatalytic degradation of dyes via semiconductor photocatalyst materials has emerged as an efficient method for removing dyes from wastewater [18]. It can degrade a wide range of organic and inorganic pollutants while avoiding the generation of undesirable byproducts [19,20,21]. Among semiconductor materials, zinc oxide (ZnO) is a superior choice for degrading organic dyes. It possesses a suitable band gap, rapid photocatalytic activity, low toxicity, easy preparation, and high efficiency [22,23]. It has a large exciton binding energy (60 meV) [24] like TiO2, suggesting a more stable and longer lifetime as a catalyst. Furthermore, ZnO has many active sites, making it a highly active photocatalyst. Therefore, it is ideal for removing toxic dyes from aqueous solutions. However, the synthesis methods used to create these nanoparticles should prioritize economic feasibility and sustainability. The size of these nanoparticles is also crucial, as it can be influenced by various factors such as temperature, stabilizers, and catalysts. These factors can impact the size of the nanoparticles, potentially leading to aggregation and the formation of larger particles due to thermodynamic instability [25,26,27]. The fast-paced development of nanomaterial production using different strategies has significantly increased the use of ZnO NPs in various applications. However, this progress has also brought several environmental issues to the forefront. As a result, there is a pressing need to find alternative synthesis strategies that can help mitigate these environmental problems. One promising approach is the green synthesis of ZnO using natural resources, such as plant extracts, microbes, and algae, as bio-stabilizer agents. By developing and implementing such green synthesis strategies, we can effectively address the environmental contamination caused by the high consumption of toxic chemicals and the discharge of hazardous waste in synthesizing ZnO NPs.
Several synthesis methods use stabilizer materials to prevent aggregation by minimizing surface energy [28]. Generally, ZnO material is not stable. Due to water’s high polarity, ZnO can agglomerate, leading to deposition. Several methods have been employed to improve its stability, including direct and homogeneous precipitations, hydrothermal/solvothermal methods, the sonochemical method, reverse micelles, thermal decomposition, and the sol-gel method. Unfortunately, most previous techniques utilized toxic chemicals as reducing and/or capping agents, generating environmentally risky by-products. Thus, this issue can be addressed by using green synthesis methods to achieve more sustainable processes, focusing on reducing or preventing the use of toxic and risky substances. In this pursuit, researchers have used several stabilizer materials, such as natural polymers [29,30]. Gum Arabic (GA), a naturally derived polymer, is utilized in several applications. It is produced from plants and has been extensively documented in many applications due to its remarkable characteristics, including its increased viscosity and stability, lack of toxicity, environmentally friendly nature, easy accessibility, cost-effectiveness, and surface activity [31,32]. It can serve the twin purposes of stabilizing and reducing metal ion agents during nanoparticle production. Depending on the pH of the solution, the polymer either forms a stiff barrier or a charge patch when it binds to the particle’s surface [33,34]. Several research studies have been published on synthesizing ZnO NPs utilizing GA derived from various plants, including cashew gum, frankincense gum, and acacia Arabic. The studies have shown impressive efficacy in preparing ZnO NPs and remarkable practical application results in the degradation of several dyes, such as direct blue 129 (DB129) and Malachite Green [35,36,37,38]. However, there are no reports in the literature about the green synthesis of ZnO NPs using GA for toxic Congo Red dye photodegradation in an aqueous solution.
Congo Red (CR) is an anionic diazo dye with the chemical formula C32H22N6Na2O6S2 [39,40]. Several studies have evidenced that its presence in the human body might have harmful effects, resulting in various illnesses that, if severe, could potentially be fatal [37], as displayed in Table 1 [41,42,43]. Possible consequences include mutagenicity, carcinogenicity, cytotoxicity, neurotoxicity, chemotoxicity, and genotoxicity. The eyes, skin, lungs, and reproductive systems are all impacted. The structural stability of the CR dye makes it non-biodegradable. Consequently, it is classified as an organic contaminant, and its removal from wastewater is necessary.
In this work, we fabricated ZnO NPs using GA material as a green synthesis method. The GA material played a crucial role as a stabilizing and reducing agent. Notably, the as-prepared ZnO NP catalyst is used for the first time to effectively photodegrade toxic CR dye in an aqueous solution. Our photodegradation studies, considering parameters such as pH level, contact time, dye concentration, and catalyst dosage, demonstrated an impressively high level of photocatalytic activity against the CR dye. Importantly, we achieved a remarkable degradation percentage of 99.5% when the concentration of the CR dye was 20 mg L−1, underscoring the potential of ZnO NPs as a catalyst.

2. Materials and Methods

2.1. Materials

Zinc (II) sulfate hexahydrate (ZnSO4.6H2O, >98%, Sigma-Aldrich, St. Louis, MO, USA), Congo Red dye (CR, Thermo Fisher Scientific, Waltham, MA, USA), sodium hydroxide pellets (NaOH, 98.0%, Loba Chemie, Mumbai, India), hydrochloric acid (HCl, 36.5%, Fluka Chemika, Seelze, Germany), and ethanol (C2H5OH, 99.9%, Techno Pharmchem, Rohtak, India) were used without further purification. Solutions were prepared by dissolving the appropriate amount of the dye in double-distilled water before each experiment.

2.2. Collection and Purification of Gum Arabic (GA)

The GA sample was collected from the Al-Mahawil district in the local area (Babylon Governorate, Iraq) based on its availability and effectiveness. After that, it was dried appropriately under sunlight and ground to a fine powder using a mortar and pestle. The GA solution (1% w/v) was prepared and purified using distilled water and hydrated for 24 h at 3 °C. The solution was then mixed with isopropanol at a 1:2 ratio and held at 3 °C for 6 h. After that, the solution was centrifugated, and the precipitate was retained. Finally, the precipitate was dried under a vacuum for 48 h.

2.3. Synthesis of ZnO NPs

A solution of GA (40 mL, 1.0% (w/v)) was added to 50 mL of 0.1 M ZnSO4 6H2O. The mixture was subjected to stirring for 30 min. The pH of the mixture was adjusted to 10 by slowly adding a 1 M NaOH solution. Subsequently, it was agitated for 180 min at room temperature. The colloidal solution acquired underwent a 24 h aging process at room temperature and was collected by centrifugation. The resultant product was washed with C2H5OH and distilled water, ensuring no impurities remained, and then dried at 45 °C in an electrical oven. The dried sample underwent calcination for 1 h at 400 °C using a muffle furnace [44]. Figure 1 shows the synthesis method for the ZnO NPs using GA material.

2.4. Material Characterization

The X-ray diffractometer (XRD) used in this study was the PW-1730 Philips model (Almelo, Netherlands), equipped with filtered Cu-Kα radiation with a wavelength of 0.15406 nm. The Bruker Vertex 70 Fourier transform infrared (FT-IR) system, Shimadzu 1800, Japan, was used to analyze the structure of GA and ZnO NPs at a resolution of 4 cm−1 throughout the 400–4000 cm−1 spectral region. The surface morphology of the GA material and ZnO NPs was thoroughly investigated using scanning electron microscopy (SEM, Tescan Vega, Czech). The absorbances for each sample were measured using a Perkin-Elmer UV-visible spectrophotometer with a wavelength range of 190 to 1100 nm. A quartz cell with a path length of 10 mm was used for the measurements, ensuring accurate and reliable data.

2.5. Photocatalytic and Kinetic Study

To evaluate the ZnO NPs sample’s ability to degrade CR dye, 20 mL of a 100 mg L−1 dye solution was added to a 250 mL beaker, which consists of two vessels (inner and outer) with water passing between them to control the temperature during the process of irradiating the solution, with 20 mg of ZnO NP powder. The mixture was then stirred for about 30 min in a dark environment, a crucial step to achieving equilibrium between adsorption and desorption, ensuring the accuracy of the results. A 60-watt lamp was used as a UV-VIS light source, positioned 10 cm from the solution’s base. At regular intervals of 15 min, a sample of the solution containing CR dye was extracted, and the absorbance was measured at a specific wavelength of 495 nm. The percentage of dye removal and the rate of removal constant were determined using Equations (1) and (2), respectively [45].
% D y e   d e c o l o r i z a t i o n = ( C o C t ) / ( C o ) × 100
k t = ln ( C o ) / C t
where Co is the initial dye concentration, Ct is the dye concentration after photodegradation, k is the reaction rate constant, and t is the time of reaction. All the experimental procedures were replicated three times under the examined parameters, including irradiation time, catalyst dose, pH level, and dye concentration.

3. Results and Discussion

The possible reaction mechanism of the as-synthesized ZnO NPs can be summarized via Equations (3)–(6) [46]. When the Zn ions (Zn2+) are formed in water, they are stabilized using the GA solution. When added, NaOH can react with the Zn2+ ions to form Zn(OH)2. Finally, the ZnO NPs were obtained when Z n O H 2 G u m   A r a b i c n ( s ) calcined at 400 °C for 1 h.
Z n S O 4 ( a q ) Z n       ( a q ) 2 + + S O 4       ( a q ) 2  
Z n       ( a q ) 2 + + G u m   A r a b i c   a q Z n G u m   A r a b i c n       ( a q ) 2 +  
Z n G u m   A r a b i c n       ( a q ) 2 + + 2 O H     ( a q ) Z n O H 2 G u m   A r a b i c n ( s )
Z n O H 2 G u m   A r a b i c n ( s )     Z n O   N P s  

3.1. Structural Characterization of the ZnO NPs

The X-ray diffraction patterns of the GA sample and ZnO NPs are represented in Figure 2a. The XRD pattern of the GA sample demonstrated that the powder lacked a distinct diffraction pattern, indicating its amorphous nature with a broad peak at 2θ ~19.5°. This is a significant observation in our research. In contrast, the diffraction peaks of ZnO NPs are located at 31.58°, 34.23°, 36.07°, 47.35°, 56.41°, 62.67°, 66.19°, 67.76°, 68.91°, 72.62°, and 76.86°, which are indexed to ZnO (JCPDS card number: 36-1451, hexagonal phase) [47]. Furthermore, the crystallite size (D, nm) of the ZnO NPs was estimated using the Debye–Scherrer Equation (7) [48].
D = 0.94 λ β c o s θ  
where 0.94 is Scherrer’s constant, λ is the X-ray wavelength, β is the full width at half maximum height (FWHM), and θ is the diffraction angle. The average crystallite size of the as-prepared ZnO NPs was 42.74 nm. The small crystallite size of ZnO NPs can speed up the photodegradation of CR dye by adding more active sites for adsorption.
FTIR analysis was used to identify the functional groups in the GA and ZnO NPs, as shown in Figure 2b. The band’s presence at 3415.9 cm−1 in the GA spectrum can be attributed to the O–H stretching vibration, a typical feature of the glycosidic ring [49]. The amino group’s distinctive absorption band within the range of 3300–3500 cm−1, though partially obscured by the broader absorption band of the O–H group, is a reliable finding. The C–H groups and specific sugar components in GA, such as galactose, arabinose, and rhamnose, are identifiable at 2926 cm−1 [50]. The spectrum peak at 1632.1 cm−1 is attributed to the vibrational stretching of the C=O bond. The typical band of C=C (stretching) was observed at around 1608.6 cm−1. The glucuronic acids exhibit distinct vibrational modes, including a band at 1426.6 cm−1 attributed to C=O symmetric stretching and another band at 1373.3 cm−1 associated with –OH bending. The peaks within 1300 to 900 cm−1 are attributed to the –C–O–C– and C–O stretching modes [51,52,53,54]. The presence of the bands at 416.9 and 617.2 cm−1 was ascribed to the stretching of the Zn–O bond [53,54]. FE-SEM analysis shows that the exterior surface of the GA (Figure 3a) appears to be an irregular and rocky-like structure with many dents. Figure 3b represents the FE-SEM image of the ZnO NPs. It demonstrates the configuration of the nanoparticle.

3.2. Catalytic Activity of ZnO NPs

The photocatalytic degradation of CR dye using ZnO NPs was detected at different contact times at λmax of 495 nm. The results of a typical run are represented in Figure 4a. The concentration of the CR dye decreased with increasing irradiation time, indicating that the CR dye was degraded photocatalytically. The control experiments were achieved and indicated that the dye is degraded only with the existence of photocatalysts upon irradiation.

3.3. Effect of Irradiation Time

UV/Vis spectrophotometric analysis was used to study the effect of irradiation time on CR degradation. The photocatalytic degradation of the CR dye was carried out by adding 20.0 mg of ZnO NPs to the solution and irradiating it at different times. The absorbance was then measured to find the CR dye’s percentage of photodegradation (%D.). Figure 4b illustrates the impact of varying irradiation times on the degradation process of the CR dye. The results demonstrated a positive correlation between irradiation time and the degradation percentage. Specifically, the degradation percentage exhibited an upward trend, reaching a peak of 97.8% after 90 min of irradiation. Conversely, the lowest degradation percentage of 80.4% was seen after 15 min of irradiation.

3.4. Effect of Variation of Catalyst Dose on Photodegradation of CR Dye

The photocatalytic degradation process was carried out with various amounts of catalyst to determine how much of the mass of the photocatalyst affected its photocatalytic activity. Different mass catalysts (0.02, 0.04, 0.06, 0.08, and 0.1 g) were used to study what happens when the mass of ZnO NPs changes during the photodegradation process of the CR dye. As shown in Figure 5a, the percentage of photodegradation increases as the catalyst is raised from 0.02 to 0.08 g. In addition, Figure 5b,c show the rate of photodegradation constant (K), which increased with increasing doses of catalyst from 0.02 to 0.08 g. This increase in catalyst mass leads to a corresponding rise in active sites, which could explain the relationship between the dye percentage and the catalyst’s weight [55]. When the weight of the substance exceeds 100 mg, the percentage of destroyed dye decreases. The solution’s light catalyst particles may be to blame for this drop’s light blocking and scattering [56]. The CR degradation efficiency on ZnO NPs is compared with other related materials, as listed in Table 2.

3.5. Effect of pH of Solution

The pH level is an essential factor in the photocatalytic degradation process [62]. It can be significantly affected by the size of particles, the charge on their surfaces, and where the boundaries of the valence and conduction bands are located [63]. The tested pH range was 5–9, as shown in Figure 6a, and the effect of pH on the rate constant (k) is listed in Figure 6b,c. The dye’s percentage of photodegradation led to an increase in free hydroxyl radical production, which raised the pH to 8. On the other hand, the dye’s photodegradation rate declined above a pH of 8, which may be due to the catalyst’s propensity for corrosion [64]. The rationale for this phenomenon is derived from the photocatalyst’s point of zero charge (PZC). Therefore, the surface charge was studied by assessing the PZC. The results revealed that the pHPZC is 8.4 (Figure S1), where above this pH value, the catalyst system will be negatively charged, and the result was confirmed from the results of different pH experiments. As shown in the effect of the pH findings, the photodegradation activity was reduced dramatically by increasing the pH value above pH 8, and this was attributed to the repulsion between the negatively charged catalyst surface at this point with the anionic CR dye, which is negative in nature [54]. Thus, the findings highlighted the crucial role of adsorption. Hydroxyl radicals are the main oxidizing species that break down materials through photocatalysis. The positively charged surface of ZnO NPs and the negatively charged CR dye interact electrostatically, which produces more hydroxyl radicals [65].

3.6. Effect of Dye Concentration

The present investigation examined the impact of varying concentrations of CR dye on the photocatalytic reaction. Different concentrations (20, 40, 60, 80, 100, and 120 mg L−1) were considered. The results are presented in Figure 7a,b. In addition, the effect of the dye concentration on the rate constant (K) is listed in Figure 7c,d. The degradation percentage and rate photodegradation constant are increased at lower dye concentrations, whereas they demonstrated a reduction at higher concentrations. The efficacy of the photocatalytic process is contingent upon the extent to which light can permeate the dye solution and then interact with the ZnO NPs, thereby enhancing the reduction of absorbed oxygen. At lower dye concentrations, the incident light can saturate the dye solution and absorb it from the photocatalyst. At elevated dye concentrations, a significant quantity of dye molecules will impede the passage of light photons to the photocatalyst. The rate of photoreaction is decreased by this phenomenon [66,67]. Conversely, when present in low quantities, it is shown that not all active sites are filled by dye molecules. Various free radicals contribute to the formation of distinct free radicals, which are responsible for the photodegradation of the dye. Consequently, the photoreaction rate significantly increases when the dye is in low quantities [68,69].

3.7. Photocatalytic Kinetic Study

Langmuir-Hinshelwood (L-H) kinetics are a common way to characterize the reaction between oxygen-containing molecules and surface-reducible reactants in photocatalysis. This is the most common kinetic model for photodegradation and heterogeneous catalytic processes. Based on certain assumptions of the Langmuir adsorption isotherm, it is utilized to determine the rate of a heterogeneous catalytic process. In addition, it is crucial to consider that the L-H model is affected by several elements, such as the characteristics of the catalyst, the light sources used, the parameters of mass transfer, and the operating circumstances. It can be expressed using Equation (8) as follows:
d C d t = k K C 1 + K C  
where C is the dye concentration (mg L−1), k is the rate constant (mg L−1 min−1), and K is the equilibrium adsorption constant (L mg−1).
The photocatalyst concentration remains unchanged during the photocatalytic reaction. Therefore, the reaction rate depends on the dye concentration. Accordingly, the pseudo-term can be used to determine the reaction rate. The surface of the photocatalyst is saturated if the dye concentration is very high; therefore, the reaction rate constant does not depend on the dye concentration (a zero-order expression). As shown in Equation (9):
d C d t = K  
In contrast, when the dye concentration is low, the rate of photodegradation depends on the change in the dye concentration, meaning the rate of photodegradation is proportional to the dye concentration, as expressed in Equation (10):
d C d t = K 1 C  
Figure 8 shows the plot of (Co-Ct), (ln C), and (1/C) versus time of irradiation (in min) of zero, first, and second-order reactions, respectively, based on the integrated equation for the CR dye at the optimum pH values of the solution and dose of the photocatalyst. In addition, Table 3 shows the values of the rate constant and regression coefficient for zero, first, and second-order reactions. Based on the findings, the reaction obeyed pseudo-first-order kinetics because it produced a straight line with a high regression coefficient (R2 = 0.998).

3.8. Photocatalytic Performance

To evaluate the recyclability and stability of the as-synthesized ZnO NP photocatalyst, tests were conducted by executing recycling reactions five times for the photodegradation of CR dye under a UV-VIS light source. As shown in Figure 9, no noticeable loss of the photocatalytic activity of the ZnO NPs was observed for the CR dye degradation reaction after repeated cycles. This indicates that ZnO has good recyclability and stability and could be a potent material for practical photocatalysis applications.

3.9. Photocatalytic Mechanism

The UV-VIS diffuse reflectance spectrum (DRS) (Figure S2a,b) was conducted to estimate the band gap (Eg) value of the ZnO NPs. It was calculated from Equation (11) to be 2.9 eV.
Eg = 1240.2/λ(nm)
In addition, from the high-resolution XPS spectra in Figure S2c,d, the valence band (VB) was calculated to be 2.42 eV, indicating enough oxidation potential to produce •OH from OH. The conduction band (CB) was also calculated from this relation, ECB = Eg − EVB, to be −0.48 eV, more than the reduction potential of O2/•O2 (−0.33 eV) [70]. Therefore, we can conclude that O2 and •OH play the main roles in the photodegradation of CR dye.
Figure 10 represents the CR dye photocatalytic degradation mechanism using a ZnO NP catalyst. Based on several studies, violet light can transfer electrons from the VB to the CB [71,72]. This creates holes (h+) and electrons (e) on the surface of the Zn NP catalyst, which is a vital part of breaking down the CR dye. The diazo groups in the CR dye are cleaved due to the generation of •OH radicals on the surface of the photocatalyst, which arises via the oxidation of the H2O molecule [73,74]. Furthermore, a photoinduced electron transforms the adsorbed O2 molecule, forming O•2− radicals via the reduction process. The O•2− radical then reacts with the H2O molecule, forming an H2O2 molecule. This molecule is further reduced by electrons, leading to the production of •OH radicals. Following a sequence of oxidation reactions, the CR dye converts into carbon dioxide (CO2), H2O, and non-toxic molecules.

4. Conclusions

In summary, the green synthesis of the zinc oxide nanoparticle (ZnO NP) catalyst was successfully demonstrated using Gum Arabic material as a natural stabilizing agent by calcination. Gum Arabic was used as an alternative to potentially harmful and toxic materials. It was employed for the photodegradation of Congo Red (CR) dye in an aqueous solution for the first time. The structural and morphological characterizations revealed that the green synthesis of the ZnO NPs via Gum Arabic improved its properties, which further helped enhance the dye adsorption performance. It possessed a small crystallite size of 42.7 nm that could speed up the photodegradation of CR dye by adding more active sites for adsorption. Furthermore, the findings indicated that the degree of CR dye degradation was affected by dye concentration, photocatalyst dose, and pH value. The optimal parameters were shown to be CR dye concentration = 20 mg L−1, amount of photocatalyst = 4 mg mL−1, and pH = 8 for a degradation percentage of 99.5% during a 30 min irradiation time. In addition, the CR dye photokinetics of the ZnO-NPs were examined systematically using various kinetic models. The results suggested that the pseudo-first-order model was deemed the most fitting, identified by maximum regression coefficient values. The as-synthesized ZnO NP catalyst exhibited excellent catalytic activity, and it is recommended as an efficient photocatalyst to remove CR dye from wastewater.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/w16152202/s1, Figure S1: Point of Zero charge plot ΔpH vs pHi for the synthesized ZnO NPs catalyst at different pH values; Figure S2: (a) DRS of ZnO catalyst. (b) valence band of ZnO. High-resolution XPS spectrum of the (c) Zn 2p and (d) O 1s for ZnO catalyst.

Author Contributions

Conceptualization, H.S.A.; methodology, H.S.A.; formal analysis, A.R.O., A.A.M. and A.H.M.; investigation, H.S.A. and M.R.T.; data curation, A.R.O., A.A.M. and A.H.M.; writing—original draft preparation, A.R.O.; writing—review and editing, M.R.T.; visualization, M.R.T.; supervision, H.S.A. and M.R.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 study are available in the article.

Acknowledgments

The authors gratefully acknowledge the Environmental Research and Studies Center at the University of Babylon for supporting this work.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Pratap, B.; Kumar, S.; Nand, S.; Azad, I.; Bharagava, R.N.; Ferreira, L.F.; Dutta, V. Wastewater generation and treatment by various eco-friendly technologies: Possible health hazards and further reuse for environmental safety. Chemosphere 2023, 313, 137547. [Google Scholar] [CrossRef] [PubMed]
  2. Jain, M.; Khan, S.A.; Sharma, K.; Jadhao, P.R.; Pant, K.K.; Ziora, Z.M.; Blaskovich, M.A.T. Current perspective of innovative strategies for bioremediation of organic pollutants from wastewater. Bioresour. Technol. 2022, 344, 126305. [Google Scholar] [CrossRef] [PubMed]
  3. Hanafi, M.F.; Sapawe, N. A review on the water problem associate with organic pollutants derived from phenol, methyl orange, and remazol brilliant blue dyes. Mater. Today Proc. 2020, 31, A141–A150. [Google Scholar] [CrossRef]
  4. Boyles, C.; Sobeck, S.J.S. Photostability of organic red food dyes. Food Chem. 2020, 315, 126249. [Google Scholar] [CrossRef] [PubMed]
  5. Yamjala, k.; Nainar, M.S.; Ramisetti, N.R. Methods for the analysis of azo dyes employed in food industry—A review. Food Chem. 2016, 192, 813–824. [Google Scholar] [CrossRef] [PubMed]
  6. Markandeya; Mohan, D.; Shukla, S.P. Hazardous consequences of textile mill effluents on soil and their remediation approaches. Clean. Eng. Technol. 2022, 7, 100434. [Google Scholar] [CrossRef]
  7. Moradihamedani, P. Recent advances in dye removal from wastewater by membrane technology: A review. Polym. Bull. 2022, 79, 2603–2631. [Google Scholar] [CrossRef]
  8. Khan, M.D.; Singh, A.; Khan, M.Z.; Tabraiz, S.; Sheikh, J. Current perspectives, recent advancements, and efficiencies of various dye-containing wastewater treatment technologies. J. Water Process Eng. 2023, 53, 103579. [Google Scholar] [CrossRef]
  9. Kasbaji, M.; Mennani, M.; Oubenali, M.; Benhamou, A.A.; Boussetta, A.; Ablouh, E.H.; Mbarki, M.; Grimi, N.; El Achaby, M.; Moubarik, A. Bio-based functionalized adsorptive polymers for sustainable water decontamination: A systematic review of challenges and real-world implementation. Environ. Pollut. 2023, 335, 122349. [Google Scholar] [CrossRef]
  10. Mane, P.V.; Rego, R.M.; Yap, P.L.; Losic, D.; Kurkuri, M.D. Unveiling cutting-edge advances in high surface area porous materials for the efficient removal of toxic metal ions from water. Prog. Mater. Sci. 2024, 146, 101314. [Google Scholar] [CrossRef]
  11. Li, W.; Mu, B.; Yang, Y. Feasibility of industrial-scale treatment of dye wastewater via bio-adsorption technology. Bioresour. Technol. 2019, 277, 157–170. [Google Scholar] [CrossRef] [PubMed]
  12. Mcyotto, F.; Wei, Q.; Macharia, D.K.; Huang, M.; Shen, C.; Chow, C.W.K. Effect of dye structure on color removal efficiency by coagulation. Chem. Eng. J. 2021, 405, 126674. [Google Scholar] [CrossRef]
  13. Slama, H.B.; Bouket, A.C.; Pourhassan, Z.; Alenezi, F.N.; Silini, A.; Cherif-Silini, H.; Oszako, T.; Luptakova, L.; Golińska, P.; Belbahri, L. Diversity of synthetic dyes from textile industries, discharge impacts and treatment methods. Appl. Sci. 2021, 11, 6255. [Google Scholar] [CrossRef]
  14. Bal, G.; Thakur, A. Distinct approaches of removal of dyes from wastewater: A review. Mater. Today Proc. 2021, 50, 1575–1579. [Google Scholar] [CrossRef]
  15. Thalji, M.R. Nanotechnologies for removal of pharmaceuticals from wastewater. Med. Pharm. Sci. 2021, 1, 25–28. [Google Scholar]
  16. Mohammed, A.M.; Thalji, M.R.; Yasin, S.A.; Shim, J.J.; Chong, K.F.; Guda, A.A.; Ali, G.A.M. Recent advances in electrospun fibrous membranes for effective chromium (VI) removal from water. J. Mol. Liq. 2023, 383, 122110. [Google Scholar] [CrossRef]
  17. Quynh, H.G.; Thanh, H.V.; Phuong, N.T.T.; Duy, N.P.T.; Hung, L.H.; Dung, N.V.; Duong, N.T.H.; Long, N.Q. Rapid removal of methylene blue by a heterogeneous photo-Fenton process using economical and simple-synthesized magnetite–zeolite composite, Environ. Technol. Innov. 2023, 31, 103155. [Google Scholar] [CrossRef]
  18. Rana, S.; Kumar, A.; Dhiman, P.; Mola, G.T.; Sharma, G.; Lai, C.W. Recent advances in photocatalytic removal of sulfonamide pollutants from waste water by semiconductor heterojunctions: A review. Mater. Today Chem. 2023, 30, 101603. [Google Scholar] [CrossRef]
  19. Rafiq, A.; Ikram, M.; Ali, S.; Niaz, F.; Khan, M.; Khan, Q.; Maqbool, M. Photocatalytic degradation of dyes using semiconductor photocatalysts to clean industrial water pollution. J. Ind. Eng. Chem. 2021, 97, 111–128. [Google Scholar] [CrossRef]
  20. Ambigadevi, J.; Senthil Kumar, P.S.; Vo, D.V.N.; Haran, S.H.; Raghavan, T.N.S. Recent developments in photocatalytic remediation of textile effluent using semiconductor based nanostructured catalyst: A review. J. Environ. Chem. Eng. 2021, 9, 104881. [Google Scholar] [CrossRef]
  21. Kumari, H.; Sonia; Suman; Ranga, R.; Chahal, S.; Devi, S.; Sharma, S.; Kumar, S.; Kumar, P.; Kumar, S.; et al. A Review on photocatalysis used for wastewater treatment: Dye degradation. Water Air Soil Pollut. 2023, 234, 349. [Google Scholar] [CrossRef]
  22. Dihom, H.R.; Al-Shaibani, M.M.; Mohamed, R.M.S.; Al-Gheethi, A.A.; Sharma, A.; Bin Khamidun, M.H. Photocatalytic degradation of disperse azo dyes in textile wastewater using green zinc oxide nanoparticles synthesized in plant extract: A critical review. J. Water Process Eng. 2022, 47, 102705. [Google Scholar] [CrossRef]
  23. Rambabu, K.; Bharath, G.; Banat, F.; Show, P.L. Green synthesis of zinc oxide nanoparticles using Phoenix dactylifera waste as bioreductant for effective dye degradation and antibacterial performance in wastewater treatment. J. Hazard. Mater. 2021, 402, 123560. [Google Scholar] [CrossRef]
  24. Barrios-Navarro, F.A.; Vilchis-Nestor, A.R.; Luque, P.A. Photocatalytic degradation of organic dyes in water using semiconductor ZnO nanoparticles synthesized using Crataegus mexicana extract. Mater. Chem. Phys. 2024, 318, 129302. [Google Scholar] [CrossRef]
  25. Shashanka, R.; Esgin, H.; Yilmaz, V.M.; Caglar, Y. Fabrication and characterization of green synthesized ZnO nanoparticle based dye-sensitized solar cells. J. Sci. Adv. Mater. Devices 2020, 5, 185–191. [Google Scholar] [CrossRef]
  26. Ndolomingo, M.J.; Bingwa, N.; Meijboom, R. Review of supported metal nanoparticles: Synthesis methodologies, advantages and application as catalysts. J. Mater. Sci. 2020, 55, 6195–6241. [Google Scholar] [CrossRef]
  27. Xu, L.; Liang, H.W.; Yang, Y.; Yu, S.H. Stability and reactivity: Positive and negative aspects for nanoparticle processing. Chem. Rev. 2018, 118, 3209–3250. [Google Scholar] [CrossRef]
  28. Ashraf, M.A.; Peng, W.; Zare, Y.; Rhee, K.Y. Effects of size and aggregation/agglomeration of nanoparticles on the interfacial/interphase properties and tensile strength of polymer nanocomposites. Nanoscale Res. Lett. 2018, 13, 214. [Google Scholar] [CrossRef]
  29. Khrenov, V.; Klapper, M.; Koch, M.; Müllen, K. Surface functionalized ZnO particles designed for the use in transparent nanocomposites. Macromol. Chem. Phys. 2005, 206, 95–101. [Google Scholar] [CrossRef]
  30. Yashni, G.; Al-Gheethi, A.; Mohamed, R.; Hossain, S.; Kamila, A.F.; Shanmugan, V.A. Photocatalysis of xenobiotic organic compounds in greywater using zinc oxide nanoparticles: A critical review. Water Environ. J. 2021, 35, 190–217. [Google Scholar] [CrossRef]
  31. Caillol, S. A Blooming season for natural polymers and biopolymers. Molecules 2023, 28, 3207. [Google Scholar] [CrossRef] [PubMed]
  32. Iravani, S. Plant gums for sustainable and eco-friendly synthesis of nanoparticles: Recent advances. Inorg. Nano-Metal Chem. 2020, 50, 469–488. [Google Scholar] [CrossRef]
  33. Samrot, A.V.; Angalene, J.L.A.; Roshini, S.M.; Raji, P.; Stefi, S.M.; Preethi, R.; Selvarani, A.J.; Madankumar, A. Bioactivity and heavy metal removal using plant gum mediated green synthesized silver nanoparticles. J. Clust. Sci. 2019, 30, 1599–1610. [Google Scholar] [CrossRef]
  34. Amiri, M.S.; Mohammadzadeh, V.; Yazdi, M.E.T.; Barani, M.; Rahdar, A.; Kyzas, G.Z. Plant-based gums and mucilages applications in pharmacology and nanomedicine: A review. Molecules 2021, 26, 1770. [Google Scholar] [CrossRef] [PubMed]
  35. Souza, J.M.; de Araujo, A.R.; de Carvalho, A.M.; Amorim, A.D.; Daboit, T.C.; de Almeida, J.R.; da Silva, D.A.; Eaton, P. Sustainably produced cashew gum-capped zinc oxide nanoparticles show antifungal activity against Candida parapsilosis. J. Clean. Prod. 2020, 247, 119085. [Google Scholar] [CrossRef]
  36. Ashwini, J.; Aswathy, T.R.; Rahul, A.B.; Thara, G.M.; Nair, A.S. Synthesis and characterization of zinc oxide nanoparticles using Acacia caesia bark extract and its photocatalytic and antimicrobial activities. Catalysts 2021, 11, 1507. [Google Scholar] [CrossRef]
  37. La, D.D.; Nguyen-Tri, P.; Le, K.H.; Nguyen, P.T.M.; Nguyen, M.D.B.; Vo, A.T.K.; Nguyen, M.T.H.; Chang, S.W.; Tran, L.D.; Chung, W.J.; et al. Effects of antibacterial ZnO nanoparticles on the performance of a chitosan/gum arabic edible coating for post-harvest banana preservation. Prog. Org. Coat. 2021, 151, 106057. [Google Scholar] [CrossRef]
  38. Mittal, H.; Morajkar, P.P.; Al Alili, A.; Alhassan, S.M. In-situ synthesis of ZnO nanoparticles using gum arabic based hydrogels as a self-template for effective malachite green dye adsorption. J. Polym. Environ. 2020, 28, 1637–1653. [Google Scholar] [CrossRef]
  39. Punnakkal, V.S.; Anila, E.I. Polypyrrole/silver/graphene ternary nanocomposite synthesis and study on photocatalytic property in degrading Congo red dye under visible light. Surf. Interfaces 2023, 42, 103342. [Google Scholar] [CrossRef]
  40. Mohebali, S.; Bastani, D.; Shayesteh, H. Equilibrium, kinetic and thermodynamic studies of a low-cost biosorbent for the removal of Congo red dye: Acid and CTAB-acid modified celery (Apium graveolens). J. Mol. Struct. 2019, 1176, 181–193. [Google Scholar] [CrossRef]
  41. Afkhami, A.; Moosavi, R. Adsorptive removal of Congo red, a carcinogenic textile dye, from aqueous solutions by maghemite nanoparticles. J. Hazard. Mater. 2010, 174, 398–403. [Google Scholar] [CrossRef] [PubMed]
  42. Oladoye, P.; Bamigboye, M.; Ogunbiyi, O.; Akano, M. Toxicity and decontamination strategies of Congo red dye. Groundw. Sustain. Dev. 2022, 19, 100844. [Google Scholar] [CrossRef]
  43. Vairavel, P.; Rampal, N.; Jeppu, G. Adsorption of toxic Congo red dye from aqueous solution using untreated coffee husks: Kinetics, equilibrium, thermodynamics and desorption study. Int. J. Environ. Anal. Chem. 2023, 103, 2789–2808. [Google Scholar] [CrossRef]
  44. Akintelu, S.A.; Folorunso, A.S. A Review on green synthesis of zinc oxide nanoparticles using plant extracts and its biomedical applications. Bionanoscience 2020, 10, 848–863. [Google Scholar] [CrossRef]
  45. Chauhan, P.S.; Kumar, K.; Singh, K.; Bhattacharya, S. Fast decolorization of rhodamine-B dye using novel V2O5-rGO photocatalyst under solar irradiation. Synth. Met. 2022, 283, 116981. [Google Scholar] [CrossRef]
  46. Pauzi, N.; Zain, N.M.; Yusof, N.A.A. Gum arabic as natural stabilizing agent in green synthesis of ZnO nanofluids for antibacterial application. J. Environ. Chem. Eng. 2020, 8, 103331. [Google Scholar] [CrossRef]
  47. Singh, A.; Wan, F.; Yadav, K.; Salvi, A.; Thakur, P.; Thakur, A. Synergistic effect of ZnO nanoparticles with Cu2+ doping on antibacterial and photocatalytic activity. Inorg. Chem. Commun. 2023, 157, 111425. [Google Scholar] [CrossRef]
  48. Thalji, M.R.; Ali, G.A.M.; Shim, J.J.; Chong, K.F. Cobalt-doped tungsten suboxides for supercapacitor applications. Chem. Eng. J. 2023, 473, 145341. [Google Scholar] [CrossRef]
  49. Ibekwe, C.A.; Oyatogun, G.M.; Esan, T.A.; Oluwasegun, K.M. Synthesis and characterization of chitosan/gum arabic nanoparticles for bone regeneration. Am. J. Mater. Sci. Eng. 2017, 5, 28–36. [Google Scholar] [CrossRef]
  50. 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, 12–16. [Google Scholar] [CrossRef]
  51. Daoub, R.M.A.; Elmubarak, A.H.; Misran, M.; Hassan, E.A.; Osman, M.E. Characterization and functional properties of some natural Acacia gums. J. Saudi Soc. Agric. Sci. 2018, 17, 241–249. [Google Scholar] [CrossRef]
  52. Ai, C.; Zhao, C.; Xiang, C.; Zheng, Y.; Zhong, S.; Teng, H.; Chen, L. Gum arabic as a sole wall material for constructing nanoparticle to enhance the stability and bioavailability of curcumin. Food Chem. X 2023, 18, 100724. [Google Scholar] [CrossRef] [PubMed]
  53. Araujo, F.P.; Trigueiro, P.; Honório, L.M.C.; Oliveira, D.M.; Almeida, L.C.; Garcia, R.P.; Lobo, A.O.; Cantanhêde, W.; Silva-Filho, E.C.; Osajima, J.A. Eco-friendly synthesis and photocatalytic application of flowers-like ZnO structures using Arabic and Karaya Gums. Int. J. Biol. Macromol. 2020, 165, 2813–2822. [Google Scholar] [CrossRef] [PubMed]
  54. Petrović, Ž.; Ristić, M.; Musić, S.; Fabián, M. The effect of gum Arabic on the nano/microstructure and optical properties of precipitated ZnO. Croat. Chem. Acta 2017, 90, 135–143. [Google Scholar] [CrossRef]
  55. Oda, A.M. Degradation of Congo red solution by zinc oxide/silver composite preheated at different temperatures. J. Thermodyn. Catal. 2015, 05, 10–13. [Google Scholar] [CrossRef]
  56. Gaur, J.; Vikrant, K.; Kim, K.H.; Kumar, S.; Pal, M.; Badru, R.; Masand, S.; Momoh, J. Photocatalytic degradation of Congo red dye using zinc oxide nanoparticles prepared using Carica papaya leaf extract. Mater. Today Sustain. 2023, 22, 100339. [Google Scholar] [CrossRef]
  57. Hairom, N.H.H.; Mohammad, A.W.; Kadhum, A.A.H. Effect of various zinc oxide nanoparticles in membrane photocatalytic reactor for Congo red dye treatment. Sep. Purif. Technol. 2014, 137, 74–81. [Google Scholar] [CrossRef]
  58. Rupa, E.J.; Kaliraj, L.; Abid, S.; Yang, D.C.; Jung, S.K. Synthesis of a zinc oxide nanoflower photocatalyst from sea buckthorn fruit for degradation of industrial dyes in wastewater treatment. Nanomaterials 2019, 9, 1692. [Google Scholar] [CrossRef]
  59. Chakraborty, S.; Farida, J.J.; Simon, R.; Kasthuri, S.; Mary, N.L. Averrhoe carrambola fruit extract assisted green synthesis of ZnO nanoparticles for the photodegradation of Congo red dye. Surf. Interfaces 2020, 19, 100488. [Google Scholar] [CrossRef]
  60. Vidya, C.; Manjunatha, C.; Chandraprabha, M.N.; Rajshekar, M.; Raj, A. Hazard free green synthesis of ZnO nano-photo-catalyst using Artocarpus Heterophyllus leaf extract for the degradation of Congo red dye in water treatment applications. J. Environ. Chem. Eng. 2017, 5, 3172–3180. [Google Scholar] [CrossRef]
  61. Yashni, G.; Al-Gheethi, A.; Mohamed, R.M.S.; Dai-Viet, N.V.; Al-Kahtani, A.A.; Al-Sahari, M.; Hazhar, N.J.; Noman, E.; Alkhadher, S. Bio-inspired ZnO NPs synthesized from Citrus sinensis peels extract for Congo red removal from textile wastewater via photocatalysis: Optimization, mechanisms, techno-economic analysis. Chemosphere 2021, 281, 130661. [Google Scholar] [CrossRef] [PubMed]
  62. Abdullah, A.H.; Yasin, S.A.; Abdullah, S.M.; Thalji, M.R.; Aziz, F.; Assiri, M.A.; Chong, K.F.; Ali, G.A.M.; Bakr, Z.H. Removal of Cr(VI) using thiol-modified cellulose nanostructure for water sustainability: Detailed adsorption study. Biomass Convers. Biorefinery 2024. [Google Scholar] [CrossRef]
  63. Essa, W.K.; Yasin, S.A.; Abdullah, A.H.; Thalji, M.R.; Saeed, I.A.; Assiri, M.A.; Chong, K.F.; Ali, G.A.M. Taguchi L25 (54) approach for methylene blue removal by polyethylene terephthalate nanofiber-multi-walled carbon nanotube composite. Water 2022, 14, 1242. [Google Scholar] [CrossRef]
  64. Kumar, S.G.; Rao, K.S.R.K. Zinc oxide based photocatalysis: Tailoring surface-bulk structure and related interfacial charge carrier dynamics for better environmental applications. RSC Adv. 2015, 5, 3306–3351. [Google Scholar] [CrossRef]
  65. Zyoud, A.H.; Zubi, A.; Zyoud, S.H.; Hilal, M.H.; Zyoud, S.; Qamhieh, N.; Hajamohideen, A.R.; Hilal, H.S. Kaolin-supported ZnO nanoparticle catalysts in self-sensitized tetracycline photodegradation: Zero-point charge and pH effects. Appl. Clay Sci. 2019, 182, 105294. [Google Scholar] [CrossRef]
  66. Zhang, Z.; Li, Y.; Du, Q.; Li, Q. Adsorption of Congo Red from Aqueous Solutions by Porous Soybean Curd Xerogels. Pol. J. Chem. Technol. 2018, 20, 95–102. [Google Scholar] [CrossRef]
  67. Anand, G.T.; Renuka, D.; Ramesh, R.; Anandaraj, L.; Sundaram, S.J.; Ramalingam, G.; Magdalane, C.M.; Bashir, A.K.H.; Maaza, M.; Kaviyarasu, K. Green synthesis of ZnO nanoparticle using Prunus dulcis (Almond Gum) for antimicrobial and supercapacitor applications. Surf. Interfaces 2019, 17, 100376. [Google Scholar] [CrossRef]
  68. Birniwa, A.H.; Mohammad, R.E.A.; Ali, M.; Rehman, M.F.; Abdullahi, S.S.; Eldin, S.M.; Mamman, S.; Sadiq, A.C.; Jagaba, A.H. Synthesis of gum arabic magnetic nanoparticles for adsorptive removal of ciprofloxacin: Equilibrium, kinetic, thermodynamics studies, and optimization by response surface methodology. Separations 2022, 9, 322. [Google Scholar] [CrossRef]
  69. Harja, M.; Buema, G.; Bucur, D. Recent advances in removal of Congo red dye by adsorption using an industrial waste. Sci. Rep. 2022, 12, 6087. [Google Scholar] [CrossRef]
  70. Mady, A.H.; Baynosa, M.L.; Tuma, D.; Shim, J.-J. Facile microwave-assisted green synthesis of Ag-ZnFe2O4@rGO nanocomposites for efficient removal of organic dyes under UV- and visible-light irradiation. Appl. Catal. B Environ. 2017, 203, 416–427. [Google Scholar] [CrossRef]
  71. Sanakousar, M.F.; Vidyasagar, C.C.; Jiménez-Pérez, V.M.; Jayanna, B.K.; Mounesh; Shridhar, A.H.; Prakash, K. Efficient photocatalytic degradation of crystal violet dye and electrochemical performance of modified MWCNTs/Cd-ZnO nanoparticles with quantum chemical calculations. J. Hazard. Mater. Adv. 2021, 2, 100004. [Google Scholar] [CrossRef]
  72. Puneetha, J.; Kottam, N.; Rathna, A. Investigation of photocatalytic degradation of crystal violet and its correlation with bandgap in ZnO and ZnO/GO nanohybrid. Inorg. Chem. Commun. 2021, 125, 108460. [Google Scholar] [CrossRef]
  73. Nasir, J.A.; Gul, S.; Khan, A.; Shah, Z.H.; Ahmad, A.; Zulfiqar; Khan, R.; Liu, Z.; Chen, W.; Lin, D.-J.; et al. Efficient solar light driven photocatalytic degradation of Congo red dye on CdS nanostructures derived from single source precursor. J. Nanosci. Nanotechnol. 2018, 18, 7405–7413. [Google Scholar] [CrossRef]
  74. Chankhanittha, T.; Watcharakitti, J.; Nanan, S. PVP-assisted synthesis of rod-like ZnO photocatalyst for photodegradation of reactive red (RR141) and Congo red (CR) azo dyes. J. Mater. Sci. Mater. Electron. 2019, 30, 17804–17819. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration for synthesizing ZnO NPs using Gum Arabic.
Figure 1. Schematic illustration for synthesizing ZnO NPs using Gum Arabic.
Water 16 02202 g001
Figure 2. (a) XRD patterns and (b) FT-IR spectra of Gum Arabic (GA) and ZnO NPs.
Figure 2. (a) XRD patterns and (b) FT-IR spectra of Gum Arabic (GA) and ZnO NPs.
Water 16 02202 g002
Figure 3. FE-SEM images of the (a) Gum Arabic (GA) and (b) ZnO NP materials.
Figure 3. FE-SEM images of the (a) Gum Arabic (GA) and (b) ZnO NP materials.
Water 16 02202 g003
Figure 4. (a) Photodegradability of CR dye using different operating conditions. (b) Effect of irradiation time (min) on the photodegradation of CR dye.
Figure 4. (a) Photodegradability of CR dye using different operating conditions. (b) Effect of irradiation time (min) on the photodegradation of CR dye.
Water 16 02202 g004
Figure 5. (a) Effect of different catalyst doses on photodegradation of 30 mg L−1 of CR dye at room temperature and pH = 8. (b,c) Effect of different catalyst doses on rate of photodegradation constant (K) of 30 mg L−1 of CR dye at room temperature, and pH = 8. Error bars represent the standard deviations of duplicate runs.
Figure 5. (a) Effect of different catalyst doses on photodegradation of 30 mg L−1 of CR dye at room temperature and pH = 8. (b,c) Effect of different catalyst doses on rate of photodegradation constant (K) of 30 mg L−1 of CR dye at room temperature, and pH = 8. Error bars represent the standard deviations of duplicate runs.
Water 16 02202 g005
Figure 6. (a) Effect of different pH values on photodegradation percentage of 100 mg L−1 of CR dye at room temperature, photocatalyst dose (0.02 g/20 mL), and contact time 90 min. (b,c) Effect of different pH on rate of photodegradation constant (K) of 100 mg L−1 of CR dye at room temperature, and photocatalyst dose (0.02 g/20 mL). Error bars represent the standard deviations of duplicate runs.
Figure 6. (a) Effect of different pH values on photodegradation percentage of 100 mg L−1 of CR dye at room temperature, photocatalyst dose (0.02 g/20 mL), and contact time 90 min. (b,c) Effect of different pH on rate of photodegradation constant (K) of 100 mg L−1 of CR dye at room temperature, and photocatalyst dose (0.02 g/20 mL). Error bars represent the standard deviations of duplicate runs.
Water 16 02202 g006
Figure 7. (a) Effect of dye concentration on photodegradation percentage at different concentrations. (b) Percentage of decolorization vs. dye concentration. (c,d) Rate constant of CR dye at room temperature and pH = 8 and contact time = 90 min. Error bars represent the standard deviations of duplicate runs.
Figure 7. (a) Effect of dye concentration on photodegradation percentage at different concentrations. (b) Percentage of decolorization vs. dye concentration. (c,d) Rate constant of CR dye at room temperature and pH = 8 and contact time = 90 min. Error bars represent the standard deviations of duplicate runs.
Water 16 02202 g007
Figure 8. Fitting procedures of the (a) zero-, (b) first-, and (c) second-order reactions at the optimum conditions (pH = 8, catalyst dose = 0.02 g, and dye concentration = 100 mg L−1). Error bars represent the standard deviations of duplicate runs.
Figure 8. Fitting procedures of the (a) zero-, (b) first-, and (c) second-order reactions at the optimum conditions (pH = 8, catalyst dose = 0.02 g, and dye concentration = 100 mg L−1). Error bars represent the standard deviations of duplicate runs.
Water 16 02202 g008
Figure 9. The stability of ZnO NPs for photodegradation of CR dye.
Figure 9. The stability of ZnO NPs for photodegradation of CR dye.
Water 16 02202 g009
Figure 10. The photocatalytic CR dye degradation mechanism by ZnO NP catalyst.
Figure 10. The photocatalytic CR dye degradation mechanism by ZnO NP catalyst.
Water 16 02202 g010
Table 1. Harmful and toxic effects of Congo Red and the affected targets.
Table 1. Harmful and toxic effects of Congo Red and the affected targets.
Harmful EffectsTargets
Undesirable mutagenicity Living organisms
Carcinogenic textile dyeHumans and animals
PhytotoxicityPlants
Allergic dermatitis, skin, eye, and gastrointestinal irritationHumans
Difficulty in breathing, chest pain, and severe headacheHumans
Table 2. The photodegradation percentage comparison of CR dye using as-synthesized ZnO NPs with other related materials.
Table 2. The photodegradation percentage comparison of CR dye using as-synthesized ZnO NPs with other related materials.
ZnO
Photocatalyst
Congo Red
Concentration
(mg L−1)
Degradation
Efficiency (%)
Time
(min)
Ref.
0.3 g L−1 of ZnO-PVP-St2065---[57]
SBT-ZnO/NF15 9980[58]
150 mg of ZnO100 93150[59]
ZnO20 <901 h[60]
0.17 g/100 mL of ZnO5 96---[61]
4 mg mL−1 of ZnO NPs20 99.530This work
Table 3. Rate constant and regression coefficient values for CR dye photodegradation obtained by plotting relationships of zero-, first-, and second-order reaction at optimum values.
Table 3. Rate constant and regression coefficient values for CR dye photodegradation obtained by plotting relationships of zero-, first-, and second-order reaction at optimum values.
Kinetic ModelRegression Coefficient (R2)Rate Constant
Zero-Order Reaction0.7230.9753 K M min−1
1st-Order Reaction0.9980.067 K (min−1)
2nd-Order Reaction0.6270.0344 K M−1 min−1
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

Alhasan, H.S.; Omran, A.R.; Al Mahmud, A.; Mady, A.H.; Thalji, M.R. Toxic Congo Red Dye Photodegradation Employing Green Synthesis of Zinc Oxide Nanoparticles Using Gum Arabic. Water 2024, 16, 2202. https://doi.org/10.3390/w16152202

AMA Style

Alhasan HS, Omran AR, Al Mahmud A, Mady AH, Thalji MR. Toxic Congo Red Dye Photodegradation Employing Green Synthesis of Zinc Oxide Nanoparticles Using Gum Arabic. Water. 2024; 16(15):2202. https://doi.org/10.3390/w16152202

Chicago/Turabian Style

Alhasan, Huda S., Alaa R. Omran, Abdullah Al Mahmud, Amr Hussein Mady, and Mohammad R. Thalji. 2024. "Toxic Congo Red Dye Photodegradation Employing Green Synthesis of Zinc Oxide Nanoparticles Using Gum Arabic" Water 16, no. 15: 2202. https://doi.org/10.3390/w16152202

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