Next Article in Journal
Assessment of the Suitability of Coke Material for Proppants in the Hydraulic Fracturing of Coals
Previous Article in Journal
Atomic Layer Deposition of HfO2 Films Using TDMAH and Water or Ammonia Water
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Modeling and Optimizing the Crystal Violet Dye Adsorption on Kaolinite Mixed with Cellulose Waste Red Bean Peels: Insights into the Kinetic, Isothermal, Thermodynamic, and Mechanistic Study

1
Pollution & Waste Treatment Laboratory (PWTL), University of Ouargla, P.O. Box 511, Ouargla 30000, Algeria
2
Department of Chemistry, Faculty of Exact Sciences, University of El-Oued, P.O. Box 789, El-Oued 39000, Algeria
3
Chemical Engineering Department, College of Engineering, University of Ha’il, P.O. Box 2440, Ha’il 81441, Saudi Arabia
4
Department of Chemical Engineering, Faculty of Engineering, Al Neelain University, Khartoum 12702, Sudan
5
Chemical Engineering Department, Faculty of Engineering, University of Blida, Blida 09000, Algeria
6
Department of Chemical and Petroleum Engineering, University Elmergib, Al-Khums P.O. Box 40161, Libya
7
Chemical Engineering Process Department, National School of Engineers Gabes, University of Gabes, Gabes 6029, Tunisia
*
Author to whom correspondence should be addressed.
Materials 2023, 16(11), 4082; https://doi.org/10.3390/ma16114082
Submission received: 11 April 2023 / Revised: 19 May 2023 / Accepted: 23 May 2023 / Published: 30 May 2023

Abstract

:
In this study, a new eco-friendly kaolinite–cellulose (Kaol/Cel) composite was prepared from waste red bean peels (Phaseolus vulgaris) as a source of cellulose to serve as a promising and effective adsorbent for the removal of crystal violet (CV) dye from aqueous solutions. Its characteristics were investigated through the use of X-ray diffraction, Fourier-transform infrared spectroscopy, scanning electron microscopy, energy-dispersive X-ray spectroscopy, and zero-point of charge (pHpzc). The Box–Behnken design was used to improve CV adsorption on the composite by testing its primary affecting factors: loading Cel into the composite matrix of Kaol (A: 0–50%), adsorbent dosage (B: 0.02–0.05 g), pH (C: 4–10), temperature (D: 30–60 °C), and duration (E: 5–60 min). The significant interactions with the greatest CV elimination efficiency (99.86%) are as follows: BC (adsorbent dose vs. pH) and BD (adsorbent dose vs. temperature) at optimum parameters (A: 25%, B: 0.05 g, C: 10, D: 45 °C, and E: 17.5 min) for which the CV’s best adsorption capacity (294.12 mg/g) was recorded. The Freundlich and pseudo-second-order kinetic models were the best isotherm and kinetic models fitting our results. Furthermore, the study investigated the mechanisms responsible for eliminating CV by utilizing Kaol/Cel–25. It detected multiple types of associations, including electrostatic, n-π, dipole–dipole, hydrogen bonding interactions, and Yoshida hydrogen bonding. These findings suggest that Kaol/Cel could be a promising starting material for developing a highly efficient adsorbent that can remove cationic dyes from aqueous environments.

1. Introduction

Today, the planet faces serious pollution issues, a global concern for the natural ecosystem. Much dye-containing wastewater relating to human activities is discharged without treatment, seriously affecting the environment and drinking water quality [1]. As contaminants in North African nations, including Algeria, dyes are dumped into the water without treatment [2]. Dyes can be divided into three categories according to their core structure: cationic dyes (basic), anionic dyes (acidic and reactive), and nonionic dyes (disperse) [3]. Organic cationic dyes, such as crystal violet (CV), are more toxic to humans and other living organisms than anionic ones [4]. Reducing or eliminating the dye content of wastewater before discharging it into the environment is a way to handle this significant ecological problem [5]. Several techniques for removing anionic dyes from wastewater have been presented, including ion exchange, membrane filtration, and irradiation, as well as chemical approaches such as oxidation, coagulation, zonation, photochemical and electrochemical destruction. Biological methods such as aerobic and anaerobic microbial degradation and photochemical treatments have also been proposed [6,7,8]. However, the adsorption method is still the method of choice for dye removal because it has several advantages over other methods other forms do not have. These advantages include a simple design, a broad pH, efficacy even at low pollutant concentrations, selectivity, the ability to be regenerated, and the design is easy to operate in moderate conditions [9,10]. Many researchers have focused on adsorption technology employing cheap adsorbents [11].
Clays and their composites have recently emerged as effective adsorbents for dye-containing water. This is due to the clays’ high microporosity, high stability, large area of a specific surface, excellent swelling ability, high cation-exchange capacity (CEC), and environmental safety [12,13]. It has been discovered that clays and the composites made from them are excellent in removing dye from water while being quite inexpensive. Kaolinite is a well-known example of a layered substance or clay, which may be found at [14]. Kaolinite is an aluminum silicate clay with a 1:1 dioctahedral layered structure consisting of a single SiO4 tetrahedral layer and an Al(OH)6 octahedral sheet in each layer. These layers are interconnected by an O-H-O bond. Kaolinite exhibits adsorption properties and is therefore used for eliminating various toxins from wastewater [15]. However, compared to illite, chitosan, activated carbon, or zeolites, raw kaolinite has a lower adsorption capacity and CEC (3–15 mEq/100 g) [16]. Adsorbing pollutants on natural minerals decreases capacity after a few cycles [9,16]. Several techniques have been used to modify clay, such as thermal, acid/base activation, intercalation, and metal oxide coatings. These modifications have led to the development of exceptional adsorbents that exhibit improved sorption capabilities by increasing surface area, pore size and volume, and surface binding sites [17]. These modifications have led to the development of special adsorbents that exhibit improved adsorption performance through increased surface area, pore size, volume, and surface binding sites [17].
Recently, biopolymer/clay bio-nanocomposites have gained attention as appealing materials for removing pollutants from contaminated water thanks to their biodegradability, biocompatibility, large surface area, and remarkable sorption performance [18]. Among the several varieties of biopolymers (such as chitosan, alginate, and starch), cellulose is ubiquitous and renewable and has abundant natural availability. As a result, it is drawing a lot of attention. Further, it is biocompatible, biodegradable, has high mechanical strength, is environmentally efficient, and does not contain toxic elements. Cellulose is made up of linear polysaccharides that are made up of two anhydroglucose rings that are associated with several β-1,4 glycosidic bonds [19]. Nevertheless, cellulose’s chemical composition makes it more likely to absorb cationic than anionic dyes. The reason for this is that the lignocellulose composition comprises multiple anions (such as hydroxyl (-OH) and carbonyl (-CO-) groups) serving as active sites. Moreover, it usually has a negative surface charge [20]. In this area, cellulose-based materials [21], hydrogels [22], and agricultural and other natural plant-derived wastes it is predominantly cellulose (40–50%), such as rice husk, sawdust, corn stalks, orange peel, wheat straw, cotton waste, banana waste, coffee waste, date palm tree spathe sheath and many other resources [6,23], and have been presented for the effective separation numerous pollutants including cationic dyes from wastewater. Several inorganic-organic hybrid composites are efficient adsorbents in a wide variety of experiments. These composites, which are used for eliminating different contaminants, i.e., cellulose/montmorillonite mesoporous beads [24], hydrogel based on cellulose and clay [25], kaolinite/biopolymer composites [26], cellulose@organically modified montmorillonite [27], cellulose/kaolinite–zeolite composite [28], are employed to remove different pollutants. They have shown promising efficiency for sequestering dyes. Because of the benefits offered by the materials described in the previous paragraph (Kaol, Cel), we decided to synthesize Kaol-Cel biopolymers as a recoverable adsorbent for removing CV dye from aquatic environments. This is the first cellulose-based compound extracted from red bean peels (Phaseolus vulgaris) (RBPs), which gives this study its innovative quality. To achieve maximum CV dye removal by Kaol/Cel, the experimental critical parameters needed to be improved and predicted using response surface techniques, included in Box–Behnken design (BBD). The investigation involved determining the synthesis conditions, including the loading of Cel into the Kaol biopolymeric matrix, as well as other crucial factors related to adsorption (such as adsorbent injection, pH, temperature, and residence period). Moreover, the study explored the kinetics, isotherms, and thermodynamic functions associated with the technique.

2. Materials and Methods

2.1. Material, Methods, and Instruments

RBPs were acquired from a local farm situated in El-Oued, Southeast Algeria. The samples underwent a thorough cleaning process to eliminate any dust or debris that may have been present on their surfaces. They were subsequently dried at 70 °C for 7–8 h and finely ground using a mixer grinder to ensure the samples were impurities-free. Raw kaolinite was obtained from Aougrout, Adrar, Southwest Algeria. The dye used in the study was CV (C25N3H30Cl, λmax = 590 nm, molecular weight: 407.979 g/mol), purchased from Sigma-Aldrich. Chemical products, such as hydrogen peroxide (H2O2), hydrochloric acid (HCl), sodium hydroxide (NaOH), acetic acid (CH3COOH), sodium carbonate (Na2CO3), sodium hexametaphosphate (NaPO3)6, thiourea (SC(NH2)2), and sodium hypochlorite (NaOCl), were furnished by Merck (Darmstadt, Germany). The chemicals employed in this research were pure and of analytical quality. The characterization of Kaol-Cel– 25 was carried out using Fourier transform–infrared (FT-IR) analysis (Cary 600 series FT-IR spectrophotometer, Agilent technologies, Santa Clara, CA, USA [29]), X-ray diffraction (XRD) analysis (PANalytical X’Pert PRO diffractometer), scanning electron microscopy (SEM)-energy-dispersive X-ray (EDX) analysis (Hitachi, TM3030Plus, Tabletop Microscope), Brunauer–Emmett–Teller (BET) analysis (Micromeritics ASAP2020 system), and zero-point of charge (pHpzc) value was derived using the methodology indicated by Kaouah et al. [30].

2.2. Pretreatment of Raw Clay

To fraction the clay samples, the method followed here was founded on several processes already applied in previous works [31,32,33]: (i) A series of sieves were used in a cascade to perform preliminary pre-sieving on the unprocessed clay to obtain particle sizes of less than 5 μm; (ii) After that, it was ground into a powder gently and then rinsed with H2O2 (40 mL, 6% w/v) to remove any organic compounds; (iii) After that, the resulting solution was transferred to an Erlenmeyer (1000 mL) was added to the buffer solution of pH 4.8 (80 mL) (16 g sodium acetate and 10 mL of acetic acid); (iv) The resulting solution was transferred to a graduated tube with the addition of a dispersible component, sodium hexametaphosphate (NaPO3)6, and allowed to decant for 7 h and 45 min after that floating layer was collected at a depth of 10 cm to produce particles smaller than 2 μm. The granules that were acquired were treated numerous times employing deionized water and centrifuged; finally, (v) One glass slide served as the control and was not subjected to any treatment, while the other slide was subjected to heating at 105 °C for two hours to pick the suitable clay. Glass slides were immersed in a glycerin bath until they became saturated and were then analyzed using XRD to identify the presence of the clay mineral Kaolinite: Al2Si2O5(OH)4, which had the following composition: 39.7% alumina, 46.2% silica, and 13.8% water. The BET procedure was utilized for calculating the material’s specific area, which was 110.786 m2/g.

2.3. Extracting Cellulose from Red Bean Peels (RBPs)

The procedure shown in Figure 1 was used to isolate cellulose from the powder obtained from RBPs. The process, detailed in several publications [29,34,35], was slightly modified. Initially, 100 g of dried powder from RBPs was cooked in 3 L of hot water with a pH of 7 for 15–25 min and then filtered. The resulting mass was subjected to acid (HCl) pretreatment by mixing it twice with a 1 M HCl solution (500 mL) at 80–90 °C for 60 min; the residue was filtered and collected. For alkali (NaOH) pretreatment, the agitation of the residue occurred three times. In a 1.5 L solution of 1 M NaOH for 60 min at 80–85 °C and then collected. The filtered residue was then subjected to bleach treatment twice with a 4% (w/v) NaOCl solution (at pH = 5 adjusted with 10% (v/v) CH3COOH) for 60 min at 80–90 °C, resulting in white-colored cellulose. The cellulose was rinsed with hot deionized water multiple times until the filtrate pH was neutralized. Finally, the cellulose was freeze-dried for seven hours and ground using a mixer grinder. It was then kept at room temperature for further investigation.

2.4. Preparation of Kaol/Cel Composite

To create Kaol/Cel nanocomposites, the researchers followed a specific procedure [36,37]. Initially, acidification is conducted by dissolving 20 g of the pretreated kaolinite into 200 mL of sulfuric acid (H2SO4) (15% W) in the beaker glass and stirring for 4 h at 80 °C. Then, 10 g of activated kaolinite was mixed with 16 mL of (46% W) NaOH in ice water and subjected to magnetic stirring for six hours. To optimize the process according to BBD, the team loaded different ratios of Cel to the modified Kaol before dissolving it with 0.6 M NaOH and 1 M thiourea solutions. Finally, the resulting mixture was vigorously stirred at 80 °C for four hours and then oven-dried at 60 °C overnight.

2.5. Experimental Design

BBD is an effective method for optimizing processes since it conforms to a quadratic surface. As a result, it was included in the overall design of the experiment [38]. A total of forty-six separate tests were conducted to assess the impact of five primary independent variables on the removal of CV. To determine the ideal parameters and specify the experimental domain, preliminary tests were conducted. These studies focused on the following aspects: the loading of Cel into Kaol (A), the adsorbent injection (B), the pH (C), the residence period (D), and the temperature (E). The various experimental levels of independent variables and their respective codes are listed in Table 1. The BBD, as well as the statistical data, were examined with the help of software called Stat-Ease Design-Expert (Version 13.0). For the purposes of assessing and forecasting CV removal, the second-order polynomial model was used. Including all square terms, linear terms, and linear-by-linear interaction items, the quadratic response model can be defined as follows:
Y = β 0 + i = 1 k β i   χ i + i = 1 k β i i   χ i 2 + i = 1 k j = 1 k β i j   χ i χ j + ε
where Y represents the objective for optimizing the response, while k denotes the number of variables being considered. The indices i and j are used to represent the variable numbers, and β0 is the constant coefficient, with βi and βii representing the linear and quadratic coefficients, respectively. The term βij refers to the interaction coefficient, and ε represents a random error. The values Xi and Xj correspond to the response of CV dye removal and coded values for the independent factors (−1, 0, and +1). In Equation (1), a positive sign implies that the variables have a synergistic effect, whereas a negative sign indicates that they have an antagonistic effect.
To evaluate the precision of the model, the researchers employed Analysis of Variance (ANOVA) to examine the coefficients. The ANOVA provided several parameters, including p-value, F-value, determination coefficient (R2), projected determination coefficient (R2pred), adjusted determination coefficient (R2adj), acceptable precision, degree of freedom (df), and standard deviation (SD). The experimental data and model precision were evaluated using these parameters [39]. The researchers used a dependable second-order quadratic model equation to predict the optimal value and describe the interactions between the elements. To determine the factors’ best values, we solved the regression equation, evaluated the counter-response surface map, and set limitations for the variable levels. To establish the variables’ extreme values, preliminary tests have been performed.

2.6. Batch Adsorption Studies

To investigate the adsorption capacity of Kaol/Cel for removal of CV dye, a BBD experimental was employed using varying amounts (0–50%) of the adsorbent with 100 mL of dye solution at concentrations of 50–300 mg/L. The experiments were conducted at different temperature intervals (30–60 °C) and pH values ranging from 4 to 10, with adjustments made using 0.1 N HCl and NaOH solutions at different intervals (5–120 min), as detailed in Table 2. Following the adsorption experiments, the residual concentrations were determined using a Cary Series UV-vis spectrophotometer at λmax = 590 nm after samples were spun at 3400 rpm for 10 min. The adsorption capacity of Kaol/Cel for CV dye removal (qe; mg/g) and the percentage of dye removal (R%) were calculated using Equations (2) and (3), respectively.
R   % = C 0 C e C 0 × 100
q e = V W C 0 C e
where: C0 (mg/L) and Ce (mg/L) are the initial and equilibrium CV dye concentrations, respectively, V (L) is the dye solution’s volume, and W is Kaol/Cel weight in grams (g).

3. Results and Discussion

3.1. Adsorbent Characterization

XRD patterns of initial cellulose are shown in Figure 2a–c. For the XRD pattern of kaolinite (a), the well-known diffraction peaks at 2θ around 6.0°, 19.6°, 20.8° 26.5°and 29.3° were attributed to the planes of (110), (002), (020) and (021), respectively, all of which were in agreement with the different characteristic orientation of the layers in the crystal structure of kaolinite (Kaol) [19]. For the XRD pattern of cellulose (b), the peaks observed at 2θ around 5.7°, 14.1°, 20.8°, 27.0°, and 29.0° are attributed to the (101), (10-1), (002), (110), and (10-2) planes, respectively, in the cellulose structure. These peak positions represent the interplanar spacing and arrangement of the cellulose chains [40]. The XRD pattern of the Kaol/Cel–25 composite(c) shows a significant decrease in peak intensities compared to the individual kaolin and cellulose patterns. This decrease in density is due to integrating the cellulose molecules into the clay framework. In addition, this work provided further evidence that Kaol particles may be successfully synthesized in the Kaol/Cel–25 formulation [19,40].
FT-IR analysis was employed to compare the functional groups on the surface of Kaol/Cel–25 before and after CV adsorption, as illustrated in Figure 3a,b, respectively. The FT-IR spectra of Kaol/Cel–25 before CV adsorption (Figure 3a) showed bands at 1633 cm−1, which could be attributed to O-H stretching and coordinated water bending vibrations, respectively [16,41]. The Si-O-Si stretching vibrations of kaolinite or quartz cause the peak at 803 cm−1, whereas the peaks at 3625 cm−1 are caused by Al-OH-Al and Fe-OH-Al deformation. The spectrum of the utilized cellulose was utilized to determine the various chemical groups of cellulose, such as β-glycosidic linkages at approximately 987 cm−1, -C-O-C pyranose rings at around 1630 cm−1, -C-H groups at around 995 cm−1 as well as approximately 2157 cm−1 [42]. The FT-IR spectrum of Kaol/Cel–25 after CV dye adsorption (Figure 3b) displayed the same bands in the spectrum before, with a slight shifting of some bands, indicating the involvement of the functional groups of Kaol in the adsorption of CV dye. Moreover, a new band that appeared at the approximately 1744 cm−1 band is attributed to the elongation of the C=N bond. The 1582 cm−1 band is due to aromatic C–C stretches, while the vibration of aromatic tertiary amine N–C is observed at 1366 cm−1, indicating that the important functional groups of Kaol/Cel–25 were responsible for the adsorption of CV dye [43].
Figure 4a depicts the surface morphology of Kaol, which may be described as an uneven and heterogonous surface with crevices. These fissures can be seen across the surface. The EDX spectrum of Kaol reveals the presence of O, Al, Si, K, Fe, and Ti. These elements are present in various minerals such as kaolinite, quartz, and other clay constituents, as identified through XRD analysis. Before CV dye adsorption, the result of the SEM-EDX analysis for Kaol/Cel can be shown in Figure 4b. The appearance of the surface morphology is a surface with protrusions of varying widths with many big spaces and fissures. The EDX analysis verifies the existence of the components O, C, Al, Si, K, Fe, and Ca. This suggests that the clay matrix effectively incorporated the cellulose particles.
Nevertheless, the morphological structure of Kaol/Cel–25 following CV dye absorption (Figure 4c) was converted into a different state, resulting in a reduced number of many tiny apertures and surface slits of the material. This observation provides evidence that molecules of CV dye have been loaded onto the surface of Kaol/Cel–25. In addition, the EDX examination reveals a spike in the carbonation rate in the related EDX spectrum, which provides more evidence that CV is being adsorbed to the Kaol/Cel–25 surface.

3.2. Fitting the Process Models

As was indicated before, BBD was responsible for the design of forty-six (46) different experiments (Table 2). The BBD method was utilized to explore the individual and interaction impacts that each of the examined factors had on the CV removal efficiency (as a response). The examined factors were the temperature (D), the pH (C), the adsorbent dose (A), and the temperature (D). These four factors were considered to be independent process factors.
The quadratic polynomial model established the mathematical relationship between the process components and the response. The regression model’s significance test for the response was conducted, and the findings of the ANOVA are presented in Table 3. The model’s F-value is 5.66, and the p-value is less than 0.0001, indicating the significance of the model terms. A p-value of < 0.05 is seen as important under the selected circumstances. The essential model terms for the response, i.e., the elimination of CV, were identified as A, B, BC, BD, and A2. However, it was observed that the residence period factor (E) had no effect on the CV elimination, possibly due to the adsorption process’s low sensitivity to time changes. The empirical relationship between CV elimination and the significant variables is expressed by the quadratic regression model given in Equation (4).
CV removal (%) = +97.97 + 11.83A + 5.89B − 8.95BC + 6.53BD − 9.66A2
Table 3. Analysis of variance (ANOVA) of the crystal violet (CV) dye removal response surface quadratic model (df: degree of freedom).
Table 3. Analysis of variance (ANOVA) of the crystal violet (CV) dye removal response surface quadratic model (df: degree of freedom).
SourceSum of
Squares
dfMeanSquareF-Valuep-ValueRemarks
Model4461.3020223.075.66<0.0001Significant
A: Cel loading2242.1112242.1156.92<0.0001Significant
B: Adsorbent dose555.691555.6914.110.0009Significant
C-pH106.311106.312.700.1130Insignificant
D-Temp.7.1917.190.18260.6728Insignificant
E-Time2.5712.570.06520.8006Insignificant
AB0.001510.00150.00000.9951Insignificant
AC3.0413.040.07720.7834Insignificant
AD49.21149.211.250.2743Insignificant
AE26.74126.740.67870.4178Insignificant
BC321.011321.018.150.0085Significant
BD170.741170.744.330.0477Significant
BE0.904610.90460.02300.8808Insignificant
CD0.712110.71210.01810.8941Insignificant
CE0.808010.80800.02050.8873Insignificant
DE1.8511.850.04690.8303Insignificant
A2814.981814.9820.690.0001Significant
B2166.501166.504.230.0504Insignificant
C23.4813.480.08820.7689Insignificant
D211.12111.120.28230.5999Insignificant
E20.179610.17960.00460.9467Insignificant
Residual984.812539.39
Cor Total5446.1145
A positive sign in Equation (4) indicates that the factor has a synergistic impact, while a negative sign means that the factor has an antagonistic effect [44]. As shown in Figure 5, the R-squared (determination coefficient) for the response factors is 0.97, which is a very high number indicating an excellent correlation between the real and projected values.

3.3. Interactions Significant for Crystal Violet (CV) Dye Removal

The influence of the adsorbent dose (B) and the solution pH (C) on the CV elimination efficacy showed a significant interaction (as evidenced by the p-value of 0.0085 from Table 3). The remaining independent variables (i.e., Cel loading (Kaol/Cel–25), residence period of 5 min, and temperature of 45 °C), were held constant during the experimentation range. Figure 6a,b depict the 3D surface and 2D contour plots for the BC interaction, respectively.
Figure 5 depicts that augmenting the maximum adsorbent injection from 0.02 g to 0.05 g and decreasing the solution pH from 10 to 4 significantly augmented CV elimination efficacy from 61.82% to 99.86%. The greatest CV removal efficiency was achieved at pH 4, with a gradual decrease in dye removal as the pH value increased toward an alkaline environment. Furthermore, Figure 7 shows that the pHpzc of Kaol/Cel–25 was 8.71, establishing that the surface of Kaol/Cel–25 begins to be positively charged at pH values less than pHpzc. Conversely, at pH values above pHpzc, the surface charge of Kaol/Cel–25 becomes negative, which implies that Kaol/Cel–25 may adsorb cationic dyes. Consequently, more essential electrostatic attractions arise between the surface functional groups of negatively charged Kaol/Cel–25 and cationic CV dye, as shown in Equation (5).
Materials 16 04082 i001
Table 3 was used to assess the significance of the interactions between the independent variables. The results showed that the interaction influence between adsorbent dosage (B) and temperature (D) had a statistically considerable impact on CV removal efficiency, with a p-value of 0.0477. The other variables, namely a 25% Cel loading, a pH of 10, and a 5 min contact time, remained constant throughout the experiment. The interaction between the adsorbent dose and the temperature was further analyzed using 3D and 2D response surface plots, as shown in Figure 8a,b, respectively. The plots indicated that augmenting the adsorbent dose from 0.02 g to 0.05 g led to higher CV dye removal (%), possibly due to increased active adsorption sites or available surface area. Nevertheless, the temperature seemed to possess no significant influence on eliminating CV dye, suggesting an exothermic adsorption process. These results will be discussed further in Section 3.7 on Adsorption Thermodynamics.
It was determined whether there was a statistically crucial interaction between each pair of independent variables (as shown in Table 3). The results revealed that the interaction influence between adsorbent dosage (B) and temperature (D) had a significant impact on the CV removal efficiency (p-value = 0.0477). The other independent factors, including Cel loading of 25%, solution pH of 10, and contact time of 5 min, remained constant throughout the experiment. Figure 8a,b illustrate the response surface plots for the interaction between temperature and adsorbent injection. The findings show that augmenting the adsorbent injection from 0.02 g to 0.05 g resulted in higher removal of CV, which can be related to an elevation in active adsorption sites and/or surface area. Moreover, the results suggest that the adsorption of dye molecules onto the Kaol/Cel–25 surface was exothermic since the temperature had no discernible influence on CV removal effectiveness. A more detailed discussion on adsorption thermodynamics is included in Section 3.7.

3.4. Adsorption Studies

In this work, the influence of residence period vs. initial CV level on the adsorption potential of Kaol/Cel–25 was explored over a range of initial levels from 50 to 300 mg/L. Figure 9 shows the experimental results, while the optimal conditions for CV dye removal were determined based on the highest achieved CV dye removal percentage, observed at a constant Kaol/Cel–25 adsorbent injection (0.035 g), pH 10, and 45 °C, as indicated in Run #24 of Table 2. As the initial CV level was augmented, the equilibrium adsorption capacity was also augmented from 51.7 to 297.7 mg/g, as presented in Figure 9. This increase is a result of a greater collision rate between CV dye and the Kaol/Cel–25 surface, resulting from the elevated initial CV concentration. Additionally, the higher concentration gradient facilitated the diffusion of dye molecules into the internal pores of the adsorbent, leading to their migration towards the active adsorption sites, thus enhancing the adsorption capacity [45].

3.5. Kinetic Modeling

To comprehend the adsorption pathway and behavior of CV on the surface of Kaol/Cel–25, further investigations were carried out, and understanding the adsorption kinetics is crucial. In this context, two kinetic models, pseudo-first-order (PFO) and pseudo-second-order (PSO), were utilized. These non-linear kinetic models are represented by Equation (6) for PFO [46] and Equation (7) for PSO [47]:
q t = q e   1 e x p k 1   t
q t = q e   k 2 2   t 1 + q e k 2 t
designated by qe (mg/g). In contrast, the quantity of CV dye adsorbed at a specific time t is designated by qt (mg/g). Additionally, the rate constant of the PFO is represented by k1 (1/min), and the rate constant of the PSO is represented by k2 (g/mg min).
Table 4 displays the correlation coefficients (R2) and model parameters for the PSO and PFO. On the basis of the kinetic adsorption data in Table 4, it can be shown that the PSO model fits the Kaol/Cel–25 dye adsorption on the Kaol/Cel–25 surface better than the PFO model. This is confirmed by the fact that the PSO model had higher R2 values than the PFO model. Furthermore, the qe (i.e., qe,cal) values estimated with the PSO model are closer to the experimental qe (i.e., qe,exp) values than the qe values calculated with the PFO model. These findings and FT-IR analysis suggest that chemical interaction between CV dye and active functional groups on the Kaol/Cel–25 surface is essential in the adsorption process. In the Kaol/Cel–25 surface, the adsorption of CV dye is thus governed by the chemisorption phenomenon.

3.6. Isotherms for Adsorption

The adsorption isotherm is a critical piece of information to have when attempting to describe the interaction between the molecules of CV dye and Kaol/Cel–25. As a result, the isotherms that are applied the most often, namely Langmuir [48], Freundlich [49], and Temkin [50], are selected for the purpose of conducting an analysis of the data pertaining to equilibrium adsorption and determining the adsorption potential of Kaol/Cel–25. The non-linear forms of the Langmuir, Freundlich, and Temkin equations are each represented in their own separate Equations (8)–(10):
C e q e = q max   K a       C e 1 + K a   C e
q e = K f   C e 1 n
q e = R T b T ( ln K T C e )
The variables used in the analysis include Ce (mg/L) for the equilibrium concentration of CV dye, qmax (mg/g) for the maximum quantity of CV dye that can be adsorbed per unit mass of Kaol/Cel–25, and qe (mg/g) for the quantity of CV dye absorbed per unit weight. The constants used include Langmuir (Ka), Freundlich (Kf), and Temkin (KT) constants (L/mg), as well as n for the dimensionless constant indicating adsorption intensity, bT (J/mol) for the heat of adsorption, T (K) for temperature, and R (8.314 J/mol K) for the universal gas constant.
The absorption of CV dye by Kaol/Cel–25 is shown in Table 5 to be best fit by the Langmuir adsorption isotherm model, which has a higher R2 (0.99) than the Freundlich and Temkin models. This implies that CV dye form a monolayer coverage on the homogeneous surface of the adsorbent that are energetically equivalent [51]. Moreover, using the Langmuir model, we determined that 294.12 mg/g is the qmax for Kaol/Cel–25. Table 6 compares the qmax of CV dye adsorption onto Kaol/Cel–25 to that of other adsorbents that have been published for removing CV, highlighting that the cationic dye removal efficiency of Kaol/Cel–25 is highly effective and shows promise for further applications.

3.7. Thermodynamic Functions Results

To assess the viability and spontaneity of the CV dye adsorption phenomenon onto the surface of Kaol/Cel–25 and estimate the level of randomness at the interface between the dye and the surface, various adsorption thermodynamic parameters were determined. These parameters include the Gibbs free energy change (ΔG°) (kJ/mol), the entropy change (ΔS°) (kJ/mol. K) and the enthalpy change (ΔH°) (kJ/mol). Such parameters thermodynamic were computed using Equations (11)–(13) [61]:
G = R T L n   K d
K d = q e C e  
L n K d = Δ S °   R   Δ H ° R T °
Figure 10 depicts a plot of lnKd against 1/T, from which the thermodynamic parameters (ΔH° and ΔS°) were calculated. The slope of this plot is ΔH°, and the intercept is ΔS°.
The results presented in Table 7 show that the Gibbs free energy change (ΔG°) for the adsorption of CV dye onto the surface of Kaol/Cel–25 is negative, indicating that the phenomenon is spontaneous [62]. Furthermore, the negative enthalpy values obtained for the adsorption process of CV by Kaol/Cel–25 suggest that the process is exothermic, which is consistent with the results obtained from the BBD parametric optimization depicted in Figure 10. Additionally, the adsorption of CV onto Kaol/Cel–25 appears to cause a greater level of disorder at the interface between the solid and solution, as indicated by the negative entropy values [63].

3.8. Mechanisms of Adsorption

A proposed adsorption mechanism for CV by Kaol/Cel–25 is illustrated in Figure 11. The various functional groups existing on the surface of Kaol/Cel–25 participate in different interactions that facilitate the adsorption pathway of CV. Notably, the most impactful interaction is an electrostatic attraction between the CV dye molecules and the surface of the Kaol/Cel–25 adsorbent. Such an adsorption pathway encompasses the electrostatic interaction between CV dye cations and the negatively charged sites on the surface of Kaol/Cel–25. Additionally, there is the possibility of n-π interaction, which typically occurs when the lone pair electrons on an O are spread out into the orbital of an aromatic ring of dye [64]. Two kinds of hydrogen bonding take place between the Kaol/Cel–25 and the molecular structure of CV. The first and more frequent kind is dipole–dipole hydrogen bonding interaction between present O on the surface of Kaol/Cel–25 with the O and N atoms existing in the CV molecular structure. This interaction is illustrated in Figure 11. The last kind is Yoshida H-bonding, which takes place between -OH groups on the surface of Kaol/Cel–25 and the CV aromatic ring [65]. Based on the information presented above, it can be inferred that these interactions had a crucial contribution in improving the adsorption phenomenon of CV dye on the surface of Kaol/Cel–25.

4. Conclusions

This research aimed to develop an environmentally friendly composite material called kaolinite-cellulose (Kaol/Cel) using waste red bean peels (Phaseolus vulgaris) as a cellulose source. A Box–Behnken design was employed to perform parametric optimization of the synthesis conditions and adsorptive performance of the Kaol/Cel composite to remove crystal violet (CV) dye. The composite’s polymeric matrix with uniform Cel distribution was found to contribute to its elevated adsorptive behavior towards CV. The best parameters for CV elimination (99.58%) were identified as 25% Cel loading, 0.035 g adsorbent dosage, pH 10, 45 °C temperature, and 5 min contact time. The equilibrium data were best represented by the Freundlich isotherm model, and the Langmuir model revealed a maximum adsorption capacity (qmax) of 294.12 mg/g at 45 °C. The adsorption mechanism involved several interactions, including electrostatic attractions, n-π stacking interactions, dipole–dipole hydrogen bonding interactions, and Yoshida H-bonding. Additionally, the adsorption process of CV by Kaol/Cel–25 was observed to be exothermic and spontaneous based on adsorption thermodynamic functions. These findings suggest that Kaol/Cel–25 could be a cost-effective composite biosorbent for removing cationic dyes from aquatic environments and could have potential applications in wastewater treatment. This refers to the elimination of organic and inorganic pollutants, as well as the diminution of the chemical oxygen demand.

Author Contributions

Conceptualization: N.E., A.A.M.S., R.M. and A.Z.; methodology: R.M., S.G., D.G. and A.Z.; formal analysis: R.M., S.A., S.N.N., M.B., N.E. and D.G.; investigation: R.M., S.A.A., N.E., A.A.M.S. and S.A.; resources: N.E. and R.M.; writing original draft preparation: R.M. and D.G.; writing review and editing: S.N.N. and M.B.; supervision: N.E. All authors have read and agreed to the published version of the manuscript.

Funding

This research has been funded by the Research Deanship of University of Ha’il, Saudi Arabia, through the Project RG-21 014.

Acknowledgments

This research has been funded by the Research Deanship of University of Ha’il, Saudi Arabia, through the Project RG-21 014. The authors express their gratitude to the General Directorate of Scientific Research and Technological Development in Algeria for providing lab facilities.

Conflicts of Interest

The authors state that they do not have any financial interest or personal relationships that may have influenced the work presented in this article.

References

  1. Al-Tohamy, R.; Ali, S.S.; Li, F.; Okasha, K.M.; Mahmoud, Y.A.-G.; Elsamahy, T.; Jiao, H.; Fu, Y.; Sun, J. A critical review on the treatment of dye-containing wastewater: Ecotoxicological and health concerns of textile dyes and possible remediation approaches for environmental safety. Ecotoxicol. Environ. Saf. 2022, 231, 113160. [Google Scholar] [CrossRef]
  2. Barasarathi, J.; Abdullah, P.S.; Uche, E.C. Application of magnetic carbon nanocomposite from agro-waste for the removal of pollutants from water and wastewater. Chemosphere 2022, 305, 135384. [Google Scholar] [CrossRef] [PubMed]
  3. Gurr, E. Synthetic Dyes in Biology, Medicine and Chemistry; Elsevier: Amsterdam, The Netherlands, 2012. [Google Scholar]
  4. Nohynek, G.J.; Fautz, R.; Benech-Kieffer, F.; Toutain, H. Toxicity and human health risk of hair dyes. Food Chem. Toxicol. 2004, 42, 517–543. [Google Scholar] [CrossRef]
  5. Kishor, R.; Purchase, D.; Saratale, G.D.; Saratale, R.G.; Ferreira, L.F.R.; Bilal, M.; Chandra, R.; Bharagava, R.N. Ecotoxicological and health concerns of persistent coloring pollutants of textile industry wastewater and treatment approaches for environmental safety. J. Environ. Chem. Eng. 2021, 9, 105012. [Google Scholar] [CrossRef]
  6. Haque, A.N.M.A.; Sultana, N.; Sayem, A.S.M.; Smriti, S.A. Sustainable adsorbents from plant-derived agricultural wastes for anionic dye removal: A review. Sustainability 2022, 14, 11098. [Google Scholar] [CrossRef]
  7. Zobeidi, A.; Bebba, A. Seasonal variations of physical, chemical parameters in a wastewater treatment plant by aerated lagoons at Southern-East of Algeria. Res. J. Pharm. Biol. Chem. Sci. 2015, 6, 1097–1102. [Google Scholar]
  8. Salleh, M.A.M.; Mahmoud, D.K.; Karim, W.A.W.A.; Idris, A. Cationic and anionic dye adsorption by agricultural solid wastes: A comprehensive review. Desalination 2011, 280, 1–13. [Google Scholar] [CrossRef]
  9. Hussain, D.; Khan, S.A.; Alharthi, S.S.; Khan, T.A. Insight into the performance of novel kaolinite-cellulose/cobalt oxide nanocomposite as green adsorbent for liquid phase abatement of heavy metal ions: Modelling and mechanism. Arab. J. Chem. 2022, 15, 103925. [Google Scholar] [CrossRef]
  10. Deliyanni, E.; Peleka, E.; Lazaridis, N. Comparative study of phosphates removal from aqueous solutions by nanocrystalline akaganéite and hybrid surfactant-akaganéite. Sep. Purif. Technol. 2007, 52, 478–486. [Google Scholar] [CrossRef]
  11. Shakya, A.; Agarwal, T. Removal of Cr(VI) from water using pineapple peel derived biochars: Adsorption potential and re-usability assessment. J. Mol. Liq. 2019, 293, 111497. [Google Scholar] [CrossRef]
  12. Haque, A.N.M.A.; Remadevi, R.; Rojas, O.J.; Wang, X.; Naebe, M. Kinetics and equilibrium adsorption of methylene blue onto cotton gin trash bioadsorbents. Cellulose 2020, 27, 6485–6504. [Google Scholar] [CrossRef]
  13. Mabrouk, S.; Bebba, A.A.; Kamarchou, A.; Zobeidi, A. Adsorption capacity of pollutants by using local clay mineral from urban wastewater Touggourt (South-East Algeria). Asian J. Res. Chem. 2020, 13, 85–90. [Google Scholar] [CrossRef]
  14. Deng, L.; Yuan, P.; Liu, D.; Annabi-Bergaya, F.; Zhou, J.; Chen, F.; Liu, Z. Effects of microstructure of clay minerals, montmorillonite, kaolinite and halloysite, on their benzene adsorption behaviors. Appl. Clay Sci. 2017, 143, 184–191. [Google Scholar] [CrossRef]
  15. Narayana Saibaba, K.V. Kaolinite-Chitosan based Nano-Composites and Applications. Adv. Appl. Micro Nano Clay Biopolym. Based Compos. 2022, 125, 87–102. [Google Scholar]
  16. Atia, D.; Bebba, A.A.; Haddad, L.; Zobeidi, A. Elimination of organic pollutants from urban wastewater by illite-kaolinite local clay from south-east of Algeria. Cienc. Tecn. Vitivinic. 2018, 33, 17–28. [Google Scholar]
  17. Khan, T.A.; Dahiya, S.; Khan, E.A. Removal of direct red 81 from aqueous solution by adsorption onto magnesium oxide-coated kaolinite: Isotherm, dynamics and thermodynamic studies. Environ. Prog. Sustain. Energy 2017, 36, 45–58. [Google Scholar] [CrossRef]
  18. del Mar Orta, M.; Martín, J.; Santos, J.L.; Aparicio, I.; Medina-Carrasco, S.; Alonso, E. Biopolymer-clay nanocomposites as novel and ecofriendly adsorbents for environmental remediation. Appl. Clay Sci. 2020, 198, 105838. [Google Scholar] [CrossRef]
  19. Putro, J.N.; Santoso, S.P.; Ismadji, S.; Ju, Y.-H. Investigation of heavy metal adsorption in binary system by nanocrystalline cellulose–bentonite nanocomposite: Improvement on extended Langmuir isotherm model. Microporous Mesoporous Mater. 2017, 246, 166–177. [Google Scholar] [CrossRef]
  20. Sajjadi, B.; Zubatiuk, T.; Leszczynska, D.; Leszczynski, J.; Chen, W.Y. Chemical activation of biochar for energy and environmental applications: A comprehensive review. Rev. Chem. Eng. 2019, 35, 777–815. [Google Scholar] [CrossRef]
  21. Jamshaid, A.; Hamid, A.; Muhammad, N.; Naseer, A.; Ghauri, M.; Iqbal, J.; Rafiq, S.; Shah, N.S. Cellulose-based Materials for the Removal of Heavy Metals from Wastewater—An Overview. ChemBioEng Rev. 2017, 4, 240–256. [Google Scholar] [CrossRef]
  22. Akter, M.; Bhattacharjee, M.; Dhar, A.K.; Rahman, F.B.A.; Haque, S.; Rashid, T.U.; Kabir, S.F. Cellulose-based hydrogels for wastewater treatment: A concise review. Gels 2021, 7, 30. [Google Scholar] [CrossRef] [PubMed]
  23. Tedjani, C.F.; Mya, O.B.; Rebiai, A. Isolation and characterization of cellulose from date palm tree spathe sheath. Sustain. Chem. Pharm. 2020, 17, 100307. [Google Scholar] [CrossRef]
  24. Pan, Y.; Xie, H.; Liu, H.; Cai, P.; Xiao, H. Novel cellulose/montmorillonite mesoporous composite beads for dye removal in single and binary systems. Bioresour. Technol. 2019, 286, 121366. [Google Scholar] [CrossRef]
  25. Wang, Q.; Wang, Y.; Chen, L. A green composite hydrogel based on cellulose and clay as efficient absorbent of colored organic effluent. Carbohydr. Polym. 2019, 210, 314–321. [Google Scholar] [CrossRef] [PubMed]
  26. Manna, M.S.; Ghanta, S. Kaolinite–Cellulose based Nano–Composites and Applications. Adv. Appl. Micro Nano Clay Biopolym. Based Compos. 2022, 125, 275–301. [Google Scholar]
  27. Cai, J.; Lei, M.; Zhang, Q.; He, J.-R.; Chen, T.; Liu, S.; Fu, S.-H.; Li, T.-T.; Liu, G.; Fei, P. Electrospun composite nanofiber mats of Cellulose@ Organically modified montmorillonite for heavy metal ion removal: Design, characterization, evaluation of absorption performance. Compos. Part A Appl. Sci. Manuf. 2017, 92, 10–16. [Google Scholar] [CrossRef]
  28. Ashraf, M.-T.; AlHammadi, A.A.; El-Sherbeeny, A.M.; Alhammadi, S.; Al Zoubi, W.; Ko, Y.G.; Abukhadra, M.R. Synthesis of cellulose fibers/Zeolite-A nanocomposite as an environmental adsorbent for organic and inorganic selenium ions; Characterization and advanced equilibrium studies. J. Mol. Liq. 2022, 360, 119573. [Google Scholar] [CrossRef]
  29. Song, Y.K.; Chew, I.M.L.; Choong, T.S.Y.; Tan, J.; Tan, K.W. Isolation of Nanocrystalline Cellulose from oil palm empty fruit bunch—A response surface methodology study. MATEC Web Conf. 2016, 60, 04009. [Google Scholar] [CrossRef]
  30. Kaouah, F.; Boumaza, S.; Berrama, T.; Trari, M.; Bendjama, Z. Preparation and characterization of activated carbon from wild olive cores (oleaster) by H3PO4 for the removal of Basic Red 46. J. Clean. Prod. 2013, 54, 296–306. [Google Scholar] [CrossRef]
  31. Al Arni, S.; Mahmoud, E.; Converti, A.; Mhamed, B.; Alsamani, A.M.S.; Saad, G.; Nadir, A. Application of Date Palm Surface Fiber as an Efficient Biosorbent for Wastewater Treatment. ChemBioEng Rev 2023, 10, 55–64. [Google Scholar] [CrossRef]
  32. Taj, R.; El Askary, M.; Saad, N.; Basyoni, M. Mineralogical investigation and some sedimentary phenomena of Ubhur Formation, north Jeddah, Saudi Arabia. J. King Abdulaziz Univ. Mar. Sci. 2002, 13, 93–110. [Google Scholar] [CrossRef]
  33. Atia, D.; Zobeidi, A. An investigation into the mineralogical and physicochemical characterization of El-Oued (Algeria) clay. Neuroquantology 2022, 20, 1048–1058. [Google Scholar]
  34. Szymańska-Chargot, M.; Chylińska, M.; Gdula, K.; Kozioł, A.; Zdunek, A. Isolation and characterization of cellulose from different fruit and vegetable pomaces. Polymers 2017, 9, 495. [Google Scholar] [CrossRef]
  35. Khawas, P.; Deka, S.C. Isolation and characterization of cellulose nanofibers from culinary banana peel using high-intensity ultrasonication combined with chemical treatment. Carbohydr. Polym. 2016, 137, 608–616. [Google Scholar] [CrossRef]
  36. Abukhadra, M.R.; AlHammadi, A.; El-Sherbeeny, A.M.; Salam, M.A.; El-Meligy, M.A.; Awwad, E.M.; Luqman, M. Enhancing the removal of organic and inorganic selenium ions using an exfoliated kaolinite/cellulose fibres nanocomposite. Carbohydr. Polym. 2021, 252, 117163. [Google Scholar] [CrossRef]
  37. Mohadi, R. Adsorption of congo red using kaolinite-cellulose adsorbent. Sci. Technol. Indones. 2017, 2, 29–36. [Google Scholar]
  38. Özer, A.; Gürbüz, G.; Çalimli, A.; Körbahti, B.K. Biosorption of copper(II) ions on Enteromorpha prolifera: Application of response surface methodology (RSM). Chem. Eng. J. 2009, 146, 377–387. [Google Scholar] [CrossRef]
  39. Zhu, S.; Shi, R.; Wang, J.; Wang, J.-F.; Li, X.-M. Unpredictable chronic mild stress not chronic restraint stress induces depressive behaviours in mice. Neuroreport 2014, 25, 1151–1155. [Google Scholar] [CrossRef]
  40. Boushehrian, M.M.; Esmaeili, H.; Foroutan, R. Ultrasonic assisted synthesis of Kaolin/CuFe2O4 nanocomposite for removing cationic dyes from aqueous media. J. Environ. Chem. Eng. 2020, 8, 103869. [Google Scholar] [CrossRef]
  41. Shaban, M.; Sayed, M.I.; Shahien, M.G.; Abukhadra, M.R.; Ahmed, Z.M. Adsorption behavior of inorganic-and organic-modified kaolinite for Congo red dye from water, kinetic modeling, and equilibrium studies. J. Sol-Gel Sci. Technol. 2018, 87, 427–441. [Google Scholar] [CrossRef]
  42. Chen, C.; Sun, K.; Wang, A.; Wang, S.; Jiang, J. Catalytic graphitization of cellulose using nickel as catalyst. BioResources 2018, 13, 3165–3176. [Google Scholar] [CrossRef]
  43. Sabna, V.; Thampi, S.G.; Chandrakaran, S. Adsorption of crystal violet onto functionalised multi-walled carbon nanotubes: Equilibrium and kinetic studies. Ecotoxicol. Environ. Saf. 2016, 134, 390–397. [Google Scholar] [CrossRef] [PubMed]
  44. Abdulhameed, A.S.; Mohammad, A.-T.; Jawad, A.H. Application of response surface methodology for enhanced synthesis of chitosan tripolyphosphate/TiO2 nanocomposite and adsorption of reactive orange 16 dye. J. Clean. Prod. 2019, 232, 43–56. [Google Scholar] [CrossRef]
  45. Zain, Z.M.; Abdulhameed, A.S.; Jawad, A.H.; ALOthman, Z.A.; Yaseen, Z.M. A pH-sensitive surface of chitosan/sepiolite clay/algae biocomposite for the removal of malachite green and remazol brilliant blue R dyes: Optimization and adsorption mechanism study. J. Polym. Environ. 2023, 31, 501–518. [Google Scholar] [CrossRef]
  46. Lagergren, S. Zur theorie der sogenannten adsorption geloster stoffe. K. Sven. Vetenskapsakademiens. Handl. 1898, 24, 1–39. [Google Scholar]
  47. Ho, Y.-S.; McKay, G. Sorption of dye from aqueous solution by peat. Chem. Eng. J. 1998, 70, 115–124. [Google Scholar] [CrossRef]
  48. Langmuir, I. The adsorption of gases on plane surfaces of glass, mica and platinum. J. Am. Chem. Soc. 1918, 40, 1361–1403. [Google Scholar] [CrossRef]
  49. Al Arni, S.; Ghareba, S.; Solisio, C.; Palma, M.S.A.; Converti, A. Methods of reactive red 141 dye decolorization, treatment, and removal from industrial wastewaters: A critical review. Environ. Eng. Sci. 2021, 38, 577–591. [Google Scholar] [CrossRef]
  50. Temkin, M. Kinetics of ammonia synthesis on promoted iron catalysts. Acta Physiochim. URSS 1940, 12, 327–356. [Google Scholar]
  51. Foo, K.Y.; Hameed, B.H. Coconut Husk Derived Activated Carbon via Microwave Induced Activation: Effects of Activation Agents, Preparation Parameters and Adsorption Performance. Chem. Eng. J. 2012, 184, 57–65. [Google Scholar] [CrossRef]
  52. El-Sayed, G.O. Removal of methylene blue and crystal violet from aqueous solutions by palm kernel fiber. Desalination 2011, 272, 225–232. [Google Scholar] [CrossRef]
  53. Chakraborty, S.; Chowdhury, S.; Saha, P.D. Adsorption of crystal violet from aqueous solution onto NaOH-modified rice husk. Carbohydr. Polym. 2011, 86, 1533–1541. [Google Scholar] [CrossRef]
  54. Çoruh, S.; Geyikçi, F.; Ergun, O.N. Dye Removal from Aqueous Solution by Adsorption onto Fly Ash; ATINER’s Conference Paper Series ENV2012-0057; Athens Institute for Education and Research: Athens, Greece, 2012. [Google Scholar]
  55. El Naeem, G.A.; Abd-Elhamid, A.; Farahat, O.O.; El-Bardan, A.A.; Soliman, H.M.; Nayl, A. Adsorption of crystal violet and methylene blue dyes using a cellulose-based adsorbent from sugercane bagasse: Characterization, kinetic and isotherm studies. J. Mater. Res. Technol. 2022, 19, 3241–3254. [Google Scholar]
  56. Khan, S.A.; Rehman, T.U.; Shah, L.A.; Ullah, M. Magnetite graphene oxide-doped superadsorbent hydrogel for efficient removal of crystal violet from wastewater. Chem. Pap. 2023, 77, 2725–2735. [Google Scholar] [CrossRef]
  57. Uddin, M.K.; Abd Malek, N.N.; Jawad, A.H.; Sabar, S. Pyrolysis of rubber seed pericarp biomass treated with sulfuric acid for the adsorption of crystal violet and methylene green dyes: An optimized process. Int. J. Phytoremediat. 2022, 25, 393–402. [Google Scholar] [CrossRef] [PubMed]
  58. Abdulhameed, A.S.; Jawad, A.H.; Kashi, E.; Radzun, K.A.; ALOthman, Z.A.; Wilson, L.D. Insight into adsorption mechanism, modeling, and desirability function of crystal violet and methylene blue dyes by microalgae: Box-Behnken design application. Algal Res. 2022, 67, 102864. [Google Scholar] [CrossRef]
  59. Hanafi, N.A.M.; Abdulhameed, A.S.; Jawad, A.H.; ALOthman, Z.A.; Yousef, T.A.; Al Duaij, O.; Alsaiari, N.S. Optimized removal process and tailored adsorption mechanism of crystal violet and methylene blue dyes by activated carbon derived from mixed orange peel and watermelon rind using microwave-induced ZnCl2 activation. Biomass Convers. Biorefinery 2022, 1–13. [Google Scholar] [CrossRef]
  60. Jani, N.A.; Haddad, L.; Abdulhameed, A.S.; Jawad, A.H.; ALOthman, Z.A.; Yaseen, Z.M. Modeling and optimization of the adsorptive removal of crystal violet dye by durian (Durio zibethinus) seeds powder: Insight into kinetic, isotherm, thermodynamic, and adsorption mechanism. Biomass Convers. Biorefinery 2022, 1–14. [Google Scholar] [CrossRef]
  61. Jawad, A.H.; Mubarak, N.S.A.; Abdulhameed, A.S. Hybrid crosslinked chitosan-epichlorohydrin/TiO2 nanocomposite for reactive red 120 dye adsorption: Kinetic, isotherm, thermodynamic, and mechanism study. J. Polym. Environ. 2020, 28, 624–637. [Google Scholar] [CrossRef]
  62. Chang, J.; Ma, J.; Ma, Q.; Zhang, D.; Qiao, N.; Hu, M.; Ma, H. Adsorption of methylene blue onto Fe3O4/activated montmorillonite nanocomposite. Appl. Clay Sci. 2016, 119, 132–140. [Google Scholar] [CrossRef]
  63. Chebli, D.; Bouguettoucha, A.; Mekhalef, T.; Nacef, S.; Amrane, A. Valorization of an agricultural waste, Stipa tenassicima fibers, by biosorption of an anionic azo dye, Congo red. Desalination Water Treat. 2015, 54, 245–254. [Google Scholar] [CrossRef]
  64. Singh, S.K.; Das, A. The n → π* interaction: A rapidly emerging non-covalent interaction. Phys. Chem. Chem. Phys. 2015, 17, 9596–9612. [Google Scholar] [CrossRef] [PubMed]
  65. Parker, H.L.; Hunt, A.J.; Budarin, V.L.; Shuttleworth, P.S.; Miller, K.L.; Clark, J.H. The importance of being porous: Polysaccharide-derived mesoporous materials for use in dye adsorption. RSC Adv. 2012, 2, 8992–8997. [Google Scholar] [CrossRef]
Figure 1. Flow process diagram for cellulose extraction from red bean peels (RBPs).
Figure 1. Flow process diagram for cellulose extraction from red bean peels (RBPs).
Materials 16 04082 g001
Figure 2. X-ray diffraction (XRD) pattern of (a) Kaol, (b) Cel, and (c) Kaol/Ce–25.
Figure 2. X-ray diffraction (XRD) pattern of (a) Kaol, (b) Cel, and (c) Kaol/Ce–25.
Materials 16 04082 g002
Figure 3. Fourier-transform infrared (FT-IR) spectra of (a) Kaol/Cel–25, and (b) Kaol/Cel–25 after adsorption of CV dye.
Figure 3. Fourier-transform infrared (FT-IR) spectra of (a) Kaol/Cel–25, and (b) Kaol/Cel–25 after adsorption of CV dye.
Materials 16 04082 g003
Figure 4. Scanning electron microscopy (SEM) and energy-dispersive X-ray (EDX) spectrum of (a) Kaol, (b) Kaol/Cel–25, and (c) Kaol/Ce–25 after CV dye adsorption.
Figure 4. Scanning electron microscopy (SEM) and energy-dispersive X-ray (EDX) spectrum of (a) Kaol, (b) Kaol/Cel–25, and (c) Kaol/Ce–25 after CV dye adsorption.
Materials 16 04082 g004
Figure 5. Linear correlation between predicted values vs. the observed values of CV adsorption on Kaol/Cel–25.
Figure 5. Linear correlation between predicted values vs. the observed values of CV adsorption on Kaol/Cel–25.
Materials 16 04082 g005
Figure 6. (a) 3D surface, (b) 2D contour plot of the influence of adsorbent dose and solution pH of CV adsorption on Kaol/Cel–25.
Figure 6. (a) 3D surface, (b) 2D contour plot of the influence of adsorbent dose and solution pH of CV adsorption on Kaol/Cel–25.
Materials 16 04082 g006
Figure 7. The zero-point of charge (pHpzc) of CV adsorption on Kaol/Cel–25.
Figure 7. The zero-point of charge (pHpzc) of CV adsorption on Kaol/Cel–25.
Materials 16 04082 g007
Figure 8. (a) 3D surface, and (b) 2D contour plot of the influence of adsorbent dose and temperature on Kaol/Cel–25 CV adsorption.
Figure 8. (a) 3D surface, and (b) 2D contour plot of the influence of adsorbent dose and temperature on Kaol/Cel–25 CV adsorption.
Materials 16 04082 g008
Figure 9. Influence of CV adsorption contact time vs. initial concentrations on Kaol/Cel–25.
Figure 9. Influence of CV adsorption contact time vs. initial concentrations on Kaol/Cel–25.
Materials 16 04082 g009
Figure 10. Plot of Van’t Hoff equation for CV adsorption onto Kaol/Cel–25.
Figure 10. Plot of Van’t Hoff equation for CV adsorption onto Kaol/Cel–25.
Materials 16 04082 g010
Figure 11. Possible mechanism between Kaol/Cel–25 surface and crystal violet (CV) dye.
Figure 11. Possible mechanism between Kaol/Cel–25 surface and crystal violet (CV) dye.
Materials 16 04082 g011
Table 1. Box–Behnken design (BBD) codes for independent factor experimental levels.
Table 1. Box–Behnken design (BBD) codes for independent factor experimental levels.
FactorsLevels
Low (−1)Medium (0)High (+1)
A: Loading (%)02550
B: Adsorbent dose (g)0.020.0350.05
C: pH4710
D: Temperature (°C)304560
E: Contact time (min)565.5120
Table 2. Box–Behnken design (BBD) matrix with five factors and experimental results for crystal violet (CV).
Table 2. Box–Behnken design (BBD) matrix with five factors and experimental results for crystal violet (CV).
RunA: Cel Loading (%)B: Adsorbent Dose (g)C: pHD: Temperature (°C)E: Contact Time (min)Dye Removal (%)
1250.03574517.597.95
200.03576017.575.88
3250.02076017.572.02
4250.050104517.599.05
5250.035103017.598.75
600.0357453072.02
7500.035104517.599.34
8500.02074517.599.48
900.05074517.579.25
10250.02044517.561.82
11250.0507453099.05
12250.05076017.599.16
13250.05073017.598.91
14250.03574517.597.95
15500.035745596.37
16250.020104517.598.19
17250.035106017.599.21
18250.035760598.55
19250.03574517.597.95
20500.0357453099.49
21500.03576017.599.37
2200.03573017.561.82
23250.0351045599.58
24250.05044517.598.51
25250.03546017.598.18
2600.03544517.575.88
27250.020745586.14
28250.02073017.597.91
29250.0357603097.40
30250.050745599.52
3100.035104517.575.88
32250.0357303097.97
33500.03573017.599.34
34250.0354453099.41
3500.035745579.25
36250.03543017.599.41
37250.035445598.44
38250.0207453083.76
39250.03574517.598.22
40500.03544517.595.86
41250.03574517.597.54
42250.03574517.598.22
43250.035730596.40
44500.05074517.599.40
4500.02074517.579.25
46250.03510453098.75
Table 4. The kinetic parameters of pseudo-first-order (PFO) and pseudo-second-order (PSO) for crystal violet (CV) sorption onto Kaol/Cel–25 at optimal conditions.
Table 4. The kinetic parameters of pseudo-first-order (PFO) and pseudo-second-order (PSO) for crystal violet (CV) sorption onto Kaol/Cel–25 at optimal conditions.
Concentration
(mg/L)
qe,exp (mg/g)PFOPSO
qe,cal (mg/g)k1 (1/min)R2qe,cal (mg/g)k2 10−2
(g/mg. min)
R2
5051.7454.230.04590.96359.520.0740.994
100131.5130.530.05310.955133.330.2980.999
150195.6180.330.02770.994198.330.0421
200203.2182.660.02150.945208.560.0160.997
250252.1192.310.03180.938294.120.0140.995
300297.7274.820.04340.954299.580.0200.996
Table 5. Langmuir, Freundlich, and Temkin constants for the adsorption of crystal violet (CV) dye onto Kaol/Cel–25 at 45 °C (318.15 K).
Table 5. Langmuir, Freundlich, and Temkin constants for the adsorption of crystal violet (CV) dye onto Kaol/Cel–25 at 45 °C (318.15 K).
ModelParametersValue
Langmuirqmax (mg/g)294.12
Ka (L/mg)0.03
R20.99
FreundlichKf (mg/g) (L/mg)1/n38.37
n2.43
R20.98
TemkinKT (L/mg)0.07
bT (JHZ[J/mol])22.40
R20.91
Table 6. Comparison of the qmax (mg/g) values for crystal violet (CV) dye adsorption to that of various adsorbents.
Table 6. Comparison of the qmax (mg/g) values for crystal violet (CV) dye adsorption to that of various adsorbents.
Adsorbentsqmax (mg/g)References
Palm kernel fiber78.9[52]
Rice husk NaOH-modified44.87[53]
Fly ash74.6[54]
Cellulose-based from sugercane bagasse107.5[55]
Magnetite graphene oxide-doped super adsorbent hydrogel 88.78[56]
Rubber seed pericarp treated with sulfuric acid302.7[57]
Microalgae243.0[58]
Zeolite–montmorillonite150.52[59]
Durian seeds powder158[60]
Kaol/Cel–25294.12This study
Table 7. Parameters thermodynamic of crystal violet (CV) dye adsorption onto Kaol/Cel–25.
Table 7. Parameters thermodynamic of crystal violet (CV) dye adsorption onto Kaol/Cel–25.
T (K)LnkdΔ (kJ/mol)Δ (kJ/mol)Δ (kJ/mol K)
303.150.0867−0.22−78.65−0.238
313.150.5931−1.54
318.151.5411−4.07
333.152.9099−8.06
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

Mecheri, R.; Zobeidi, A.; Atia, S.; Neghmouche Nacer, S.; Salih, A.A.M.; Benaissa, M.; Ghernaout, D.; Arni, S.A.; Ghareba, S.; Elboughdiri, N. Modeling and Optimizing the Crystal Violet Dye Adsorption on Kaolinite Mixed with Cellulose Waste Red Bean Peels: Insights into the Kinetic, Isothermal, Thermodynamic, and Mechanistic Study. Materials 2023, 16, 4082. https://doi.org/10.3390/ma16114082

AMA Style

Mecheri R, Zobeidi A, Atia S, Neghmouche Nacer S, Salih AAM, Benaissa M, Ghernaout D, Arni SA, Ghareba S, Elboughdiri N. Modeling and Optimizing the Crystal Violet Dye Adsorption on Kaolinite Mixed with Cellulose Waste Red Bean Peels: Insights into the Kinetic, Isothermal, Thermodynamic, and Mechanistic Study. Materials. 2023; 16(11):4082. https://doi.org/10.3390/ma16114082

Chicago/Turabian Style

Mecheri, Razika, Ammar Zobeidi, Salem Atia, Salah Neghmouche Nacer, Alsamani A. M. Salih, Mhamed Benaissa, Djamel Ghernaout, Saleh Al Arni, Saad Ghareba, and Noureddine Elboughdiri. 2023. "Modeling and Optimizing the Crystal Violet Dye Adsorption on Kaolinite Mixed with Cellulose Waste Red Bean Peels: Insights into the Kinetic, Isothermal, Thermodynamic, and Mechanistic Study" Materials 16, no. 11: 4082. https://doi.org/10.3390/ma16114082

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