Next Article in Journal
Increasing the Thermostability of Luciferase from Antarctic Krill by Rational Design for Biotechnological Applications
Previous Article in Journal
Games with a Purpose for Part-of-Speech Tagging and the Impact of the Applied Game Design Elements on Player Enjoyment and Games with a Purpose Preference
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel ZIF-67@Bentonite (ZIF-67@BNT) Nanocomposite for Aqueous Methyl Orange Removal

by
Kashif Faheem
1,
Sagheer A. Onaizi
2,3 and
Muhammad S. Vohra
1,4,*
1
Civil and Environmental Engineering Department, King Fahd University of Petroleum & Minerals (KFUPM), Dhahran 31261, Saudi Arabia
2
Chemical Engineering Department, King Fahd University of Petroleum & Minerals (KFUPM), Dhahran 31261, Saudi Arabia
3
Interdisciplinary Research Center for Hydrogen and Energy Storage, King Fahd University of Petroleum & Minerals (KFUPM), Dhahran 31261, Saudi Arabia
4
Interdisciplinary Research Center for Construction and Building Materials, King Fahd University of Petroleum & Minerals (KFUPM), Dhahran 31261, Saudi Arabia
*
Author to whom correspondence should be addressed.
Appl. Sci. 2025, 15(7), 3562; https://doi.org/10.3390/app15073562
Submission received: 20 February 2025 / Revised: 18 March 2025 / Accepted: 19 March 2025 / Published: 25 March 2025

Abstract

:

Featured Application

1. Developed highly efficient novel ZIF-67@BNT nanocomposite adsorbent. 2. The ZIF-67@BNT showed high MO adsorption capacity. 3. Advanced RSM and ML modeling were successfully used for process optimization.

Abstract

The indiscriminate discharge of synthetic dyes into wastewater streams poses a severe threat to the environment as well as to human well-being. Among all these dyes, methyl orange (MO) attracts attention due to its widespread use and persistence in industrial effluents. This study investigated the use of zeolitic imidazolate framework and bentonite (ZIF-67@BNT) nanocomposite material for the removal of MO from the aqueous phase. Various characterization techniques were employed such as FTIR, XRD, and TGA to verify the successful synthesis of the ZIF-67@BNT adsorbent, which was subsequently utilized to investigate the adsorption of MO. Batch adsorption studies demonstrated a high MO adsorption capacity of 187 mg/g. A response surface methodology (RSM)-based modeling exercise was used to optimize the adsorption process. While assessing the impact of various operational factors, the initial MO concentration followed by ZIF-67@BNT dose were noted to be important. Adsorption kinetics and isotherm studies were also completed. The ZIF-67@BNT nanocomposite after adsorption analysis indicated multiple mechanisms facilitating MO uptake. Additionally, various machine learning (ML) models such ANN, SVR, RF, and GPR were also utilized to predict MO adsorption onto ZIF-67@BNT nanocomposite under a varying set of conditions.

1. Introduction

Human activities with an ever-increasing introduction of pollutants continue to diminish the availability of clean water [1,2]. Such aquatic pollution results from a range of inorganic and organic contaminants including heavy metals and dyes [2,3,4,5,6,7,8,9,10,11,12,13,14,15,16]. For example, the treatment of wastewater with dyes has been looked into to reduce the respective environmental concerns [2]. One such dye is methyl orange (MO). Its chemical structure is provided in Figure 1 and its IUPAC name is sodium-4-(4-dimethylamino phenyl diazenyl) benzenesulfonate [6]. MO dye is also noted in several industrial discharges, thus prompting the need for an effective treatment for its removal from wastewater streams. Considering this, several techniques have been used for MO removal from wastewater, including photocatalysis and adsorption processes [8,13,14,15,16].
Further, the adsorption capacity of a given adsorbent is contingent upon its inherent characteristics such as the specific surface area and surface functional groups [10]. Recently, there has been a surge of interest in novel materials known as metal organic frameworks (MOFs). MOFs have various types including zeolite imidazole framework-67 (ZIF-67). It comprises cobalt ions coordinated with imidazole ligands. This porous material is employed as an adsorbent for various treatments including wastewater treatment, and it is also used as catalyst [11,12]. Researchers have shown that different ZIFs can adsorb dyes, including methylene blue (MB), methyl orange (MO), and rhodamine B (RB) [13,14]. In one study, ZIF-67 with layered double hydroxide (ZIF-67@LDH) showed higher MO and MB removal compared to LDH-only systems [15]. Furthermore, several natural clay minerals have also been used for removing contaminants from the water phase [16] and the use of clay surface modification techniques to improve pollutant removal has also been reported [17,18,19]. Additionally, the response surface methodology (RSM) and artificial intelligence (AI)-based techniques have also been used for process optimization and environmental modeling [5,20]. Further, various ZIF-67-based materials have been explored for the removal of different pollutants and adsorption studies [21,22,23,24,25,26,27,28,29,30] and the aforementioned clay [17,18,19,20] materials’ application for water decontamination show promising results that can be further explored for the application of nanocomposites of ZIFs with clay material. With this impetus, this study looked into the application of novel ZIF-67@BNT for aqueous methyl orange adsorption, which offers a viable solution for the treatment of aqueous waste streams. In summary, the aforementioned applications of ZIFs and clay-based materials for water decontamination applications support that both materials can be used as adsorbents for water decontamination. Furthermore, the literature review also indicates that the nanocomposite ZIF-67@BNT has not been used for the removal of MO from aqueous streams. Considering this, the present work investigated the use of ZIF-67@BNT-based adsorption for the removal of aqueous phase MO under a varying set of process conditions. Additionally, RSM and AI-based modeling were also employed for process optimization purposes, and the respective details are provided in the sections below.

2. Materials and Methods

2.1. Materials and Equipment

This investigation used pure bentonite (BNT) (Fisher, Waltham, MA, USA) and other chemicals of high purity, including ammonium hydroxide (NH4OH), 2-methyl imidazole (2-MIM), cobalt (II) nitrate hexahydrate (Co(NO3)2∙6H2O), and methyl orange [(CH3)2NC6H4N6H4SO3Na)]. All respective stock solutions were prepared using de-ionized water.
A Rigaku Miniflex-II diffractometer was used for X-ray diffraction (XRD) analysis to evaluate the crystalline structure of the synthesized ZIF-67@BNT. The respective materials were scanned at a rate of 7° per minute throughout a 2-theta angle range of 5° to 40°. Further, thermogravimetric analysis (TGA) was conducted using TA Instruments to evaluate the thermal stability of the synthesized nanocomposite within a temperature range of 25 to 800 °C under an inert environment (i.e., in nitrogen) at a heating rate of 10 °C per minute. Also, ZIF-67@BNT surface functional groups were determined by employing a Fourier transform infrared spectroscopy (FTIR) setup (Smart iTR NICOLET iS10). Furthermore, a Shimadzu UV-Vis spectrophotometer setup was used to determine the residual methyl orange (MO) concentration in aqueous samples (that had been duly pre-filtered using 0.2 µm (Whatman, Dassel, Germany) filters to exclude the adsorbent particles) at 464 nm. Standards for MO ranged from 12.5 to 100 mg/L and the same were used to develop the calibration curve. Further details on the synthesis and application of ZIF-67@BNT for the adsorptive removal of MO from wastewater samples are provided in the sections below and also summarized in Figure 2.

2.2. Synthesis of ZIF-67 and ZIF-67@BNT

The ZIF-67@BNT nanocomposite synthesis procedure is sketched in Figure 3. First, a cobalt nitrate salt aqueous solution was prepared by dissolving 11.64 g of Co(NO3)2∙6H2O in 150 mL de-ionized water. To this solution, 0.24 g BNT was added, followed by stirring at room temperature for 10 min. To thoroughly disperse BNT in the cobalt solution, the mixture was sonicated at 50% amplitude for 30 min, with intervals of 30 s of pulse on and 3 s off. Meanwhile, in a separate beaker, 13.12 g of 2-methyl imidazole was mixed with 150 mL of 28% NH4OH aqueous solution, maintaining a metal-to-linker ratio (i.e., Co:2-MIM) of 1:4. This was followed by pouring the imidazole-containing solution quickly into the bentonite–cobalt solution along with stirring the resulting liquid mixture continuously for 2 h. The produced ZIF-67@BNT nanocomposite was, then, recovered and purified through washing 5 times with de-ionized water. The ZIF-67@BNT nanocomposite was recovered after each washing cycle using 20 min of centrifugation at 6000× g rpm. The purified ZIF-67@BNT nanocomposite was dried at 50 °C.

2.3. Response Surface Methodology (RSM)

As shown in Equation (1), the response surface methodology (RSM)-based design of experiments and modeling facilitates the connections between input variables/factors (XiS) and the desired output/response (y) [31,32].
y = f(X1, X2, X3, X4, …, Xn)
Further, in RSM, the face-centered central composite design (FCCD) enables a comprehensive examination of all related system operation variables/factors in the desired response [33]. For the present MO adsorption work, the ZIF-67@BNT dosage, initial MO concentration, and contact time were the independent factors, whereas the ZIF-67@BNT MO adsorption capacity (mg/g) was the desired response. Table 1 lists those factors and the respective levels. The statistical parameters to assess the RSM model validity included coefficient of determination (R2), adjusted R2, and projected R2 along with the analysis of variance (ANOVA). The RSM modeling was completed employing Design-Expert software (version 13).

2.4. Batch Adsorption Experiment

The adsorption process is a complex sequence of events that begins with external diffusion, where the adsorbate migrates across the liquid film to the surface of the adsorbent. The final stage of the adsorption process involves the interaction between the adsorbate and the active sites of the adsorbent, leading to adsorption equilibrium. For this study, each batch adsorption test a 150 mL MO solution was mixed with the desired ZIF-67@BNT adsorbent dose followed by placing it on a Bioevopeak mechanical shaker. All suspension samples were filtered using a 0.2 µm filter (Whatman, Dassel, Germany), followed by MO analysis and Equations (2) and (3) were then used to calculate the specific removal efficiency (%) and the adsorption capacity (mg/g), respectively:
R = C 0 C e C e × 100
Q e = C o C e m × V
where the ‘Co’ is the initial MO concentration (mg/L), ‘Ce’ is the MO concentration after adsorption at equilibrium, ‘m’ is the mass of the adsorbent in grams, and ‘V’ is the sample volume in liters. For the adsorption kinetics, the ZIF-67@BNT dose was 5 mg at an initial MO concentration of 100 ppm. For the adsorption isotherm tests, the starting initial MO concentration was varied from 5 to 125 mg/L at a ZIF-67@BNT dose of 5 mg.

2.5. Machine Learning (ML) and Artificial Neural Network (ANN) Modeling

Machine learning techniques are increasingly being used to forecast adsorbents to remove pollutants from aqueous streams [34,35]. However, due to operational and financial difficulties, there is still a lack of experimental data that support the application of machine learning and other artificial intelligence techniques in wastewater treatment. In order to assess the prediction capabilities of several machine learning algorithms, this study collected experimental data on batch adsorption. The machine learning (ML) modeling technique has been explored for adsorption work [36,37].

2.5.1. Artificial Neural Network

The ANN is another widely employed artificial intelligence (AI)-based modeling technique [38]. As shown in Figure 4A, the input layer, output layer, and one or more hidden layers make up the ANN structure and multiple neurons grouped as linked nodes make up each layer, which are trained as per the multi-layer perceptron (MLP) along with the experimentally acquired data and the final feedforward backpropagation (FFBP) calculated findings are used in order to improve model performance. Equation (4) displays the anticipated ANN output:
Y = f(W + B)
where Y stands for the output, B for the bias, W for the weight, and f for the activation function.

2.5.2. Support Vector Regression

Support vector regression (SVR) is an extension of support vector machines (SVMs) originally designed for classification but adapted for regression tasks. It maps input vectors into a high-dimensional feature space using a hyperplane to maximize the separation between data points, leading to accurate predictions. Various kernel functions can be used for this mapping, making SVR suitable for a wide range of applications, even with small datasets. Equation (5) produces the output and Figure 4B illustrates SVR operations [35,39].
y = i = 1 n α I k ( x i ,   x j ) + b
where y is predicted output, n is number of support vectors, α is Lagrange multipliers, k is kernel function, b is bias term, and xi and xj are the support vectors and test inputs.

2.5.3. Random Forest

Further, the random forest (RF) is a tree-based ensemble learning model that improves upon the bagging tree approach. It constructs multiple decision trees using bootstrap sampling during training. Each decision tree consists of internal nodes (which assess predictions), branch nodes (which display evaluation results), and leaf nodes (which show class labels in regression). The RF model builds regression trees on bootstrap samples, selects attributes randomly at each branch node, and stops tree expansion once the mean squared error reaches an optimal level. Finally, the RF model aggregates predictions from multiple trees, averaging them to produce the final result. The RF model formulation is shown in Equation (6) [38].
f ^ r f N x = 1 N i = 1 N t i ( x )
where ti(x) is the individual regression tree algorithm, N is the number of regression algorithms, and f r f N (x) is the included regression tree. Figure 4C shows the RF method’s graphical display.

2.5.4. Gaussian Process

The Gaussian process (GP) is a non-parametric probabilistic model used in machine learning for regression, classification, and uncertainty quantification. It models a collection of random variables with a joint Gaussian distribution, allowing for flexible data representation while providing predictions with uncertainty estimates. The key characteristics of the GP include handling non-parametric data, probabilistic predictions, interpolation, smoothing, and hyperparameter marginalization and the model is evaluated using N-fold cross-validation, correlation coefficient (R), coefficient of determination (R2), and RMSE. Figure 4D shows the visual presentation of the GP [40].

3. Results and Discussion

3.1. ZIF-67@BNT Characterization

3.1.1. X-Ray Diffraction Analysis

The X-ray diffraction (XRD) patterns of the ZIF-67@BNT nanocomposite, ZIF-67, and BNT are shown in Figure 5A. The diffraction pattern of BNT (shown in blue) shows some characteristic peaks related to the clay mineral structure. Major reflections seen at 2θ: 6.1° (001), 19.8° (110), 26.7° (020), 35.0° (220), 54.9° (240), and 68.1° (330) correspond to different planes of crystals existing in BNT [41] The basal reflection determined close to 6.1° (001) is an indicator of the montmorillonite interlayer spacing, which is a fundamental structural property of BNT. Some peaks established near 20–30° correspond to quartz impurities predominating in natural BNT, while their broad character signifies an amorphous phase of BNT owing to its hydrated layered silicate structure. The calculated crystallite size of BNT is 2810.19 nm; therefore, it shows structural features in relatively large particle dimensions and a less defined crystalline nature than ZIF-67.
The X-ray diffraction (XRD) pattern of ZIF-67 (red pattern) can be assigned to characteristic peaks at 2θ = 7.3°, 10.4°, 12.7°, 14.7°, 16.4°, 18.1°, 22.2°, 24.5°, 29.4°, and 33.5°, typical of the crystalline metal–organic framework (MOF) structure of ZIF-67 [42]. They are assigned as (011), (002), (013), (112), (222), (233), (044), (134), (244), and (235) crystal planes, respectively [43]. The calculated crystallite size of ZIF-67 and ZIF-67@BNT is 44.48 and 79.28 nm, respectively, with the ZIF-67@BNT nanocomposite being significantly larger than ZIF-67 alone, suggesting a structural reorganization upon composite formation. The decline in crystallinity in the composite might potentially point to surface changes and flaws, which would improve the adsorption capacity by raising the number of accessible active sites [44]. Furthermore, BNT in the composite is projected to increase the general structural stability of ZIF-67, hence avoiding its collapse in aqueous surroundings [45].
With all factors taken into account, the XRD study verifies the effective synthesis of the ZIF-67@BNT composite by maintaining the main structural characteristics of ZIF-67 and including modifications caused by interactions with BNT. The noted rise in crystallite size points to BNT’s stabilizing action for ZIF-67, hence enabling better textural qualities. Moreover, the retention of unique ZIF-67 crystal planes in the composite shows that its crystalline structure stays intact; on the other hand, the addition of BNT brings structural changes advantageous for adsorption uses. The minimal shifting and widening of BNT’s peaks point to exfoliation or intercalation, which improves ZIF-67 dispersion in the composite [46].

3.1.2. Fourier Transform Infrared Spectroscopy Analysis

Besides the crystallinity, Fourier transform infrared spectroscopy (FTIR) results of BNT, ZIF-67, and the ZIF-67@BNT composite are shown in Figure 5B, providing insight into the functional groups and interactions present between the components. BNT displays a wide O-H stretching band at 3100 cm−1 due to adsorbed water, as well as peaks at 2370 cm−1 (C-H stretching), 1408 cm−1 (C=N stretching), 1170 cm−1 (C=C stretching), and 970 cm−1 (Si-OH bending), which are a consequence of its layered silicate structure. These peaks indicate the presence of montmorillonite and silicate groups, responsible for the BNT’s adsorption capacity [47,48].
The FTIR spectrum of ZIF-67 exhibits some characteristic vibrational modes related to its MOF structure. The Co-N stretching vibration coincides with a clear peak at 538 cm⁻¹, therefore verifying the coordination between cobalt ions and the 2-methylimidazolate linker. Strong O-H stretching vibrations imply the hydrophobic character of ZIF-67, which is anticipated from the non-polar imidazolate linkers. Since C=N and C=C stretching modes compose the fundamental structure of the MOF, their presence promotes the stability of the ZIF-67 framework even more. Crucially for preserving adsorption efficiency, the well-defined peaks of ZIF-67 point to a highly crystalline and stable framework [49].
In the ZIF-67@BNT, significant spectrum alterations noticed in the ZIF-67@BNT composite indicate robust interactions between ZIF-67 and BNT. N-H stretching at 3232 cm−1 points to interactions between the hydroxyl (-OH) groups of BNTs and the imidazolate linkers of ZIF-67. New C-H stretching vibrations at 2738 cm−1 (aromatic) and 2600 cm−1 (aliphatic) point to changes in the organic framework of the MOF, maybe resulting from surface alterations following integration with BNT or electrostatic interactions. Retention of the Co-N peak at 538 cm−1 guarantees composite structural integrity of ZIF-67 [41,50].
Due to the high interfacial interactions between the two adsorbent materials, ZIF-67@BNT exhibits additional distinctive peaks when compared to ZIF-67 and BNT. Electrostatic interactions between the Co2+ metal centers in the ZIF-67 framework and the negatively charged BNT surface cause structural modifications, while hydrogen bonding between the hydroxyl groups of BNT and the imidazolate linkers of ZIF-67 causes changes in vibrational frequencies. Spectral fluctuations may also be caused by the emergence of new functional groups or changes to the electrical environment of existing groups brought about by ZIF-67’s integration into the BNT matrix. Alterations to the peak intensities and movement of the peaks point to altered adsorption sites, which may improve the composite’s efficacy in pollutant removal [51,52,53].

3.1.3. Thermogravimetric Analysis

The thermogravimetric analysis (TGA) of the ZIF-67@BNT nanocomposite as shown in Figure 6 reveals its thermal stability profile. The initial segment of the TGA curve, up to approximately 150 °C, shows minimal weight loss, likely due to the evaporation of surface-adsorbed moisture or other volatile components. However, a pronounced weight loss occurs between 150 and 600 °C, with the primary decomposition phase observed from around 300 °C to 500 °C. This substantial weight reduction is attributed to the degradation of the organic framework within ZIF-67 and possibly the release of structurally bound water associated with the BNT component. Beyond 600 °C, a slower rate of weight loss is noted, indicating the breakdown of more thermally resilient components within the nanocomposite, with stabilization occurring at around 800 °C [54].
Hence, the characterization results given above, using XRD, FTIR, and TGA, regarding the structural, chemical, and thermal characterizations of the ZIF-67@BNT composite, confirm its successful synthesis. They also demonstrate that the ZIF-67 composite retains the crystalline integrity of ZIF-67, as it undergoes modifications because of its interactions with BNT and shows enhanced thermal stability. These characteristics make the composite a good material for uses in environmental remediation and adsorption systems.

3.2. Benchmarking the Performance of the ZIF-67@BNT Nanocomposite with Its Constituents

The results depicted in Figure 7 contrast the MO adsorption performance of ZIF-67@BNT nanocomposite with its parental materials (i.e., ZIF-67 and BNT). As shown in Figure 7, ZIF-67 exhibits an MO adsorption capacity of 222.9 mg/g. The ZIF-67@BNT nanocomposite, on the other hand, shows the highest MO adsorption capacity with a value of 275.3 mg/g. The enhanced performance of ZIF-67@BNT can be attributed to a synergistic effect between ZIF-67 and BNT, where the integration of BNT with ZIF-67 results in improved structural stability, increased active sites, or optimized interaction with the targeted pollutant (MO in this case). This superior capacity highlights the potential of nanocomposite materials to outperform their individual components by leveraging the strengths of their constituents. The superiority of nanocomposites generated from different materials has also been demonstrated in the literature [10,32,55,56,57,58]. Such superiority underscores the importance of material hybridization in developing advanced materials with enhanced properties, paving the way for further exploration and optimization in the wastewater treatment field.
Based on the results presented in Figure 7, the ZIF-67@BNT nanocomposite offers a greater performance, making it a promising candidate for applications in water decoloring such as the MO removal reported herein. Thus, subsequent studies will focus on optimizing the performance of ZIF-67@BNT nanocomposite using RSM techniques, studying the adsorption kinetics and adsorption isotherm of MO onto this nanocomposite, and developing predictive machine learning models. Plausible adsorption mechanism(s) will also be highlighted wherever applicable throughout the discussion in the subsequent sections.

3.3. Analysis of Methyl Orange Adsorption Using ZIF-67@BNT

3.3.1. Experimental Design and Analysis of Variance

Results from the MO adsorption experiments conducted using the ZIF-67@BNT nanocomposite are given in Table 2. The adsorption findings show that the uptake of MO by ZIF-67@BNT is a function of the process parameters. Also, the MO adsorption capacity using ZIF-67@BNT ranges from 38 to 187 mg/g, indicating that the MO removal can be significantly enhanced and optimized by varying the process conditions. Furthermore, among the various RSM-based models generated, the quadratic model fitted well to the noted MO adsorption results using ZIF-67@BNT, as given below (Equation (7)):
Adsorption capacity (mg/g) = 2.57351 + 0.185881A + 0.057277B + 0.115436C − 0.002436AB − 0.009655A2
The R2, adjusted R2, and predicted R2 values for the present RSM model were noted to be 0.953, 0.927, and 0.822, respectively. The quadratic model’s (Equation (7)) suitability is thus demonstrated by less than 20% difference between the adjusted and predicted R2 values [59]. This is also supported by the actual and predicted MO removal as given in Figure 8a. Further, some other statistical parameters for the suggested RSM model (Equation (7)) are also provided in Table 3. The ANOVA findings as shown in Table 3, with a p-value of 0.05 or less, also support the suggested RSM-based model for MO adsorption by ZIF-67@BNT. Furthermore, the Box–Cox plot for total MO removal (Figure 8b), Figure 8c that illustrates that the residuals have a normal distribution, and Figure 8d that reveals an absence of correlation between the run sequence and residuals (as shown by the random dispersion of points in the plot) also support the suggested RSM-based model’s (Equation (7)) suitability to predict the MO uptake by the novel ZIF-67@BNT under a varying set of process conditions.

3.3.2. Examine the Effect of Process Parameters Using ANOVA

Additionally, the process variables could have a significant effect on the adsorption capacity [60] and Figure 9 illustrates the effect of the present process variables on MO adsorption capacity. To that end, it is noted that an increase in the MO concentration from 10 to 30 mg/L at an adsorbent dose of 5 mg increases the MO adsorption capacity. Such an enhancement in the MO uptake by ZIF-67@BNT is suggested to result from an enhanced driving force for the mass transfer of MO from the bulk aqueous phase to the bulk adsorbent surface. Figure 9 also shows that extended contact time delivers better MO uptake, as more process time permits more adsorbate molecules to diffuse to the adsorbent surface sites until reaching the adsorption equilibrium. Similarly, Figure 10 and Figure 11 also indicate that the MO adsorption capacity (Qmax) is affected by both the adsorbent ZIF-67@BNT dose and the methyl orange (MO) initial concentration; at higher MO concentrations (30 ppm) an increase in the ZIF-67@BNT adsorbent dose equilibrates more MO molecules on the adsorbent surface and significantly increases the Qmax [61]. Hence, these findings can be used to establish parameters for the enhancement of MO uptake by ZIF-67@BNT. Furthermore, this study also provides a comparison with the other adsorption studies on the removal of MO from aqueous solution as shown in Table 4.

3.3.3. Adsorption Kinetics and Isotherm Analysis

The pseudo-first order (PFO), pseudo-second order (PSO), and Avrami models were also employed for the kinetic modeling. Out of these three kinetic models, the Avrami model showed the highest R2 value of 0.9949 (Figure 12 and Table 5). This suggests that MO adsorption by ZIF-67@BNT involves a potential variation in the adsorption mechanism with the Avrami model indicating a non-linear kinetics and a better capturing of the overall MO adsorption trend as a function of time.
The interactions between the MO adsorbate and ZIF-67@BNT adsorbent were further evaluated using the adsorption isotherm approach. To that end, the respective MO adsorption isotherm results were fitted to the Langmuir isotherm, Freundlich isotherm, and Redlich–Peterson isotherm and the output was analyzed for possible adsorbate–adsorbent interactions as described earlier [67]. The respective experimental and modeling findings as given in Figure 13 and Table 6 indicate that the three-parameter Redlich–Peterson (R–P) isotherm shows a good match to the experimental MO adsorption isotherm results, which is also confirmed by the “g” parameter value of more than 1 [68]. This finding indicates that, instead of a plateau, the adsorption capacity would tend to increase with the mass transfer driving force, i.e., the MO aqueous phase concentration, though the change at higher equilibrium MO concentrations is low.
This also indicates a complex adsorption mechanism with several adsorbent surface parameters (including the type of surface functional group) initiating the MO adsorption. This topic is probed further.

3.3.4. Post-Analysis of MO Adsorption on ZIF-67@BNT Nanocomposite

Fourier transform infrared (FTIR) spectra of ZIF-67@BNT before and after methyl orange adsorption were analyzed to investigate the changes in functional groups and interactions involved in the adsorption process. The findings as shown in Figure 14 exhibit the effects of methyl orange adsorption on the nanocomposite’s functional groups. The spectrum before adsorption exhibited a broad peak around 3400 cm−1, corresponding to the stretching vibrations of hydroxyl (-OH) groups, which are typically associated with adsorbed water molecules and surface hydroxyl groups in BNT [69]. Peaks observed from 2900–3100 cm−1 can be attributed to C-H stretching vibrations from organic components in the ZIF-67 framework [58,70], while the band around 1600–1700 cm−1 corresponds to C = O stretching or H-O-H bending vibrations [71]. Additionally, strong peaks in the range of 1000–1500 cm−1 are characteristic of metal–oxygen (Co-O, Al-O-Si) stretching and bending vibrations from the BNT structure and ZIF-67 framework. The region below 1000 cm⁻¹ represents metal–ligand interactions, specifically Co-N stretching in ZIF-67 [72,73]. After methyl orange adsorption, notable spectral changes were observed, confirming the interaction between the adsorbent and the dye molecules. A shift and reduction in the intensity of the hydroxyl (-OH) peak at ~3400 cm−1 suggest possible hydrogen bonding between the dye and surface hydroxyl groups of ZIF-67@BNT [74]. The emergence of new peaks and enhanced intensities in the 1500–1650 cm−1 region correspond to the aromatic C=C stretching and sulfonate (-SO3−1) vibrations of methyl orange, indicating its successful attachment to the surface of ZIF-67@BNT [75]. Furthermore, broadening and shifts in the 1000–1200 cm−1 region suggest interactions between the BNT structure and methyl orange functional groups, likely through electrostatic attractions or coordination interactions with Co2+ sites in ZIF-67. Additionally, changes in the fingerprint region (below 1000 cm−1) indicate modifications in the metal–ligand framework due to adsorption. The observed spectral changes suggest that the adsorption of methyl orange onto ZIF-67@BNT occurs through multiple mechanisms, including electrostatic interactions between the negatively charged sulfonate (-SO3⁻) groups of methyl orange and the positively charged Co2+ or Al3+ sites in ZIF-67@BNT and hydrogen bonding between hydroxyl groups of BNT and methyl orange molecules [76].
These interactions validate the effectiveness of ZIF-67@BNT as an adsorbent for methyl orange, highlighting its potential utility in applications aimed at removing organic dyes from contaminated water. Furthermore, the noted stability and adsorption efficiency support the nanocomposite’s reusability and robustness, making it a promising candidate for practical environmental remediation efforts [75].

3.4. Predicting MO Adsorption Capacity with Machine Learning Techniques

In this study, we also used ML modeling using the four-regression model, namely GPR, SVR, ANN-based MLP, and RF, to optimize and analyze the obtained adsorption data and the descriptive analysis shown in Table 7. The MLP Regressor model with two hidden layers (65 and 42 neurons) and the ReLU activation function effectively captures non-linear patterns in regression tasks. Using the L-BFGS solver ensures faster convergence, while the adaptive learning rate optimizes performance dynamically. Further improvements can be made by fine-tuning hyperparameters, exploring alternative activation functions, and integrating feature selection techniques. The RF model used 100 estimators, a minimum of three samples for splitting nodes, a squared error threshold, and a maximum depth of six. This collection of parameter combinations was chosen to maximize the model’s performance while balancing complexity and general applicability of results. SVR used a 5-degree polynomial kernel with an auto gamma and regularization parameter (C = 1.0). The GPR was set with a given kernel with 20 optimization restarts, a learning rate of 0.014 for regularization, and target value normalization enabled.
Metrics including coefficient of determination (R2), root mean square error (RMSE), and mean absolute error (MAE) for both training and testing were analyzed and compared in order to assess the model fit, as seen in Table 8. The actual and predicted outcomes showed a very significant connection with the enhanced models, indicating R2 values for training and testing as follows: GPR achieved 0.996 and 0.962, RF reached 0.955 and 0.927, while SVR had 0.998 and 0.907, and finally ANN had 0.999 and 0.872, respectively. With the highest R2 and the lowest RMSE and MAE scores, as demonstrated in Table 8, the GPR model outperforms the other three models on testing datasets, demonstrating its superior accuracy in forecasting the MO adsorption capacity of the synthesized ZIF-67@BNT. The validation observations validate that the models generalize well to new data and underline the effectiveness of the measures taken to avoid data leaking. This suggests that, in comparison to the ANN, RF, and SVR models, the GPR model offers the greatest fit, the most precise forecasts, and the most efficiency in outcome prediction. Figure 15 provides a visual comparison of the performance of each model.
A Taylor diagram was used in this work for both training and testing datasets in order to identify the best model and assess its dependability. The standard deviation, root mean square error (RMSE), and correlation coefficient of the prediction models are all graphically depicted in the Taylor diagram (Figure 16). The GPR model (represented by the violet circle) is positioned closest to the observed data point in the Taylor diagram displayed in Figure 16, whereas the ANN model exhibits overfitting with high accuracy for the training dataset and comparatively moderate accuracy for the testing dataset. These assessment techniques provided convincing proof that the GPR model is the best option for adsorption capacity prediction, outperforming the independent MLP-based ANN, RF, and SVR models.
The results suggest that machine learning algorithms offer useful insights into adsorbent efficiency for contaminant removal in water systems. Increasing the dataset used for training these models may further improve their accuracy in predicting adsorbent performance for water purification applications.

4. Conclusions

This study examined the removal of MO dye from polluted water through adsorption using a ZIF-67@Bentonite nanocomposite (ZIF-67@BNT). The synthesized ZIF-67@BNT was characterized using Fourier transform infrared spectroscopy (FTIR), X-ray diffraction (XRD), and thermogravimetric analysis (TGA). Furthermore, the batch adsorption studies demonstrated ZIF-67@BNT has a high MO adsorption capacity of 187 mg/g at an initial concentration of 30 mg/L, 5 mg dosage, and a 5 h duration. The response surface methodology (RSM)-based modeling exercise was used to optimize the adsorption process. While assessing the impact of various operational factors, the initial MO concentration, followed by ZIF-67@BNT dose, was noted to be important. Adsorption kinetics and isotherm studies were also completed. The adsorption kinetics results were best described by the Avrami model, whereas the MO adsorption isotherm data fitted well to the Redlich–Peterson (R-P) model. The ZIF-67@BNT nanocomposite ‘after adsorption analysis’ indicated multiple mechanisms facilitating MO uptake onto ZIF-67@BNT, including electrostatic and π–π interactions. Additionally, various machine learning (ML) models, such ANN, SVR, RF, and GPR, were also utilized to predict MO adsorption onto ZIF-67@BNT nanocomposite under a varying set of conditions. In conclusion, the novel ZIF-67@BNT adsorbent can effectively treat MO-contaminated water, and both the RSM- and ML-based modeling approaches yielded good optimization results.

Author Contributions

Conceptualization, S.A.O. and M.S.V.; Methodology, K.F.; Software, K.F., S.A.O. and M.S.V.; Validation, K.F., S.A.O. and M.S.V.; Formal analysis, K.F., S.A.O. and M.S.V.; Investigation, K.F., S.A.O. and M.S.V.; Resources, S.A.O. and M.S.V.; Data curation, K.F., S.A.O. and M.S.V.; Writing—original draft, K.F., S.A.O. and M.S.V.; Writing—review & editing, K.F., S.A.O. and M.S.V.; Visualization, K.F., S.A.O. and M.S.V.; Supervision, S.A.O. and M.S.V.; Project administration, M.S.V.; Funding acquisition, M.S.V. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Interdisciplinary Research Center for Construction and Building Materials (IRC-CBM) at the King Fahd University of Petroleum & Minerals (KFUPM) under Research Grant # INCB2420.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Acknowledgments

The authors would like to thank the Civil and Environmental Engineering Department and the Chemical Engineering Department at King Fahd University of Petroleum & Minerals (KFUPM) for providing the lab facilities.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Aniagor, C.O.; Igwegbe, C.A.; Ighalo, J.O.; Oba, S.N. Adsorption of doxycycline from aqueous media: A review. J. Mol. Liq. 2021, 334, 116124. [Google Scholar] [CrossRef]
  2. Hu, H.; Liu, J.; Xu, Z.; Zhang, L.; Cheng, B.; Ho, W. Hierarchical porous Ni/Co-LDH hollow dodecahedron with excellent adsorption property for Congo red and Cr(VI) ions. Appl. Surf. Sci. 2019, 478, 981–990. [Google Scholar] [CrossRef]
  3. Sabbagh, M.; Hussaini, M.; Ismail, U.; Ahmed, H.R.; Al-Suwaiyan, M.; Vohra, M. Novel nafion-palygorskite composite for Pb/Lead treatment. Int. J. Environ. Sci. Technol. 2024, 22, 879–894. [Google Scholar] [CrossRef]
  4. Ismail, U.M.; Vohra, M.S.; Onaizi, S.A. Adsorptive removal of heavy metals from aqueous solutions: Progress of adsorbents development and their effectiveness. Environ. Res. 2024, 251, 118562. [Google Scholar] [CrossRef]
  5. Ibrahim, A.I.; Onaizi, S.A.; Vohra, M.S. Novel CTAB functionalized graphene oxide for selenium removal: Adsorption results and ANN & RSM modeling. Emergent Mater. 2023, 7, 547–564. [Google Scholar] [CrossRef]
  6. Zaggout, F.R.; El-Ashgar, N.M.; Zourab, S.M.; El-Nahhal, I.M.; Motaweh, H. Encapsulation of methyl orange pH-indicator into a sol-gel matrix. Mater. Lett. 2005, 59, 2928–2931. [Google Scholar] [CrossRef]
  7. Faheem, K.; Khan, S.U.; Washeem, M.; Khan, S.U. Energy efficient removal of COD from landfill leachate wastewater using electrocoagulation: Parametric optimization using RSM. Int. J. Environ. Sci. Technol. 2021, 19, 3625–3636. [Google Scholar] [CrossRef]
  8. Anil, K.; Surenjan, A. Green synthesis, characterisation, and performance evaluation of ZnO/TiO2 nanocomposite for cationic and anionic dye removal from aqueous solutions under solar irradiation. Water Pr. Technol. 2024, 19, 824–838. [Google Scholar] [CrossRef]
  9. Haddaji, C.; Rifi, S.K.; Digua, K.; Madinzi, A.; Chatoui, M.; Driouich, A.; Ettaloui, Z.; Kurniawan, T.A.; Anouzla, A.; Souabi, S. Optimization of the coagulation-flocculation process using response surface methodology for wastewater pretreatment generated by vegetable oil refineries: A path towards environmental sustainability. Environ. Nanotechnol. Monit. Manag. 2024, 22, 100973. [Google Scholar] [CrossRef]
  10. Al-Nowaiser, W.K.; Onaizi, S.A.; Vohra, M.S. Synthesis and application of novel GO@ZIF-7 nanocomposite for the removal of lead from aqueous solutions. Emergent Mater. 2023, 7, 1043–1055. [Google Scholar] [CrossRef]
  11. Nazir, M.A.; Ullah, S.; Shahid, M.U.; Hossain, I.; Najam, T.; Ismail, M.A.; Rehman, A.U.; Karim, R.; Shah, S.S.A. Zeolitic imidazolate frameworks (ZIF-8 & ZIF-67): Synthesis and application for wastewater treatment. Sep. Purif. Technol. 2024, 356, 129828. [Google Scholar] [CrossRef]
  12. Zhong, G.; Liu, D.; Zhang, J. The application of ZIF-67 and its derivatives: Adsorption, separation, electrochemistry and catalysts. J. Mater. Chem. A 2017, 6, 1887–1899. [Google Scholar] [CrossRef]
  13. Uddin, M.J.; Ampiaw, R.E.; Lee, W. Adsorptive removal of dyes from wastewater using a metal-organic framework: A review. Chemosphere 2021, 284, 131314. [Google Scholar] [CrossRef] [PubMed]
  14. Ganiyu, S.A.; Suleiman, M.A.; Al-Amrani, W.A.; Usman, A.K.; Onaizi, S.A. Adsorptive removal of organic pollutants from contaminated waters using zeolitic imidazolate framework Composites: A comprehensive and Up-to-date review. Sep. Purif. Technol. 2023, 318, 123765. [Google Scholar] [CrossRef]
  15. Nazir, M.A.; Khan, N.A.; Cheng, C.; Shah, S.S.A.; Najam, T.; Arshad, M.; Sharif, A.; Akhtar, S.; Rehman, A.U. Surface induced growth of ZIF-67 at Co-layered double hydroxide: Removal of methylene blue and methyl orange from water. Appl. Clay Sci. 2020, 190, 105564. [Google Scholar] [CrossRef]
  16. Bahrudin, N.N.; Nawi, M.A.; Jawad, A.H.; Sabar, S. Adsorption Characteristics and Mechanistic Study of Immobilized Chitosan-Montmorillonite Composite for Methyl Orange removal. J. Polym. Environ. 2020, 28, 1901–1913. [Google Scholar] [CrossRef]
  17. Barakan, S.; Aghazadeh, V. The advantages of clay mineral modification methods for enhancing adsorption efficiency in wastewater treatment: A review. Environ. Sci. Pollut. Res. 2020, 28, 2572–2599. [Google Scholar] [CrossRef]
  18. Ewis, D.; Ba-Abbad, M.M.; Benamor, A.; El-Naas, M.H. Adsorption of organic water pollutants by clays and clay minerals composites: A comprehensive review. Appl. Clay Sci. 2022, 229, 106686. [Google Scholar] [CrossRef]
  19. Khan, S.; Ajmal, S.; Hussain, T.; Rahman, M.U. Clay-based materials for enhanced water treatment: Adsorption mechanisms, challenges, and future directions. J. Umm Al-Qura Univ. Appl. Sci. 2023, 1, 1–16. [Google Scholar] [CrossRef]
  20. Ismail, U.M.; Onaizi, S.A.; Vohra, M.S. Novel MgCuAl-layered triple hydroxide for aqueous selenite and selenate treatment. Emergent Mater 2024, 7, 521–532. [Google Scholar] [CrossRef]
  21. Nazir, M.A.; Najam, T.; Zarin, K.; Shahzad, K.; Javed, M.S.; Jamshaid, M.; Bashir, M.A.; Shah, S.S.A.; Rehman, A.U. Enhanced adsorption removal of methyl orange from water by porous bimetallic Ni/Co MOF composite: A systematic study of adsorption kinetics. Int. J. Environ. Anal. Chem. 2021, 103, 4841–4856. [Google Scholar] [CrossRef]
  22. Tu, N.T.T.; Thien, T.V.; Du, P.D.; Chau, V.T.T.; Mau, T.X.; Khieu, D.Q. Adsorptive removal of Congo red from aqueous solution using zeolitic imidazolate framework-67. J. Environ. Chem. Eng. 2018, 6, 2269–2280. [Google Scholar] [CrossRef]
  23. Prabhakar, N.; Isloor, A.M.; Padaki, M.; Ismail, A.F. Fabrication of TiO2@ZIF-67 metal organic framework composite incorporated PVDF membranes for the removal of hazardous reactive black 5 and Congo red dyes from contaminated water. Chem. Eng. J. 2024, 498, 155270. [Google Scholar] [CrossRef]
  24. Saghir, S.; Xiao, Z. Synthesis of novel Ag@ZIF-67 rhombic dodecahedron for enhanced adsorptive removal of antibiotic and organic dye. J. Mol. Liq. 2021, 328, 115323. [Google Scholar] [CrossRef]
  25. Li, Y.; Xiang, P.; Chen, H.; Zhou, Y. Adsorption performance of one- and two-component anionic dyes using core-shell ZIF-8@ZIF-67. J. Solid State Chem. 2022, 315, 123538. [Google Scholar] [CrossRef]
  26. Haghighat, G.A.; Sadeghi, S.; Saghi, M.H.; Ghadiri, S.K.; Anastopoulos, I.; Giannakoudakis, D.A.; Colmenares, J.C.; Shams, M. Zeolitic imidazolate frameworks (ZIFs) of various morphologies against eriochrome black-T (EBT): Optimizing the key physicochemical features by process modeling. J. Solid State Chem. 2022, 315, 125391. [Google Scholar] [CrossRef]
  27. Kumar, S.; Jain, S.; Nehra, M.; Dilbaghi, N.; Marrazza, G.; Kim, K.H. Green synthesis of metal–organic frameworks: A state-of-the-art review of potential environmental and medical applications. Coord. Chem. Rev. 2020, 420, 213407. [Google Scholar] [CrossRef]
  28. Guo, X.; Kong, L.; Ruan, Y.; Diao, Z.; Shih, K.; Su, M.; Hou, L.; Chen, D. Green and facile synthesis of cobalt-based metal–organic frameworks for the efficient removal of Congo red from aqueous solution. J. Colloid Interface Sci. 2020, 578, 500–509. [Google Scholar] [CrossRef]
  29. Yang, Q.; Lu, R.; Ren, S.; Zhou, H.; Wu, Q.; Zhen, Y.; Chen, Z.; Fang, S. Magnetic beads embedded in poly (sodium-p-styrenesulfonate) and ZIF-67: Removal of nitrophenol from water. J. Solid State Chem. 2018, 265, 200–207. [Google Scholar] [CrossRef]
  30. Saghir, S.; Xiao, Z. Hierarchical mesoporous ZIF-67@LDH for efficient adsorption of aqueous Methyl Orange and Alizarine Red S. Powder Technol. 2020, 377, 453–463. [Google Scholar] [CrossRef]
  31. Rafiee, A.; Nasab, S.G.; Teimouri, A. Synthesis and characterization of pistachio shell/nanodiopside nanocomposite and its application for removal of Crystal Violet dye from aqueous solutions using central composite design. Int. J. Environ. Anal. Chem. 2019, 100, 1624–1649. [Google Scholar] [CrossRef]
  32. Ismail, U.M.; Ibrahim, A.I.; Onaizi, S.A.; Vohra, M.S. Synthesis and application of MgCuAl-layered triple hydroxide/carboxylated carbon nanotubes/bentonite nanocomposite for the effective removal of lead from contaminated water. Results Eng. 2024, 24, 102991. [Google Scholar] [CrossRef]
  33. Marahel, F.; Goodajdar, B.M.; Niknam, L.; Faridnia, M.; Pournamdari, E.; Doost, S.M. Ultrasonic assisted adsorption of methylene blue dye and neural network model for adsorption of methylene blue dye by synthesised Mn-doped PbS nanoparticles. Int. J. Environ. Anal. Chem. 2021, 103, 3059–3080. [Google Scholar] [CrossRef]
  34. Kooh, M.R.R.; Thotagamuge, R.; Chau, Y.F.C.; Mahadi, A.H.; Lim, C.M. Machine learning approaches to predict adsorption capacity of Azolla pinnata in the removal of methylene blue. J. Taiwan Inst. Chem. Eng. 2022, 132, 104134. [Google Scholar] [CrossRef]
  35. Ismail, U.M.; Hussaini, M.; Vohra, M.S. H2S and SO2 toxic gases removal using date palm-tree branches based activated carbon: Experimental findings and machine learning (ML) modeling. Environ. Dev. Sustain. 2024, 1–32. [Google Scholar] [CrossRef]
  36. McCulloch, W.S.; Pitts, W. A logical calculus of the ideas immanent in nervous activity. Bull. Math. Biophys. 1943, 5, 115–133. [Google Scholar] [CrossRef]
  37. Ismail, U.M.; Onaizi, S.A.; Vohra, M.S. Crystal violet removal using ZIF-60: Batch adsorption studies, mechanistic & machine learning modeling. Environ. Technol. Innov. 2023, 33, 103456. [Google Scholar] [CrossRef]
  38. Beigzadeh, B.; Bahrami, M.; Amiri, M.J.; Mahmoudi, M.R. A new approach in adsorption modeling using random forest regression, Bayesian multiple linear regression, and multiple linear regression: 2,4-D adsorption by a green adsorbent. Water Sci. Technol. 2020, 82, 1586–1602. [Google Scholar] [CrossRef]
  39. Ismail, U.M.; Onaizi, S.A.; Vohra, M.S. Aqueous Pb(II) Removal Using ZIF-60: Adsorption Studies, Response Surface Methodology and Machine Learning Predictions. Nanomaterials 2023, 13, 1402. [Google Scholar] [CrossRef]
  40. Hamza, M.A.; Althobaiti, M.M.; Al-Wesabi, F.N.; Alabdan, R.; Mahgoub, H.; Hilal, A.M.; Motwakel, A.; Al Duhayyim, M. Gaussian Process Regression and Machine Learning Methods for Carbon-Based Material Adsorption. Adsorpt. Sci. Technol. 2022, 2022, 3901608. [Google Scholar] [CrossRef]
  41. Bhattacharyya, K.G.; Gupta, S.S. Adsorption of a few heavy metals on natural and modified kaolinite and montmorillonite: A review. Adv. Colloid Interface Sci. 2008, 140, 114–131. [Google Scholar] [CrossRef] [PubMed]
  42. Wang, M.; Liu, J.; Guo, C.; Gao, X.; Gong, C.; Wang, Y.; Liu, B.; Li, X.; Gurzadyan, G.G.; Sun, L. Metal–organic frameworks (ZIF-67) as efficient cocatalysts for photocatalytic reduction of CO2: The role of the morphology effect. J. Mater. Chem. A 2018, 6, 4768–4775. [Google Scholar] [CrossRef]
  43. Gu, X.; Zhou, K.; Feng, Y.; Yao, J. Morphology Control of Zeolitic Imidazolate Framework by Addition of Amino Acid l-Histidine. Chem. Lett. 2015, 44, 1080–1082. [Google Scholar] [CrossRef]
  44. Mamdouh, N.; Farghali, A.A.; El Rouby, W.M.A.; Abdelwahab, A. Effect of ZIF-67-derived Co3O4 on the activity of CNTs/NiCo2O4 nanocomposite for methanol oxidation reaction. Int. J. Hydrogen Energy 2024, 93, 878–887. [Google Scholar] [CrossRef]
  45. Yeerken, A.; Lin, J.; Wang, X.; Luo, Y.; Ma, H. Fabrication of novel porous ZIF-67/PES composite microspheres and the efficient adsorption of triphenylmethane dyes from water. CrystEngComm 2023, 25, 1076–1089. [Google Scholar] [CrossRef]
  46. Narimbi, J.; Balakrishnan, S.; Perova, T.S.; Dee, G.; Swiegers, G.F.; Gun’ko, Y.K. XRD and Spectroscopic Investigations of ZIF—Microchannel Glass Plates Composites. Materials 2023, 16, 2410. [Google Scholar] [CrossRef]
  47. Wi, W.; Nowicki, P.; Bujdák, J. Controversial Issues Related to Dye Adsorption on Clay Minerals: A Critical Review. Molecules 2023, 28, 6951. [Google Scholar] [CrossRef]
  48. Hu, Y.; Song, X.; Zheng, Q.; Wang, J.; Pei, J. Zeolitic imidazolate framework-67 for shape stabilization and enhanced thermal stability of paraffin-based phase change materials. RSC Adv. 2019, 9, 9962–9967. [Google Scholar] [CrossRef]
  49. Ahmed, S.; Vohra, M.S. Selenocyanate removal using combined system of TiO2 photocatalysis and Fe(III)/SiO2 adsorption {UV-TiO2/[Fe(III)/SiO2]}: Complex destruction and simultaneous selenium species adsorption. Chem. Eng. J. 2024, 500, 156525. [Google Scholar] [CrossRef]
  50. Cravillon, J.; Münzer, S.; Lohmeier, S.J.; Feldhoff, A.; Huber, K.; Wiebcke, M. Rapid room-temperature synthesis and characterization of nanocrystals of a prototypical zeolitic imidazolate framework. Chem. Mater. 2009, 21, 1410–1412. [Google Scholar] [CrossRef]
  51. Wu, R.; Qian, X.; Rui, X.; Liu, H.; Yadian, B.; Zhou, K.; Wei, J.; Yan, Q.; Feng, X.; Long, Y.; et al. Zeolitic imidazolate framework 67-derived high symmetric porous Co3O4 hollow dodecahedra with highly enhanced lithium storage capability. Small 2014, 10, 1932–1938. [Google Scholar] [CrossRef]
  52. Madejová, J. FTIR techniques in clay mineral studies. Vib. Spectrosc. 2003, 31, 1–10. [Google Scholar] [CrossRef]
  53. Brigatti, M.F.; Galan, E.; Theng, B.K.G. Chapter 2 Structures and Mineralogy of Clay Minerals. Dev. Clay Sci. 2006, 1, 19–86. [Google Scholar] [CrossRef]
  54. Alam, S.N.; Sharma, N.; Kumar, L. Synthesis of Graphene Oxide (GO) by Modified Hummers Method and Its Thermal Reduction to Obtain Reduced Graphene Oxide (rGO) *. Graphene 2017, 6, 1–18. [Google Scholar] [CrossRef]
  55. Guan, W.; Gao, X.; Ji, G.; Xing, Y.; Du, C.; Liu, Z. Fabrication of a magnetic nanocomposite photocatalysts Fe3O4@ZIF- 67 for degradation of dyes in water under visible light irradiation. J. Solid State Chem. 2017, 255, 150–156. [Google Scholar] [CrossRef]
  56. Yang, Q.; Ren, S.; Zhao, Q.; Lu, R.; Hang, C.; Chen, Z.; Zheng, H. Selective separation of methyl orange from water using magnetic ZIF-67 composites. Chem. Eng. J. 2018, 333, 49–57. [Google Scholar] [CrossRef]
  57. Yang, Q.; Lu, R.; Ren, S.S.; Chen, C.; Chen, Z.; Yang, X. Three dimensional reduced graphene oxide/ZIF-67 aerogel: Effective removal cationic and anionic dyes from water. Chem. Eng. J. 2018, 348, 202–211. [Google Scholar] [CrossRef]
  58. Sarker, T.; Tahmid, I.; Sarker, R.K.; Dey, S.C.; Islam, M.T.; Sarker, M. ZIF-67-based materials as adsorbent for liquid phase adsorption-a review. Polyhedron 2024, 260, 117069. [Google Scholar] [CrossRef]
  59. Hussaini, M.; Vohra, M. LDH-TiO2 Composite for Selenocyanate (SeCN−) Photocatalytic Degradation: Characterization, Treatment Efficiency, Reaction Intermediates and Modeling. Nanomaterials 2022, 12, 2035. [Google Scholar] [CrossRef]
  60. Balarak, D.; Zafariyan, M.; Igwegbe, C.A.; Onyechi, K.K.; Ighalo, J.O. Adsorption of Acid Blue 92 Dye from Aqueous Solutions by Single-Walled Carbon Nanotubes: Isothermal, Kinetic, and Thermodynamic Studies. Environ. Process. 2021, 8, 869–888. [Google Scholar] [CrossRef]
  61. Haghdoust, S.; Arabkhani, P.; Ghaderi, S.; Ghaedi, M.; Asfaram, A. Rapid microwave-assisted synthesis of a magnetic biochar@ZIF-67: An efficient nanocomposite-based adsorbent for the dye-contaminated water cleanup. New J. Chem. 2023, 47, 9257–9270. [Google Scholar] [CrossRef]
  62. Rattanapan, S.; Srikram, J.; Kongsune, P. Adsorption of Methyl Orange on Coffee grounds Activated Carbon. Energy Procedia 2017, 138, 949–954. [Google Scholar] [CrossRef]
  63. Xiao, B.; Liu, S.Q. Photocatalytic Oxidation of Ammonia via an Activated Carbon-Nickel Ferrite Hybrid Catalyst under Visible Light Irradiation. Acta Phys. Chim. Sin. 2014, 30, 1697–1705. [Google Scholar] [CrossRef]
  64. AlNaimat, S.; Ismail, U.M.; Ibrahim, A.I.; Muse, A.; Faheem, K.; Mustafa, M.; Vohra, M.S.; Onaizi, S.A. Machine learning modeling and statistical optimization of dye removal from contaminated water using CTAB-functionalized graphene oxide. Water Air Soil Pollut. 2024, 235, 671. [Google Scholar] [CrossRef]
  65. Sharma, A.; Mona, S.; Sharma, P. Greenly synthesised novel microporous ZnO/GO for adsorption of methyl orange and malachite green. Int. J. Environ. Anal. Chem. 2024, 1–25. [Google Scholar] [CrossRef]
  66. Sun, J.; Semenchenko, L.; Lim, W.T.; Rivas, M.F.B.; Varela-Guerrero, V.; Jeong, H.K. Facile synthesis of Cd-substituted zeolitic-imidazolate framework Cd-ZIF-8 and mixed-metal CdZn-ZIF-8. Microporous Mesoporous Mater. 2018, 264, 35–42. [Google Scholar] [CrossRef]
  67. Wang, J.; Guo, X. Adsorption isotherm models: Classification; physical meaning, application and solving method. Chemosphere 2020, 258, 127279. [Google Scholar] [CrossRef]
  68. Ayawei, N.; Ebelegi, A.N.; Wankasi, D. Modelling and Interpretation of Adsorption Isotherms. J. Chem. 2017, 2017, 3039817. [Google Scholar] [CrossRef]
  69. Zhao, G.; Huang, X.; Tang, Z.; Huang, Q.; Niu, F.; Wang, X. Polymer-based nanocomposites for heavy metal ions removal from aqueous solution: A review. Polym. Chem. 2018, 9, 3562–3582. [Google Scholar] [CrossRef]
  70. Gil, A.; Santamaría, L.; Korili, S.; Vicente, M.; Barbosa, L.; de Souza, S.; Marçal, L.; de Faria, E.; Ciuffi, K. A review of organic-inorganic hybrid clay based adsorbents for contaminants removal: Synthesis, perspectives and applications. J. Environ. Chem. Eng. 2021, 9, 105808. [Google Scholar] [CrossRef]
  71. Huang, P.; Miao, X.; Wu, J.; Zhang, P.; Zhang, H.; Bai, S.; Liu, W. Facile synthesis of an ultrathin ZIF-67 layer on the surface of Sn/Ti co-doped hematite for efficient photoelectrochemical water oxidation. Dalton Trans. 2022, 51, 8848–8854. [Google Scholar] [CrossRef] [PubMed]
  72. Zhang, T.; Wang, W.; Zhao, Y.; Bai, H.; Wen, T.; Kang, S.; Song, G.; Song, S.; Komarneni, S. Removal of heavy metals and dyes by clay-based adsorbents: From natural clays to 1D and 2D nano-composites. Chem. Eng. J. 2021, 420, 127574. [Google Scholar] [CrossRef]
  73. Zaffino, C. Exploiting the Potentiality of Surface-Enhanced Raman Scattering (sers) for the Identification of Natural Dyes in Ancient Textiles: From Hyphenated Techniques to In-Situ Analysis. Ph.D. Thesis, Università Degli Studi di Milano, Milano, Italy, 2015. [Google Scholar] [CrossRef]
  74. Yang, J.-M.; Ying, R.-J.; Han, C.-X.; Hu, Q.-T.; Xu, H.-M.; Li, J.-H.; Wang, Q.; Zhang, W. Adsorptive removal of organic dyes from aqueous solution by a Zr-based metal-organic framework: Effects of Ce(III) doping. Dalton Trans. 2018, 47, 3913–3920. [Google Scholar] [CrossRef]
  75. Gao, R. Mechanism and Application of Metal-Organic Framework in Wastewater Treatment. Highlights Sci. Eng. Technol. 2022, 6, 25–32. [Google Scholar] [CrossRef]
  76. Tang, S.; Wang, Y.; He, P.; Wang, Y.; Wei, G. Recent Advances in Metal–Organic Framework (MOF)-Based Composites for Organic Effluent Remediation. Materials 2024, 17, 2660. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The chemical structure of methyl orange (MO).
Figure 1. The chemical structure of methyl orange (MO).
Applsci 15 03562 g001
Figure 2. A general framework for evaluating MO adsorption using ZIF-67@BNT adsorbent along with the ML modeling.
Figure 2. A general framework for evaluating MO adsorption using ZIF-67@BNT adsorbent along with the ML modeling.
Applsci 15 03562 g002
Figure 3. A schematic illustration depicting the preparation procedure of the ZIF-67@Bentonite (ZIF-67@BNT) nanocomposite.
Figure 3. A schematic illustration depicting the preparation procedure of the ZIF-67@Bentonite (ZIF-67@BNT) nanocomposite.
Applsci 15 03562 g003
Figure 4. (A) ANN algorithm structure, (B) linear hyperplane-based SVR classification, (C) Random Forest algorithm architecture, and (D) Regressor for the Gaussian process.
Figure 4. (A) ANN algorithm structure, (B) linear hyperplane-based SVR classification, (C) Random Forest algorithm architecture, and (D) Regressor for the Gaussian process.
Applsci 15 03562 g004
Figure 5. The ZIF-67@BNT nanocomposite’s crystalline structure and functional group characterization are shown in (A) XRD patterns and (B) FTIR spectra, respectively.
Figure 5. The ZIF-67@BNT nanocomposite’s crystalline structure and functional group characterization are shown in (A) XRD patterns and (B) FTIR spectra, respectively.
Applsci 15 03562 g005
Figure 6. Thermogravimetric analysis results for the ZIF-67@BNT nanocomposite.
Figure 6. Thermogravimetric analysis results for the ZIF-67@BNT nanocomposite.
Applsci 15 03562 g006
Figure 7. Benchmarking the adsorption of MO onto ZIF-67@BNT with its constituents (i.e., ZIF-67 and BNT). The adsorption took place from 100 mg/L MO solution containing 50 mg/mL of each adsorbent.
Figure 7. Benchmarking the adsorption of MO onto ZIF-67@BNT with its constituents (i.e., ZIF-67 and BNT). The adsorption took place from 100 mg/L MO solution containing 50 mg/mL of each adsorbent.
Applsci 15 03562 g007
Figure 8. The RSM Model (for MO adsorption by ZIF-67@BNT) check plots. (a) Predicted Vs. Actual adsorption capacity, (b) Box–Cox plot, (c) Normality distribution plot, and (d) Residuals Vs. Run plot.
Figure 8. The RSM Model (for MO adsorption by ZIF-67@BNT) check plots. (a) Predicted Vs. Actual adsorption capacity, (b) Box–Cox plot, (c) Normality distribution plot, and (d) Residuals Vs. Run plot.
Applsci 15 03562 g008
Figure 9. MO adsorption onto ZIF-67@BNT as a function of (a) ZIF-67@BNT dose, (b) MO concentration, and (c) Time.
Figure 9. MO adsorption onto ZIF-67@BNT as a function of (a) ZIF-67@BNT dose, (b) MO concentration, and (c) Time.
Applsci 15 03562 g009
Figure 10. The 3D interaction plots for MO adsorption capacity: (a) MO concentration vs. ZIF-67@BNT dose, (b) Time vs. ZIF-67@BNT dose, (c) ZIF-67@BNT dose vs. MO concentration.
Figure 10. The 3D interaction plots for MO adsorption capacity: (a) MO concentration vs. ZIF-67@BNT dose, (b) Time vs. ZIF-67@BNT dose, (c) ZIF-67@BNT dose vs. MO concentration.
Applsci 15 03562 g010
Figure 11. Results from the RSM-based response optimization exercise for MO adsorption using ZIF-67@BNT.
Figure 11. Results from the RSM-based response optimization exercise for MO adsorption using ZIF-67@BNT.
Applsci 15 03562 g011
Figure 12. Adsorption kinetic model for MO adsorption by ZIF-67@BNT nanocomposite.
Figure 12. Adsorption kinetic model for MO adsorption by ZIF-67@BNT nanocomposite.
Applsci 15 03562 g012
Figure 13. Adsorption Isotherm model for MO adsorption by ZIF-67@BNT nanocomposite.
Figure 13. Adsorption Isotherm model for MO adsorption by ZIF-67@BNT nanocomposite.
Applsci 15 03562 g013
Figure 14. FTIR Characterization of ZIF-67@BNT nanocomposite Before and After Adsorption.
Figure 14. FTIR Characterization of ZIF-67@BNT nanocomposite Before and After Adsorption.
Applsci 15 03562 g014
Figure 15. Actual vs. predicted machine learning models.
Figure 15. Actual vs. predicted machine learning models.
Applsci 15 03562 g015
Figure 16. Taylor diagram for both testing and training.
Figure 16. Taylor diagram for both testing and training.
Applsci 15 03562 g016
Table 1. The RSM modeling factors and levels used for design of experiments (DOE).
Table 1. The RSM modeling factors and levels used for design of experiments (DOE).
FactorsLevels
−101
AAdsorbent dose (mg)51015
BMethyl orange (mg/L)102030
CTime (hours)135
Table 2. The RSM-based DOE and the MO adsorption results using the ZIF-67@BNT nanocomposite adsorbent.
Table 2. The RSM-based DOE and the MO adsorption results using the ZIF-67@BNT nanocomposite adsorbent.
RunA: Adsorbent Dose (mg)B: Methyl Orange (mg/L)C: Time (Hours)Adsorption Capacity (mg/g)
11530582
2520389
31530146
410205105
51010362
6510584
71020385
85301109
91020175
105305187
111510544
12510141
1310303124
141510138
151520360
Table 3. The respective statistical parameters for RSM modeling.
Table 3. The respective statistical parameters for RSM modeling.
SourceSum of Squares Degree of Freedom (df)Mean SquareF-Valuep-Value
Model2.7150.542436.79<0.0001Significant
A: Adsorbent dose0.782310.782353.06<0.0001
B: Methyl Orange1.0811.0873.49<0.0001
C: Time0.533010.533036.15<0.000
AB0.118710.11878.050.0002
A20.194210.194213.170.0195
Residual 0.132790.0147 0.0055
Core Total2.8414
Table 4. Comparison with the other known adsorbent materials for the removal of MO.
Table 4. Comparison with the other known adsorbent materials for the removal of MO.
AdsorbentInitial ConcentrationAdsorbent DoseInitial pHTemperatureTimeAdsorption CapacityReference
Ni@ZIF-6740206Room18033.17[21]
Ag@ZIF-6720020630180994.6[24]
ZIF-67@LDH 20630500663.4[30]
CGAC3005034090658[62]
AC/NiFe2O4100300330180182.82[63]
CS-MT Composite403206.33060154.4[17]
CTAB@GO300504Room14401301.5[64]
ZnO/GO Nanocomposite402046050296.73[65]
Cd-ZIF-8 Composite20050525120425.8[66]
This Study305-Room300 187-
Table 5. Kinetic modeling parameters for MO adsorption by ZIF-67@BNT nanocomposite.
Table 5. Kinetic modeling parameters for MO adsorption by ZIF-67@BNT nanocomposite.
Models
Pseudo-First Order
q t = q e 1 e k 1 t
Pseudo-Second Order
q t = q e 2 k 2 t 1 + q e k 2 t
Avrami
q t = q e 1 e k a t n n
qe (mg/g)k1 (min−1)R2qe (mg/g)k2 (g/mg.min)R2qe (mg/g)ka (min−1)nR2
3380.02160.8837360.79.464 × 10−50.9585386.80.00130.4010.9949
Table 6. Adsorption isotherm modeling parameters for MO adsorption by ZIF-67@BNT nanocomposite.
Table 6. Adsorption isotherm modeling parameters for MO adsorption by ZIF-67@BNT nanocomposite.
Models
Langmuir Adsorption Isotherm
q e = q max K L C e 1 + K L C e
Freundlich Adsorption Isotherm
q e = K F C e 1 / n
Redlich–Peterson Adsorption Isotherm
q e = K R C e 1 + a R C e g
qmax (mg/g)KL (L/mg)R2nKFR2KR (L/g)aR (L/mg)gR2
528.70.0340.97462.4859.80.902414.060.00811.230.9823
Table 7. Descriptive statistics for the machine learning dataset.
Table 7. Descriptive statistics for the machine learning dataset.
IndexAdsorbent Dose (mg)Methyl Orange (mg/L)Time (Hours)Adsorption Capacity (mg/g)
Count22222222
Mean8.4026.819.68112.49
Median520585.41
Mode51024-
Minimum55115.18
Maximum1510024241.80
Range109523226.61
Standard deviation4.1922.6010.1073.91
Sample variance17.58510.82102.135463.98
Table 8. Performance of machine learning models for the adsorption capacity by ZIF-67@BNT.
Table 8. Performance of machine learning models for the adsorption capacity by ZIF-67@BNT.
TestingTraining
AlgorithmR2RMSEMAER2RMSEMAE
ANN0.87225.58221.1430.9990.8330.553
Random forest0.92738.83533.8880.95518.82213.374
Support vector regression0.907110.95696.3850.9983.9971.476
Gaussian process regression0.96221.68714.9890.9964.1793.051
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

Faheem, K.; Onaizi, S.A.; Vohra, M.S. Novel ZIF-67@Bentonite (ZIF-67@BNT) Nanocomposite for Aqueous Methyl Orange Removal. Appl. Sci. 2025, 15, 3562. https://doi.org/10.3390/app15073562

AMA Style

Faheem K, Onaizi SA, Vohra MS. Novel ZIF-67@Bentonite (ZIF-67@BNT) Nanocomposite for Aqueous Methyl Orange Removal. Applied Sciences. 2025; 15(7):3562. https://doi.org/10.3390/app15073562

Chicago/Turabian Style

Faheem, Kashif, Sagheer A. Onaizi, and Muhammad S. Vohra. 2025. "Novel ZIF-67@Bentonite (ZIF-67@BNT) Nanocomposite for Aqueous Methyl Orange Removal" Applied Sciences 15, no. 7: 3562. https://doi.org/10.3390/app15073562

APA Style

Faheem, K., Onaizi, S. A., & Vohra, M. S. (2025). Novel ZIF-67@Bentonite (ZIF-67@BNT) Nanocomposite for Aqueous Methyl Orange Removal. Applied Sciences, 15(7), 3562. https://doi.org/10.3390/app15073562

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