Next Article in Journal
Polystyrene–Carbon Nanotube Composites: Interaction Mechanisms, Preparation Methods, Structure, and Rheological Properties—A Review
Previous Article in Journal
Random Forest-Based Machine Learning Model Design for 21,700/5 Ah Lithium Cell Health Prediction Using Experimental Data
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Cadmium Removal by Raw Argan Shell Adsorbent: Experimental and Theoretical Investigations for Ecological Applications

1
Laboratory of Spectroscopy, Molecular, Modeling, Materials, Nanomaterials, Water and Environment, (LS3MNWE), Department of Chemistry, Faculty of Sciences, Mohammed V University in Rabat, Av. Ibn Battouta, B.P. 1014, Rabat 10000, Morocco
2
Laboratory of Spectroscopy, Molecular, Modeling, Materials, Nanomaterials, Water and Environment, (LS3MNWE), ENSAM-Rabat, Mohammed V University in Rabat, Av. Ibn Battouta, B.P. 1014, Rabat 10000, Morocco
3
Higher School of Education and Training, Chouaib Doukkali University, El Jadida 24000, Morocco
4
Higher School of Education and Training, Mohammed I University, Oujda 60000, Morocco
*
Author to whom correspondence should be addressed.
Physchem 2025, 5(1), 13; https://doi.org/10.3390/physchem5010013
Submission received: 15 January 2025 / Revised: 15 February 2025 / Accepted: 19 February 2025 / Published: 19 March 2025
(This article belongs to the Section Kinetics and Thermodynamics)

Abstract

:
The removal of cadmium ions (Cd2+) using raw argan shells (ArS) was optimized through experimental and theoretical studies. Adsorption experiments revealed optimal conditions at an adsorbent dose of 3.5 g, an initial Cd2+ concentration of 20 mg·L−1, and a pH of 8, achieving a maximum sorption capacity of 3.92 mg·g−1. The kinetic analysis showed that the adsorption followed a pseudo-second-order model (R2 = 0.98), and the Langmuir isotherm model predicted a maximum adsorption capacity of 4 mg·g−1. Thermodynamic analysis indicated an endothermic adsorption process, with ΔG° shifting from positive to negative as temperature increased, confirming that adsorption is favored at higher temperatures. Desorption studies demonstrated that HCl was the most effective eluting agent, achieving a desorption efficiency of 90.02%, followed by HNO3 (76.65%) and CH3COOH (71.59%). The varying desorption efficiencies were attributed to differences in acid strength and ionic interactions with Cd2+. This study demonstrates the potential of raw argan shells as an efficient, reusable, and sustainable biosorbent for cadmium removal, offering a promising solution for water treatment and environmental remediation.

1. Introduction

Heavy metals, including both transition and post-transition elements like lead, cad-mium, thallium, and antimony, are commonly found in industrial environments and are significant environmental pollutants due to their hazardous nature [1]. Due to the growth of the mining sector and industrial activities, the discharge of heavy metals into the environment has increased [2]. These metal cations are commonly found in effluents from mining drainage, dye waste, paint and ink manufacturing, electroplating and porcelain polishing, and battery production. High concentrations of metal ions negatively affect water quality, aquatic life, terrestrial flora and fauna, and human health [3]. Heavy metal ions do not biodegrade like organic contaminants do; they cannot be broken down into harmless by-products, leading to bioaccumulation in living organisms and eventual transmission across the food chain to humans [4]. Among these heavy metals, cadmium stands out due to its prevalence in various industrial processes and its particularly harmful impact on ecosystems and public health [5]. Indeed, the World Health Organization (WHO) categorizes cadmium as a highly toxic substance [6], and it is released through multiple industrial procedures such as the fabrication of electronic components and batteries. Additionally, cadmium pollution originates from agricultural sources, as cadmium is often present in phosphate fertilizers due to its occurrence in phosphate rock, the primary raw material used to produce these fertilizers [7]. When phosphate rock is processed to extract phosphoric acid (the key component of phosphate fertilizers), cadmium, which can be found in concentrations ranging from 10 to 80 g per ton in phosphate minerals, may also be released and remain in the final product [6]. Despite being a relatively minor constituent of phosphate rock, the accumulation of cadmium in fertilizers poses significant environmental and health risks. When such fertilizers are used on farmland, cadmium has the potential to infiltrate soil and water systems [8]. Several studies have shown that cadmium accumulates in soils and seeps into groundwater and surface waters, exhibiting bioaccumulative properties. Eutrophication is one of the issues associated with the use of phosphate fertilizers, significantly impacting the environment [9]. However, cadmium pollution in aquatic ecosystems continues to be a worldwide problem. Research has established a direct link between cadmium exposure and bone and kidney issues, as it disrupts filtration mechanisms and can potentially lead to liver cancer [6]. Even at low quantities, cadmium can be harmful to human health, hence it is imperative that contaminated water sources be treated to ensure environmental and public health. Given the negative effects associated with this situation, international regulatory bodies have urged the implementation of strict regulations limiting the discharge of heavy metals into sewers and receiving waterways [10]. Therefore, it is crucial to develop techniques that are both cost-efficient and environmentally friendly to remove metallic cations from wastewater [11]. Although challenging, the removal of these metal cations can be achieved through various traditional treatment methods, including ion exchange, chemical precipitation, coagulation–flocculation, and flotation [12]. These techniques have several disadvantages, namely expensive implementation, insufficient selectivity, and only partial elimination of metal ions, particularly when the concentration is less than 50 mg·L−1 [9]. Consequently, it is essential to create a new affordable technique to minimize damage. For this purpose, biosorption presents a potential method. Several research studies have been released regarding the possibility of biosorbents to remove cadmium ions from an aqueous medium, such as pine bark, walnut shells, cashew activated carbon, raw walnut shells, coffee beans, rice husk ash, activated carbon obtained from the husk of rice, and Pachira aquatica Aubl. fruit peels [13,14,15,16,17,18,19,20]. This process is technically an effective method for eliminating inorganic and organic contaminants in an aqueous solution. Recent research has also explored the use of mesoporous silica derivatives for the selective recovery of cadmium, cobalt, and nickel from Ni-Cd batteries [21], highlighting the importance of porous materials in heavy metal adsorption. However, since the argan tree is a naturally abundant and unique resource in Morocco, argan shells are readily accessible and affordable [22]. Utilizing this agricultural waste not only meets a local need by adding value to an unexploited by-product but also lowers the total cost of the adsorption process. Furthermore, argan shells possess physicochemical properties favorable for heavy metal adsorption. Their porous structure and chemical composition allow effective interaction with metal ions, maximizing adsorption capacity [23]. These characteristics make argan shells an excellent candidate compared to other biomaterials, contributing to an eco-friendly and sustainable solution for decontaminating polluted water sources.
To our knowledge, this is the first study to specifically evaluate the adsorption potential of raw argan nut shells for cadmium removal from aqueous solutions. This study highlights the untapped potential of a local, natural resource from Morocco, contributing to the development of an eco-friendly and cost-effective biosorbent for heavy metal remediation. Consequently, the aim of this study is to remove cadmium cations (Cd2+) from contaminated water through biosorption using natural adsorbents, specifically raw argan nut shells, which are considered an untapped plant waste with significant potential. This experiment was conducted in a controlled cadmium solution, rather than actual wastewater, to precisely evaluate the adsorption capacity of argan powder for cadmium.

2. Materials and Methods

2.1. Natural Adsorbent: Preparation of Argan Nut Shell Powder (ArS)

Argan shells, harvested near Aoulouz, were first broken into small fragments using a Retsch RM 200 mortar grinder in their natural state (referred to as ArS). The primary purpose of mechanical grinding was to reduce crystallinity levels [24]. The resulting powder was then sieved to achieve particle sizes between 125 and 63 µm in diameter. After sieving, the powder was cleaned with hot bi-distilled water to effectively remove all contaminants, followed by a drying phase.

2.2. Adsorbate: Synthetic Solution Preparation

To prepare the cadmium stock solution, 10 mg of cadmium chloride (CdCl2) was dis-solved in 100 mL of ultrapure water sourced from a Millipore Milli-Q system, where the resistivity is more than 18 MΩ·cm at ambient temperature. The CdCl2, sourced from Sigma Aldrich, possesses a molecular weight of 183.32 g·mol−1, a water solubility of 119.6 g/100 mL at ambient temperature, and a purity of 99.99%. From this stock solution, three working solutions were prepared with Cd2+ concentrations of 5 mg·L−1, 12.5 mg·L−1, and 20 mg·L−1 to evaluate the adsorption of cadmium by argan shell powder. pH adjustments were per-formed directly in the working solutions using high-quality reagents, NaOH (99.0%) and HCl (37%), obtained from Merck & Co. (Darmstadt, Germany).

2.3. Characterization

2.3.1. Scanning Electron Microscopy (SEM-EDX)

A SEM-EDX study was performed to examine the surface elemental composition and morphology of ArS. The measurements were performed utilizing a JEOL JSM-IT100 scanning electron microscope, operated at an acceleration voltage of 20 kV.

2.3.2. FTIR Analysis

Infrared transmission spectra were recorded using a PerkinElmer RX-1 spectrometer. A sample pellet with 1 wt% concentration was dispersed in KBr for analysis. The infrared spectrum was measured between 4000 and 500 cm−1, with a 4 cm−1 spectral resolution.

2.3.3. pH Zero Charge Point: pHzpc

pHzpc provides valuable insight into the electrostatic forces involved in adsorption [25]. In this analysis, 0.15 g of ArS was suspended in 50 mL of NaCl solution at a concentration of 0.01 M, with the initial pH adjusted from 2 to 12. Suspensions were agitated for 48 h and then filtered, after which the final pH (pHf) was measured. The variation in pH (ΔpH = pHf − pHi) was graphed relative to the starting pH (pHi), and the pHzpc was determined at the point where the curve crosses the axis of zero.
The pHzpc provides useful information on the electronic and chemical characteristics of the surface groups [26].

2.3.4. Boehm Titration

The functional groups present on the surface of ArS were analyzed using the method of Boehm, which allows for a global determination of basic functions while measuring acidic functions individually. In this method, 0.15 g of ArS was added to 50 mL of aqueous solutions of NaOH, Na2CO3, and NaHCO3, as well as a 0.01 M HCl solution. NaOH neutralizes all lactonic, phenolic, and carboxylic functions; sodium carbonate neutralizes lactonic and carboxylic functions; and sodium bicarbonate neutralizes only carboxylic functions on the surface of the material. Each solution was stirred for 24 h to ensure that the majority of the functional groups reacted, after which the solutions were filtered. The final step involved titrating a 10 mL sample from each filtrate using alkaline solutions against 0.01 M HCl and acidic solutions against 0.01 M NaOH [27].

2.3.5. Specific Area

The knowledge of the mass area of a sample, referred to as the specific surface area, is of great importance in physicochemical studies, particularly in the adsorption field. The specific surface area is essential for characterizing finely divided solids and porous materials, as it is directly related to the reactivity of the solid. To determine the surface area of ArS, the methylene blue technique was utilized. In this experiment, 0.1 g of ArS was combined with 50 mL of prepared mixture at various concentrations, agitated for 2 h, and then filtered. The filtrate was subsequently analyzed. The Langmuir model was applied, using the equation below, to evaluate the available surface area [28]:
Sa = (Qm × NA × S)/M
where Sa represents the surface area (m2·g−1); Qm denotes the highest adsorption capacity, expressed in mg·g−1; NA represents the Avogadro number which is 6.023 × 1023 mol−1; S refers to the area occupied by a single methylene blue molecule, which is equal to 119 Å2; and M represents the molar mass of hydrated methylene blue which is 319.86 g·mol−1.

2.3.6. Inductively Coupled Plasma Mass Spectrometry (ICP-MS)

ICP-MS was utilized to perform all trace element analyses in this study, employing a Thermo Scientific X-SERIES 2 instrument (Waltham, MA, USA). This highly sensitive and precise technique allows for the detection of trace elements at low concentrations, making it particularly valuable for analyzing heavy metals, such as cadmium, in environmental samples.

2.4. Response Surface Methodology (RSM): Batch Experiments

RSM is a powerful statistical tool used to optimize batch adsorption experiments by modeling how independent factors and response are related. It efficiently determines optimal conditions for adsorption systems, reducing the need for extensive experimentation [29]. In this study, analyses were conducted using RSM via JMP 17.2 software, which combines statistical and mathematical approaches [26]. This software enabled the evaluation of parameter influences on process efficiency and interaction effects through a predictive model [26]. The experimental design applied a second-degree polynomial function, utilizing the Box–Behnken optimization design with three factors: X1 (initial concentration), X2 (pH), and X3 (adsorbent dose), with coded levels listed in Table 1.

2.5. Batch Adsorption Process

Three input factors with three evenly distributed values, represented by −1, 0, and 1, were investigated during the batch adsorption studies, in the manner described in Table 1. The ideal adsorption conditions were determined by varying the Cd2+ concentration from 5 to 20 mg·L−1, the pH within the range of 4 to 8, and the ArS dose between 1 and 4 g·L−1 of 50 mL. The blend was agitated at 250 rpm under a temperature of 25 °C. Following the adsorption experiment, the ArS was isolated from the mixture using a 185 mm diameter Whatman Grade 595 filter paper. Residual Cd2+ concentrations in the liquid phase were analyzed using AES HORIBA ULTIMA EXPERT 2 (ICP-AES) (Loos, France). The adsorption capacity, expressed in milligrams of Cd2+ ions for every gram of ArS adsorbent, was determined utilizing the Formula (2) below Es-said [30]:
qt = (C0 − Ct)·V/m
where C0 and Ct represent the initial and final concentrations of heavy metal ions, respectively (mg·L−1), V is the volume of the entire solution (L), and m represents the mass of ArS adsorbent (g).
Various kinetic models, including pseudo-first-order and pseudo-second-order intra-particle diffusion, were employed to assess the kinetics of adsorption. Additionally, adsorption isotherm parameters were determined utilizing the Temkin, Dubinin–Radushkevich, Freundlich, and Langmuir models. Furthermore, the effect of temperature on the adsorption process was analyzed at 25 °C, 35 °C, 50 °C, and 65 °C, allowing us to calculate the associated thermodynamic factors.

3. Results and Discussion

3.1. Characterization

To characterize the ArS surface including its morphology and functional groups, SEM and FTIR techniques were employed, as adsorption is primarily a surface phenomenon.

3.1.1. SEM Analysis of Argan Nutshell Powder

The SEM image of argan nut shells in Figure 1 reveals key surface characteristics. The outer surface exhibits an irregular, heterogeneous texture with visible grooves and fissures contributing to a rough appearance. Small, uniformly distributed pores are also visible, suggesting the presence of intercellular cavities and micropores, which are advantageous for trapping and adsorbing metal cations. The ArS chemical composition was revealed using energy-dispersive X-ray spectroscopy (EDX), showing the material is primarily plant-based. Carbon is predominant at 53.09% of the weight, followed by oxygen at 45.98% and nitrogen in a smaller amount, with 0.94% by weight. This elemental profile reflects the organic nature of the material and highlights the potential of argan nut shells as an adsorbent due to the high carbon content, which is often associated with enhanced adsorption capabilities.

3.1.2. Fourier Transform Infrared Spectroscopy (FTIR)

The ArS FTIR spectrum, measured from 500 to 4000 cm−1 and shown in Figure 2, reveals characteristic absorbance bands corresponding to various functional groups. A prominent broad band at 3340.37 cm−1 suggests the detection of hydroxyl functional groups [31]. The presence of the OH band in the FTIR spectrum of raw argan shell, even without pretreatment, can be attributed to the intrinsic hydroxyl groups found in lignocellulosic components such as cellulose, hemicellulose, and lignin [32]. Additionally, the natural adsorption of atmospheric moisture and the formation of hydrogen bonds between hydroxyl groups contribute to the observed signal. Similar findings have been reported in other lignocellulosic biomass studies, confirming that untreated plant-based materials can exhibit OH stretching vibrations due to their inherent structure and surface properties.
The peak detected around 1731.60 cm−1 is attributed to the stretching vibrations of C=O, likely from carboxylic acids or ester groups associated with lignin or hemicelluloses [31]. A peak at 1594.59 cm−1 indicates that the cellulose structure has absorbed water [33]. The band at 1422.27 cm−1 corresponds to CH2 symmetric deformations in cellulose [33]. The existence of CH and C=O groups in aromatic rings of polysaccharides is designated by the band at 1372.17 cm−1 [34]. Additionally, the absorbance band around 1234.82 cm−1 is linked to C=O stretching vibrations in acetyl groups within hemicellulose and lignin [31]. The band at 1031.13 cm−1 arises from C-O and O-H bond vibrations in the cellulosic structure [35], while a peak at 871.19 cm−1 corresponds to glycosidic bonds in polysaccharides. The last feature observed at 670.41 cm−1 is related to C-OH deformations in cellulose [36].

3.1.3. pH of Point Zero Charge: pHzpc

Surface properties of the adsorbent play a crucial role in the relationship with the adsorbate during the sorption mechanism [37]. The pHzpc represents the pH where the surface charge of the adsorbent becomes neutral, offering insights into reactive groups on the surface of the biosorbent and whether acidic or basic characteristics predominate. The pHzpc of ArS was found experimentally at pH 3.6 (Figure 3). For pH levels exceeding the pHzpc, the surface of ArS carries a negative charge, increasing the electrostatic attraction between metal cations and the surface, which increases cation adsorption [37]. Conversely, at pH values below pHzpc, the biomass surface becomes positively charged, favoring interactions with anions. Additionally, decreased metal cation sorption can occur due to competitive sorption between H3O+ and Cd2+ ions. Therefore, pHzpc plays a key role in controlling and enhancing the adsorption mechanism.

3.1.4. Boehm Titration

According to Boehm titration, the amount of functional groups that are acidic (0.133 mmol·g−1) on the adsorbent surface is greater than that of basic functional groups (0.092 mmol/g) (Table 2). This indicates that the majority of reactive sites within the adsorbent show acidity, suggesting a predominance of oxygenated surface-bound groups [38]. which are essential to the process of adsorption by enhancing the attraction of the adsorbent toward positively charged ions. The acidic nature of these functional groups contributes to a more negative surface charge once the pH level of the solution exceeds the ArS pHPZC which is equal to 3.6. This negative surface charge increases the attraction of metal cations (Cd2+) through electrostatic forces and the adsorbent, thereby improving the adsorption efficiency. In contrast, the lower quantity of basic functional groups suggests they play a less significant role in this context. Along with the functional groups, ArS surface area (31.99 m2·g−1) is an important factor that influences its adsorption capacity. A larger surface area generally enhances interaction with adsorbate molecules by offering additional adsorption active sites [39]. This relatively large surface area enables ArS to accommodate significant amounts of Cd2+, which could improve adsorption process efficiency. Together, the large surface area and the acidic functional groups make ArS an effective adsorbent for targeting metal ions in aqueous solutions.

3.2. Batch Experiments: Response Surface Methodology

3.2.1. Study of the Parameters: Statistical Study

To optimize cadmium removal efficiency, 15 experiments were conducted simultaneously. The response (Y) represents the amount of cadmium adsorbed. A second-degree linear polynomial regression model was employed to establish the connection between (Y) and the three independent variables along with their interactions [40] (Tong et al., 2011). The following Equation (3) represents this model:
Y = a0 + b1X1 + b2X2 + b3X3 + b11(X1)2 + b22(X2)2 + b33(X3)2 + b12(X1X2) + b13(X1X3) + b23(X2X3)
In this equation, a0 represents the constant term, bi denotes the linear coefficients, and bii signifies the coefficients for quadratic interactions. The variables Xi (where i = 1, 2, 3) correspond to the independent factors in the process, while Y is the dependent variable representing the amount of cadmium adsorbed.
As shown in Table 3, different combinations of pH, ArS dose, and initial Cd2+ concentration were tested to determine adsorption effectiveness. This matrix serves as a foundation for analyzing how variations in pH, ArS dose, and initial Cd2+ concentration influence adsorption capacity, supporting RSM optimization of the adsorption process.

3.2.2. Correlation Coefficient

According to the results from the JMP software (17.2 version) and illustrated in Figure 4, a high correlation was found between the predicted (based on the mathematical model) and observed Cd2+ adsorption values, having a strong correlation coefficient (R2 = 0.988), reflecting that the actual experimental results and the model predictions correspond well. This high R2 value, being close to 1, suggests that the second-degree polynomial model is well-suited for accurately describing the relation between Cd2+ adsorption and the selected variables. The combination of a high R2 value, a low RMSE of 0.1864, and a low p-value of 0.0003 demonstrate that the model fits well and is reliable for describing the adsorption process.

3.2.3. Analysis of Variance

In this analysis, ANOVA is applied to examine the statistical significance of the predictive model for cadmium adsorption. This approach determines whether the independent variables (pH, ArS dose, and [Cd2+]i) significantly impact the amount of Cd2+ adsorbed. The ANOVA findings of the model are summarized in Table 4. The findings indicate that the model significance is statistically meaningful with 95% confidence in predicting Cd2+ adsorption, with a p-value corresponding to 0.003 (p < 0.05). This result is further supported by an F-ratio of 47.8409, which suggests a significant impact of the independent factors on Cd2+ adsorption. Both the low p-value and the high F-ratio confirm that the second-degree polynomial model is statistically significant and reliable for predicting cadmium adsorption under the tested conditions.

3.2.4. Residue Review

Figure 5 presents the residual values from the various experiments. The residual plot shows a dispersed pattern, with residuals randomly scattered around zero and lacking any specific trend. This random distribution suggests that the model accurately describes the data, as it aligns with the expectation that residuals should be randomly distributed for a well-fitting model. All residual values fall within the range (−0.20, 0.25), indicating the absence of outliers [41]. This narrow range of residuals confirms the reliability of the model and stability in predicting Cd2+ adsorption, as extreme values could otherwise signal potential issues with model fit. Overall, the plot supports that the model provides a good fit without systematic errors, making it appropriate for the data.

3.2.5. Estimation of the Parameters

The p-value analysis shows that certain factors significantly influence the amount of Cd2+ adsorbed by argan nut shells (ArS), specifically the pH (X1), adsorbent mass (X2), initial metal cation concentration (X3), the quadratic term for concentration (X32), and their interaction terms (X1X3, X2X3) (Table 5). Each of these terms has a probability value lower than 5%, indicating a statistically significant impact on adsorption [41].
Factors with p-values above 0.05 are not statistically significant, indicating minimal effect on the response variable. The final model equation (Equation (4)) incorporates only the significant terms, which include the main effects of pH, ArS mass, and [Cd2+]i, in addition to selected interaction and quadratic terms. The coefficient of each term represents its impact on the predicted cadmium adsorption (Y):
Y= 2.154 + 0.698X1 + 0.26X2 + 1.0357X3 − 0.0219X12 − 0.245X22 − 0.517 X32 − 0.0369 X1X2 + 0.344 X1X3 + 0.282 X2X3

3.2.6. Optimization of the Parameters

This study aimed to identify optimal conditions for maximizing the adsorption of cadmium onto raw argan nut shells. To achieve this, we used numerical optimization based on desirability to identify the best combination of pH, adsorbent mass, and initial [Cd2+]. Figure 6 presents the prediction profiler, which identifies the optimal conditions for maximizing cadmium removal. By maximizing the desirability, the optimal parameters were identified as pH = 8, ArS dose = 3.5 g·L−1, and [Cd2+] = 20 mg·L−1. Under these conditions, desirability is close to 1, indicating an optimal solution, with a maximum adsorbed quantity of cadmium of approximately 3.92 mg·g−1. The prediction profiler also shows that all three parameters (pH, adsorbent mass, and cadmium concentration) positively influence Cd2+ adsorption, with higher values generally enhancing the adsorption process.

3.3. Batch Experiments: Kinetic Study

To comprehend the Cd2+ removal kinetics utilizing ArS as a biosorbent, the results were tested with several kinetic models, such as intra-particle diffusion and pseudo-first-order and pseudo-second-order. The study was conducted on ArS under optimal conditions determined by the Box–Behnken design at 25 °C for a duration of 4 h.
Figure 7 illustrates the study of kinetics, showing (a) the adsorption kinetics of Cd2+, (b) the pseudo-first-order model, (c) the pseudo-second-order model, and (d) the intra-particle diffusion model. Analysis of Table 6 and Figure 7 indicates that the pseudo-second-order model most accurately depicts the relationship between t/qt and time. Key parameters, including the coefficient of determination (R2), estimated equilibrium capacity (qe), and rate constant (k2) at [Cd2+]i = 20 mg·L−1, can be found in Table 7. The estimated qe values closely match the practical results, yielding a strong R2 value (≥0.987) [42], indicating that the pseudo-second-order model delivers a precise description of Cd2+ removal by ArS. This implies that the mechanism of adsorption is primarily chemical, involving electron exchange between the adsorbent and adsorbate [43]. In addition, intra-particle transport within the ArS/Cd2+ system was assessed using the Weber–Morris equation. Figure 7d shows that cadmium adsorption on argan shells occurs in four distinct phases. The initial phase, marked by a steep slope, indicates rapid adsorption on the external surface. This is followed by three phases with slower diffusion rates, suggesting that intra-particle diffusion contributes to the process of adsorption [44].
qt and qe represent cadmium quantities adsorbed at time (t) and at equilibrium, respectively (mg·g−1); t represents the contact time in minutes, K1, K2, and Kd are rate constants for pseudo-first-order (min−1), pseudo-second-order (g/mg·min), and intra-particle diffusion (mg/g·min0.5), respectively, and I represents the intercept, indicating the boundary layer impact on the process of adsorption (mg·g−1).

3.4. Batch Experiments: Isotherms

Figure 8 and Table 8 illustrate the adsorption behavior of Cd2+ on ArS and the corresponding isotherm model parameters. The plots in Figure 8 represent the Langmuir (a), Freundlich (b), Temkin (c), and Dubinin–Radushkevich (d) isotherm models. The Langmuir equilibrium constants were calculated through the graphical representation of 1/qe against 1/Ce, according to the linear representation of the model of Langmuir. Since R2 is equal to 0.99, close to 1, the model of Langmuir fits the data quite well in contrast to the Freundlich isotherm (R2 = 0.859), highlighting the accuracy of Langmuir model. The Qads value from the Langmuir model (4 mg·g−1) aligns with the experimental adsorption capacity (3.26 mg·g−1) [46]. The dimensionless constant RL = 0.05 (ranging between 0 and 1) suggests that the adsorption process is favorable [47]. As the concentration increases, RL approaches zero, suggesting the adsorption process is becoming irreversible [48]. The log(qe) versus log(Ce) figure was used to determine the Freundlich constants yielding lower R2 (0.859) than the Langmuir isotherm. The 1/nF value of 0.33 (less than 1) implies typical adsorption behavior under the Freundlich model [49]. Figure 8c illustrates the Temkin isotherm (Qe vs. ln Ce), which describes adsorption on heterogeneous surfaces. It describes the decline in adsorption energy as surface coverage increases and assumes uniform binding energy distribution [50]. The R2 of 0.902 indicates a moderate fit, with constants BT = 0.718 J·mol−1 and KT = 1.432 L/g−1 reflecting the heat of adsorption. The Dubinin–Radushkevich model, as illustrated in Figure 8d, achieves a strong correlation (R2 = 0.984), suggesting a reliable fit with experimental data. The computed energy E = 3.488 KJ·mol−1, below 8 KJ·mol−1, implies that the Cd2+ sorption on ArS is mostly physical [50]. Table 8 summarizes these findings, underscoring the Langmuir model’s superior fit (R2 = 0.99) compared to the other models. The Dubinin–Radushkevich model follows closely with R2 = 0.984, while the Freundlich and Temkin models show moderate correlations (R2 = 0.859 = 0.859 and 0.902, respectively).
Figure 8. Plots of isothermal models by Langmuir (a), Freundlich (b), Temkin (c), as well as Dubinin–Radushkevich (d) for Cd2+ adsorption on ArS.
Figure 8. Plots of isothermal models by Langmuir (a), Freundlich (b), Temkin (c), as well as Dubinin–Radushkevich (d) for Cd2+ adsorption on ArS.
Physchem 05 00013 g008
Table 7. Isotherm parameters.
Table 7. Isotherm parameters.
IsothermsParametersValues
Langmuirqm (mg·g−1)4
KL (L·g−1)0.91
RL0.0521
R20.99
Freundlich [50]1/nf0.33
Kf (mg(1−n)Ln·g−1)1.8516
R20.859
TemkinBT (J·mol−1)0.718
KT (L·g−1)1.432
R20.902
bT (J/molK)1731.13
Dubinin–Radushkevich [51]E (KJ·mol−1)3.488
Qe (mg·g−1)3.525
β (J2/mol2)4.933
R20.984
Table 8. Comparison of cadmium capacities of adsorption on various lignocellulosic adsorbents.
Table 8. Comparison of cadmium capacities of adsorption on various lignocellulosic adsorbents.
SamplesAdsorption Capacity (mg·g−1)pHReference
Pine bark (P. pinaster Ait)3.233.4[14]
Walnut shells4.366[20]
Cashew activated carbon 2.875[18]
Raw walnut shells (RWS)5.386[15]
Coffee beans3.808[16]
Ash produced from rice husks3.046[19]
Rice-hull-derived AC2.494.5[17]
Raw argan shells48This work
Qe and qm represent the quantity of adsorbate adsorbed under equilibrium conditions and the maximum adsorption capacity, respectively, both measured in milligrams per gram (mg·g−1); Ce denotes the equilibrium concentration of the unadsorbed adsorbate in the liquid phase (mg·L−1); T represents the absolute temperature in Kelvin; R denotes the universal gas constant, valued at 8.314 J/mol·K; β represents a constant from the Dubinin–Radushkevich isothermal model that relates to adsorption capacity. It represents the average free energy of adsorption for a single molecule as it moves from an infinite distance to the solid surface. E = 1/√2β is the formula used to determine the free energy. KL, KF, and KT are constants corresponding to the Langmuir (L·mg−1), Freundlich (mg(1−n)Ln·g−1), and Temkin (L·mg−1) models, respectively; RL is the equilibrium factor, often known as the dimensionless separation factor and expressed as RL = 1/(1 + KL·C0) [52]; BT represents the constant related to the heat of adsorption, defined by the equation BT = R·T/KT [52]; bT represents the isothermal constant of Temkin and 1/nf represents the heterogeneity factor. Table 8 compares the adsorption capacities of various lignocellulosic biosorbents, showing that raw argan shells, with a capacity of 4 mg/g at pH 8, performed better than some biosorbents such as cashew activated carbon (CAC) (2.87 mg·g−1) [19] and rice husk ash (3.04 mg·g−1). However, walnut shells (4.36 mg·g−1) and raw walnut shells (5.38 mg·g−1) [15] had higher capacities. This analysis highlights that, while ArS shows significant adsorption potential, certain materials like walnut shells may offer superior performance.

3.5. Batch Experiments: Thermodynamic Parameters

The experimental temperature range for this study was between 25 °C and 65 °C, with 175 mg of ArS used in 50 mL of a 20 mg·L−1 cadmium mixture. Notably, the adsorption capacity of Cd2+ was augmented from 3.45 mg·g−1 at 25 °C to 4.74 mg·g−1 at 65 °C. To assess the spontaneity, feasibility, and nature of the adsorbate and adsorbent, different thermodynamic factors of the reaction, namely, entropy (ΔS°), enthalpy (ΔH°), and Gibbs free energy (ΔG°), were determined by using Equations (5) and (6) [53]:
ln(Kd) = ΔS°/R − ΔH°/RT
ΔGo = −RT ln(Kd)
Using T as the Kelvin temperature and R as the gas constant, Kd may be found using qe/Ce, and ΔH° and ΔS° can be found through the graphical representation of ln(Kd) vs. T−1.
The thermodynamic parameters for Cd2+ adsorption on argan shells (Table 9) reveal key insights. The Gibbs free energy (ΔG°) values shift from positive at lower temperatures (2.26 KJ·mol−1 at 298.15 K) to negative (−0.43 KJ·mol−1 at 338.15 K), showing that, at lower temperatures, the process is not spontaneous but becomes spontaneous as temperature increases, suggesting that higher temperatures favor adsorption. The positive enthalpy (ΔH° = 20.26 KJ·mol−1) shows that the process is endothermic [53], implying that energy input is needed for adsorption, which is consistent with a chemical adsorption mechanism where Cd2+ ions likely form strong interactions with the argan shell surface. The positive entropy (ΔS° = 60.36 J/mol·K) suggests a dissociative mechanism, where the disorder at the solid–liquid interface increases as ions interact with the adsorption sites [54]. Together, these observations indicate that the adsorption of Cd2+ onto ArS is endothermic, likely chemical, and is enhanced by higher temperatures, with a dissociative mechanism at play.

3.6. Desorption Study of Cd2+ from ArS

The desorption study is a crucial step to evaluate the reusability of the adsorbent and the potential for cost-effective and sustainable heavy metal removal in water treatment processes [55]. Cd2+ desorption was carried out under optimum conditions. After equilibration, ArS was filtered and rinsed with 10 mL of distilled water. It was then added to various acid solutions (1 M HCl, 1 M HNO3, and 1 M CH3COOH) to evaluate the efficiency of the different eluents. Metal ion desorption was carried out over a 12 h period. Cd2+ desorption (DP) is calculated as follows:
D P = C d e ( C d 2 + ) C a d ( C d 2 + ) × 100
where Cad represents [Cd2+] adsorbed on the ArS adsorbent and Cde corresponds to [Cd2+] desorbed by the various acid solutions.
The results of Cd2+ desorption are shown in Figure 9. Among the acidic agents tested, HCl proved the most effective, with a desorption rate of 90.02%, followed by HNO3 with 76.65%, and finally CH3COOH with 71.59%. The desorption efficiencies with the different acids can be correlated with their relative acid strengths. HCl, being a strong monoprotic acid, is fully dissociated in solution, generating a high H+ ion concentration, which interferes with the interaction between Cd2+ and the ArS surface. HNO3, another strong acid, was found to have a somewhat lower efficiency, likely due to the difference in ionic interactions with Cd2+. CH3COOH, being a weak acid, might have formed more stable complexes with Cd2+ and thus reduced its desorption efficiency.

3.7. Proposed Adsorption Mechanisms

Effective removal of impurities from aqueous solutions requires understanding the mechanisms underlying solute adsorption on solid surfaces. This adsorption can involve ionic interactions between opposite charges, such as chemical bonding, H bonding, dipole–dipole, dipole-induced dipole, and ion exchange. FTIR is a valuable technique for defining the ArS chemical structure, which primarily consists of cellulose, hemicellulose, lignin, and ash. It has proven effective as a biosorbent for cadmium ions (Cd2+) in solution. FTIR analysis before and after adsorption (Figure 10) reveals significant shifts in the bands of specific functional groups after Cd2+ adsorption, indicating modifications to both the surface characteristics and functional groups of the biosorbent. These shifts likely result from changes in counterions related to hydroxylate and carboxylate anions, suggesting that hydroxyl and carboxyl functional sites are key in Cd2+ biosorption [20]. Experimental findings demonstrate that optimal adsorption of Cd2+ onto the ArS surface occurs at pH equal to 8, greater than the zero point of charge (pHzpc = 3.6) [26]. Above this pH, the ArS surface becomes negatively charged because of the deprotonation of cellulose-OH and lignin-OH sites, facilitated by OH anions in the solution. This negative charge enhances the attraction of Cd2+ cations [56]. Boehm theoretical analysis suggests that acidic groups predominate on the ArS surface compared to basic groups, highlighting oxygen-based reactive sites contributing to the adsorption process. This outcome agrees with the FTIR examination, confirming the roles of hydroxyl and carboxyl groups in Cd2+ biosorption.
Figure 11 illustrates the adsorption mechanism of Cd2+ on cellulose and lignin. In lignin, deprotonated phenolic hydroxides interact with Cd2+, forming stable complexes, while in cellulose, the deprotonation of hydroxyl groups creates active sites that coordinate with Cd2+. This highlights the significant roles of surface functional groups and pH in driving effective adsorption. The FTIR results, which show shifts in functional group bands after Cd2+ adsorption, further support these interactions, with hydrogen bonding and coordination bonds playing central roles.

4. Conclusions

Cd2+ ion removal from aqueous solutions utilizing ArS as an adsorbent was analyzed through various kinetic models, isotherms, and the response surface methodology in conjunction with the Box–Behnken design. This approach helped identify the optimal conditions for adsorption, which involves pH, ArS dosage, and [Cd2+]i, as well as the factors influencing adsorption capacity. The optimization results indicate that, to achieve an adsorbed amount of 3.92 mg, the ideal parameters are a pH of 8, an ArS mass of 3.5 g·L−1, and an initial [Cd2+]i = 20 mg·L−1. The isotherm analysis revealed that the Langmuir model, with an R2 of 0.96, provided the best fit, indicating that adsorption takes place as a single layer on a homogeneous surface. Additionally, the adsorption kinetics demonstrated that metal ion binding happens rapidly in the initial phase, gradually slowing down as intra-particle diffusion takes effect. Furthermore, the pseudo-second-order kinetic model showed remarkable agreement with the findings from the experiment (R2 = 0.987), confirming the model’s suitability for describing the Cd2+ adsorption process.

Author Contributions

Conceptualization, F.-Z.A.; Methodology, F.-Z.A., M.B. (Maria Benbouzid), K.B., H.N., M.B. (Meryem Bensemlali), N.L. and S.E.H.; Software, F.-Z.A., M.B. (Maria Benbouzid), K.B., H.N., M.B. (Meryem Bensemlali), N.L. and S.E.H.; Validation, F.-Z.A., M.B. (Maria Benbouzid), K.B., H.N., M.B. (Meryem Bensemlali), N.L. and S.E.H.; Formal analysis, F.-Z.A., M.B. (Maria Benbouzid), K.B., H.N., M.B. (Meryem Bensemlali), N.L. and S.E.H.; Investigation, F.-Z.A., M.B. (Maria Benbouzid), K.B., H.N., M.B. (Meryem Bensemlali), N.L. and S.E.H.; Resources, F.-Z.A., M.B. (Maria Benbouzid), K.B., H.N., M.B. (Meryem Bensemlali), N.L. and S.E.H.; Data curation, F.-Z.A., M.B. (Maria Benbouzid), K.B., H.N., N.L. and S.E.H.; Writing—original draft, F.-Z.A., M.B. (Maria Benbouzid), H.N., M.B. (Meryem Bensemlali), N.L. and S.E.H.; Writing—review & editing, F.-Z.A., H.N., N.L. and S.E.H.; Visualization, F.-Z.A., H.N., N.L. and S.E.H.; Supervision, N.L. and S.E.H. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to thank all of the project participants “Wastewater reuse in countries with high water stress—challenges and opportunities” for their scientific support and their comments, which significantly improved the manuscript. Financial support from the Swedish Research Council VR (Grant No. 2018-03476), is gratefully acknowledged.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hamid, N.H.A.; Rushdan, A.I.; Nordin, A.H.; Husna, S.M.N.; Faiz Norrrahim, M.N.; Knight, V.F.; Tahir, M.I.H.M.; Li, G.X.; Quan, T.L.; Abdullah, A.M.; et al. A state-of-art review on the sustainable technologies for cadmium removal from wastewater. Water Reuse 2024, 14, 312–341. [Google Scholar] [CrossRef]
  2. Liang, Q.; Luo, H.; Geng, J.; Chen, J. Facile one-pot preparation of nitrogen-doped ultra-light graphene oxide aerogel and its prominent adsorption performance of Cr(VI). Chem. Eng. J. 2018, 338, 62–71. [Google Scholar] [CrossRef]
  3. Fulke, A.B. Understanding heavy metal toxicity: Implications on human health, marine ecosystems and bioremediation strategies. Mar. Pollut. Bull. 2024, 206, 116707. [Google Scholar] [CrossRef]
  4. Guarín-Romero, J.R.; Rodríguez-Estupiñán, P.; Giraldo, L.; Moreno-Piraján, J.C. Simple and Competitive Adsorption Study of Nickel(II) and Chromium(III) on the Surface of the Brown Algae Durvillaea antarctica Biomass. ACS Omega 2019, 4, 18147–18158. [Google Scholar] [CrossRef]
  5. Dhanda, A.; Lamba, R.; Deepika, S.; Prakash, R. Cadmium in Environment—An Overview; Springer: Berlin/Heidelberg, Germany, 2024; pp. 3–20. [Google Scholar] [CrossRef]
  6. Kouzbour, S.; Gourich, B.; Gros, F.; Vial, C.; Allam, F.; Stiriba, Y. Comparative analysis of industrial processes for cadmium removal from phosphoric acid: A review. Hydrometallurgy 2019, 188, 222–247. [Google Scholar] [CrossRef]
  7. Suciu, N.A.; De Vivo, R.; Rizzati, N.; Capri, E. Cd content in phosphate fertilizer: Which potential risk for the environment and human health? Curr. Opin. Environ. Sci. Health 2022, 30, 100392. [Google Scholar] [CrossRef]
  8. Kubier, A.; Wilkin, R.T.; Pichler, T. Cadmium in soils and groundwater: A review. Appl. Geochem. J. Int. Assoc. Geochem. Cosmochem. 2019, 108, 104388. [Google Scholar] [CrossRef]
  9. Sadeek, S.A.; Negm, N.A.; Hefni, H.H.H.; Wahab, M.M.A. Metal adsorption by agricultural biosorbents: Adsorption isotherm, kinetic and biosorbents chemical structures. Int. J. Biol. Macromol. 2015, 81, 400–409. [Google Scholar] [CrossRef]
  10. Godt, J.; Scheidig, F.; Grosse-Siestrup, C.; Esche, V.; Brandenburg, P.; Reich, A.; Groneberg, D.A. The toxicity of cadmium and resulting hazards for human health. J. Occup. Med. Toxicol. 2006, 1, 22. [Google Scholar] [CrossRef]
  11. Rashid, A.; Bhatti, H.N.; Iqbal, M.; Noreen, S. Fungal biomass composite with bentonite efficiency for nickel and zinc adsorption: A mechanistic study. Ecol. Eng. 2016, 91, 459–471. [Google Scholar] [CrossRef]
  12. Qasem, N.; Mohammed, R.; Lawal, D. Removal of heavy metal ions from wastewater: A comprehensive and critical review. NPJ Clean Water 2021, 4, 36. [Google Scholar] [CrossRef]
  13. Nascimento, T.L.; Oliveira, K.F.; Junior, J.O.; Pimenta, A.S.; Melo, D.M.; Melo, M.A.; Braga, R.M. Biosorption of nickel and cadmium using Pachira aquatica Aubl. peel biochar. Sci. Rep. 2024, 14, 5086. [Google Scholar] [CrossRef]
  14. Cutillas-Barreiro, L.; Ansias-Manso, L.; Fernández-Calviño, D.; Arias-Estévez, M.; Nóvoa-Muñoz, J.C.; Fernández-Sanjurjo, M.J.; Álvarez-Rodríguez, E.; Núñez-Delgado, A. Pine bark as bio-adsorbent for Cd, Cu, Ni, Pb and Zn: Batch-type and stirred flow chamber experiments. J. Environ. Manag. 2014, 144, 258–264. [Google Scholar] [CrossRef]
  15. Gondhalekar, S.C.; Shukla, S.R. Biosorption of cadmium metal ions on raw and chemically modified walnut shells. Environ. Prog. Sustain. Energy 2015, 34, 1613–1619. [Google Scholar] [CrossRef]
  16. Kaikake, K.; Hoaki, K.; Sunada, H.; Dhakal, R.; Baba, Y. Removal Characteristics of Metal Ions Using Degreased Coffee Beans: Adsorption Equilibrium of Cadmium(II). Bioresour. Technol. 2007, 98, 2787–2791. [Google Scholar] [CrossRef]
  17. Lee, S.-M.; Lalhmunsiama, C.S.-I.; Tiwari, D. Manganese and iron oxide immobilized activated carbons precursor to dead biomasses in the remediation of cadmium-contaminated waters. Environ. Sci. Pollut. Res. Int. 2013, 20, 7464–7477. [Google Scholar] [CrossRef]
  18. N’goran, P.D.A.; Diabaté, D.; Yao, K.M.; Berenger, K.; Gnonsoro, U.; Kinimo, K.; Trokourey, A. Lead and cadmium removal from natural freshwater using mixed activated carbons from cashew and shea nut shells. Arab. J. Geosci. 2018, 11, 498. [Google Scholar] [CrossRef]
  19. Srivastava, N.K.; Majumder, C.B. Novel biofiltration methods for the treatment of heavy metals from industrial wastewater. J. Hazard. Mater. 2008, 151, 1–8. [Google Scholar] [CrossRef]
  20. Kamsonlian, S.; Suresh, S.; Majumder, C.B.; Chand, S. Biosorption of As(III) from contaminated water onto low cost palm bark biomass. Int. J. Curr. Eng. Technol. 2012, 2, 153–158. [Google Scholar]
  21. Weshahy, A.R.; Sakr, A.K.; Gouda, A.A.; Atia, B.M.; Somaily, H.H.; Hanfi, M.Y.; Gado, M.A. Selective re-covery of cadmium, cobalt, and nickel from spent Ni–Cd batteries using Adogen® 464 and mesoporous silica deriva-tives. Int. J. Mol. Sci. 2022, 23, 8677. [Google Scholar] [CrossRef]
  22. Najah El Idrissi, A.; Benbrahim, M.; Rassai, N. Effect of the particle size argan nut shell (ANS) biomass on combustion parameters in stirling engine in Morocco. Results Eng. 2023, 18, 101202. [Google Scholar] [CrossRef]
  23. Boundati, Y.; Ziat, K.; Naji, A.; Saidi, M.; Rghioui, L. Copper biosorption on argan nut shell: Kinetic and thermodynamic studies. Int. J. Adv. Res. 2018, 6, 300–309. [Google Scholar] [CrossRef]
  24. Millett, M.A.; Baker, A.J.; Satter, L.D. Physical and chemical pretreatments for enhancing cellulose saccharification. Biotechnol. Bioeng. Symp. 1976, 6, 125–153. [Google Scholar]
  25. Benadjemia, M.; Milliere, L.; Reinert, L.; Benderdouche, N.; Duclaux, L. Preparation, characterization and Meth-ylene Blue adsorption of phosphoric acid activated carbons from globe artichoke leaves. Fuel Energy Abstr. 2011, 92, 1203–1212. [Google Scholar] [CrossRef]
  26. Bencheikh, I.; Azoulay, K.; Mabrouki, J.; El Hajjaji, S.; Moufti, A.; Labjar, N. The use and the performance of chemically treated artichoke leaves for textile industrial effluents treatment. Chem. Data Collect. 2021, 31, 100597. [Google Scholar] [CrossRef]
  27. Es-said, A.; Hamdaoui, L.; El Moussaouiti, M.; Bchitou, R. Esterification optimization of cellulose with p-Iodobenzoyl chloride using experimental design method. J. Polym. Res. 2019, 26, 237. [Google Scholar] [CrossRef]
  28. Benallou Benzekri, M.; Benderdouche, N.; Bestani, B.; Douara, N.; Duclaux, L. Valorization of olive stones into a granular activated carbon for the removal of Methylene blue in batch and fixed bed modes. J. Mater. Environ. Sci. 2018, 9, 272–284. [Google Scholar] [CrossRef]
  29. Olabinjo, O.O. Response Surface Techniques as an Inevitable Tool in Optimization Process. In Response Surface Methods—Theory, Applications and Optimization Techniques; IntechOpen: London, UK, 2024. [Google Scholar] [CrossRef]
  30. Duan, R.; Fedler, C.; Jiao, X. Adsorption of pyridine from aqueous solutions onto polyaluminium chloride and anionic polyacrylamide water treatment residuals. Water Sci. Technol. 2021, 83, 1753–1763. [Google Scholar] [CrossRef]
  31. Güray, T.; Menevşe, B.; Yavuz, A.A. Determination of optimization parameters based on the Box-Behnken design for cloud point extraction of quinoline yellow using Brij 58 and application of this method to real samples. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2020, 243, 118800. [Google Scholar] [CrossRef]
  32. Zhuang, J.; Li, M.; Pu, Y.; Ragauskas, A.J.; Yoo, C.G. Observation of potential contaminants in processed biomass using fourier transform infrared spectroscopy. Appl. Sci. 2020, 10, 4345. [Google Scholar] [CrossRef]
  33. Olsson, A.-M.; Salmén, L. The association of water to cellulose and hemicellulose in paper examined by FTIR spectroscopy. Carbohydr. Res. 2024, 339, 813–818. [Google Scholar] [CrossRef]
  34. Paiva, M.; Ammar, I.; Campos, A.R.; Cheikh, R.; Cunha, A. Alfa fibres: Mechanical, morphological and interfacial characterization. Compos. Sci. Technol. 2007, 67, 1132–1138. [Google Scholar] [CrossRef]
  35. Arrakhiz, F.Z.; El Achaby, M.; Malha, M.; Bensalah, M.O.; Fassi-Fehri, O.; Bouhfid, R.; Benmoussa, K.; Qaiss, A. Mechanical and thermal properties of natural fibers reinforced polymer composites: Doum/low density polyethylene. Mater. Des. 2013, 43, 200–205. [Google Scholar] [CrossRef]
  36. Liu, D.; Han, G.; Huang, J.; Zhang, Y. Composition and structure study of natural Nelumbo nucifera fiber. Carbohydr. Polym. 2009, 75, 39–43. [Google Scholar] [CrossRef]
  37. Sen, T.K. Agricultural Solid Wastes Based Adsorbent Materials in the Remediation of Heavy Metal Ions from Water and Wastewater by Adsorption: A Review. Molecules 2023, 28, 5575. [Google Scholar] [CrossRef]
  38. 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]
  39. Benbouzid, M.; Azoulay, K.; Bencheikh, I.; Al-Jadab, N.; Bensemlali, M.; Aarfane, A.; Nasrellah, H.; El Hajjaji, S.; Labjar, N. Evaluation of natural porous material as media filters for domestic wastewater treatment using infiltration percolation process. Euro-Mediterr. J. Environ. Integr. 2024, 1–16. [Google Scholar] [CrossRef]
  40. Tong, K.S.; Kassim, M.J.; Azraa, A. Adsorption of copper ion from its aqueous solution by a novel biosorbent Uncaria gambir: Equilibrium, kinetics, and thermodynamic studies. Chem. Eng. J. 2011, 170, 145–153. [Google Scholar] [CrossRef]
  41. Fadil, M.; Farah, A.; Ihssane, B.; Haloui, T.; Rachiq, S. Optimisation des paramètres influençant l’hydrodistillation de Rosmarinus officinalis L. par la méthodologie de surface de réponse Optimization of parameters influencing the hydrodistillation of Rosmarinus officinalis L. by response surface methodology. J. Mater. Environ. Sci. 2015, 6, 2346–2357. [Google Scholar]
  42. Azoulay, K.; Bencheikh, I.; Mabrouki, J.; Samghouli, N.; Moufti, A.; Dahchour, A.; El Hajjaji, S. Adsorption mecha-nisms of azo dyes binary mixture onto different raw palm wastes. Int. J. Environ. Anal. Chem. 2023, 103, 1633–1652. [Google Scholar] [CrossRef]
  43. Vishan, I.; Saha, B.; Sivaprakasam, S.; Kalamdhad, A. Evaluation of Cd(II) biosorption in aqueous solution by using lyophilized biomass of novel bacterial strain Bacillus badius AK: Biosorption kinetics, thermodynamics and mechanism. Environ. Technol. Innov. 2019, 14, 100323. [Google Scholar] [CrossRef]
  44. Ma, P.; Ma, M.; Wu, J.; Qian, Y.; Wu, D.; Zhang, X. The effect of plastic on performance of activated carbon and study on adsorption of methylene blue. J. Mater. Res. 2019, 34, 3040–3049. [Google Scholar] [CrossRef]
  45. Abdolali, A.; Ngo, H.H.; Guo, W.; Lu, S.; Chen, S.-S.; Nguyen, N.C.; Zhang, X.; Wang, J.; Wu, Y. A breakthrough biosorbent in removing heavy metals: Equilibrium, kinetic, thermodynamic and mechanism analyses in a lab-scale study. Sci. Total Environ. 2016, 542, 603–611. [Google Scholar] [CrossRef]
  46. Putro, J.N.; Kurniawan, A.; Ismadji, S.; Ju, Y.-H. Nanocellulose based biosorbents for wastewater treatment: Study of isotherm, kinetic, thermodynamic and reusability. Environ. Nanotechnol. Monit. Manag. 2017, 8, 134–149. [Google Scholar] [CrossRef]
  47. Kluczka, J.; Pudło, W.; Krukiewicz, K. Boron adsorption removal by commercial and modified activated carbons. Chem. Eng. Res. Des. 2019, 147, 30–42. [Google Scholar] [CrossRef]
  48. Kavakli, C.; Akkaş Kavaklı, P.; Turan, B.; Hamurcu, A.; Güven, O. Quaternized dimethylaminoethyl methacrylate strong base anion exchange fibers for As(V) adsorption. Radiat. Phys. Chem. 2014, 102, 84–95. [Google Scholar] [CrossRef]
  49. Abid, M.; Niazi, N.K.; Bibi, I.; Farooqi, A.; Ok, Y.S.; Kunhikrishnan, A.; Ali, F.; Ali, S.; Igalavithana, A.D.; Arshad, M. Arsenic(V) biosorption by charred orange peel in aqueous environments. Int. J. Phytoremediat. 2016, 18, 442–449. [Google Scholar] [CrossRef]
  50. Foo, K.Y.; Hameed, B.H. Insights into the modeling of adsorption isotherm systems. Chem. Eng. J. 2010, 156, 2–10. [Google Scholar] [CrossRef]
  51. Yang, Z.; Gluesenkamp, K.R.; Frazzica, A. Equilibrium vapor pressure properties for absorbent and adsorbent materials. Int. J. Refrig. 2021, 124, 134–166. [Google Scholar] [CrossRef]
  52. Shehu, A.; Ibrahim, M.B. Equilibrium Adsorption Isotherm of Methylene Blue and Rhodamine B Using Shea Butter Leaves as a Low-Cost Adsorbent. Appl. J. Environ. Eng. Sci. 2023, 9. [Google Scholar] [CrossRef]
  53. Zhang, M.; Wang, H. Raw and calcination-modified coal waste as adsorbents to remove cadmium from simulated mining wastewater. Int. J. Environ. Sci. Technol. 2014, 11, 987–996. [Google Scholar] [CrossRef]
  54. Gupta, A.; Sharma, V.; Mishra, P.K.; Ekielski, A. A Review on Polyacrylonitrile as an Effective and Economic Constituent of Adsorbents for Wastewater Treatment. Molecules 2022, 27, 8689. [Google Scholar] [CrossRef] [PubMed]
  55. Isaac, R.; Siddiqui, S.; Aldosari, O.F.; Uddin, M.K. Magnetic biochar derived from Juglans regia for the adsorption of Cu2+ and Ni2+: Characterization, modelling, optimization, and cost analysis. J. Saudi Chem. Soc. 2023, 27, 101749. [Google Scholar] [CrossRef]
  56. Al-Ghouti, M.A.; Li, J.; Salamh, Y.; Al-Laqtah, N.; Walker, G.; Ahmad, M.N.M. Adsorption mechanisms of re-moving heavy metals and dyes from aqueous solution using date pits solid adsorbent. J. Hazard. Mater. 2010, 176, 510–520. [Google Scholar] [CrossRef] [PubMed]
Figure 1. SEM images of Raw Argan shells and EDX analysis.
Figure 1. SEM images of Raw Argan shells and EDX analysis.
Physchem 05 00013 g001
Figure 2. FTIR spectrum of ArS.
Figure 2. FTIR spectrum of ArS.
Physchem 05 00013 g002
Figure 3. pHzpc determination graph.
Figure 3. pHzpc determination graph.
Physchem 05 00013 g003
Figure 4. Correlation of predicted and observed values of cadmium adsorption.
Figure 4. Correlation of predicted and observed values of cadmium adsorption.
Physchem 05 00013 g004
Figure 5. Residual trace of adsorbed cadmium quantity.
Figure 5. Residual trace of adsorbed cadmium quantity.
Physchem 05 00013 g005
Figure 6. Prediction profiler for optimizing Cd2+ adsorption conditions.
Figure 6. Prediction profiler for optimizing Cd2+ adsorption conditions.
Physchem 05 00013 g006
Figure 7. Kinetic study showing: (a) the adsorption kinetics of Cd2+, (b) the pseudo-first-order model, (c) the pseudo-second-order model, and (d) the intra-particle diffusion model.
Figure 7. Kinetic study showing: (a) the adsorption kinetics of Cd2+, (b) the pseudo-first-order model, (c) the pseudo-second-order model, and (d) the intra-particle diffusion model.
Physchem 05 00013 g007
Figure 9. Cd2+ desorption using different acid solutions (HCl, HNO3, and CH3COOH).
Figure 9. Cd2+ desorption using different acid solutions (HCl, HNO3, and CH3COOH).
Physchem 05 00013 g009
Figure 10. ArS FTIR spectra.
Figure 10. ArS FTIR spectra.
Physchem 05 00013 g010
Figure 11. Adsorption mechanism of Cd2+ on cellulose and lignin.
Figure 11. Adsorption mechanism of Cd2+ on cellulose and lignin.
Physchem 05 00013 g011aPhyschem 05 00013 g011b
Table 1. Selected trial levels applied to cadmium adsorption on ArS.
Table 1. Selected trial levels applied to cadmium adsorption on ArS.
VariablesCode−10+1
[Cd2+]i (mg·L−1)X1512.520
pHX2468
ArS dose (g·L−1)X312.54
With [Cd2+]i = Cd2+ initial concentration.
Table 2. Surface group quantification by Boehm titration of the surface area and pHzpc value.
Table 2. Surface group quantification by Boehm titration of the surface area and pHzpc value.
Carboxylic
Groups (mmol·g−1)
Phenolic
Groups (mmol·g−1)
Lactonic
Groups (mmol·g−1)
Total Acid (mmol·g−1)Total Basic (mmol·g−1)pHzpcSurface
Area (m2·g−1)
ArS0.0250.0150.0930.1330.0853.631.99
Table 3. Experimental design matrix for cadmium adsorption.
Table 3. Experimental design matrix for cadmium adsorption.
ExperimentpHArS Dose (g·L−1)[Cd2+]i (mg·L−1)q(Cd2+)
(mg·g−1)
162.512.52.121
24112.50.951
38412.52.748
461201.756
582.5203.824
642.5201.678
764202.898
842.550.095
96150.449
108112.52.361
1182.550.863
1262.512.51.973
136450.464
1462.512.52.367
154412.51.486
[Cd2+]i is the initial concentration of Cd2+ and q(Cd2+) represents the Cd2+ adsorbed quantity.
Table 4. Variance analysis of the adsorbed amount of cadmium.
Table 4. Variance analysis of the adsorbed amount of cadmium.
Degrees of
Freedom
Sum of SquaresMean SquareF-Ratio
Model914.9672471.6630347.8409
Residual50.1738080.03476Prob. > F
Total1415.141055 0.0003 *
* p < 0.005.
Table 5. Variance analysis of the adsorbed cadmium quantity.
Table 5. Variance analysis of the adsorbed cadmium quantity.
TermEstimationStandard Errort-RatioProb. > |t|
Constant2.15382350.10764420.01<0.0001 *
pH (4.8)0.69822060.06591810.590.0001 *
m (1.4)0.26003680.0659183.940.0109 *
cc (5.20)1.03571320.06591815.71<0.0001 *
pH × m−0.0369850.093222−0.400.7079
pH × cc0.34436760.0932223.690.0141 *
m × cc0.28205880.0932223.030.0292 *
pH × pH−0.0219260.097029−0.230.8302
m × m−0.2452060.097029−2.530.0527
cc × cc−0.5167060.097029−5.330.0031 *
Table 6. Parameters of the kinetic models.
Table 6. Parameters of the kinetic models.
Kinetic ModelParametersArS/Cd2+
Pseudo-first-order [45] (5)R20.171
k1 (min−1)0.373
qetheo (mg·g−1)2.30
qeexp (mg·g−1)4.73
Pseudo-second-order [42]R20.987
k2 (g/mg·min)−0.065
qetheo (mg·g−1)4.166
qeexp (mg·g−1)4.73
Intra-particle diffusion [44]Type IR21
kd (mg·g1 min0.5)1.14
I−0.335
Type IIR20.942
kd (mg·g1 min0.5)0.003
I4.698
Type IIIR21
kd (mg/g1 min0.5)0.025
I3.413
Type IVR20.566
kd (mg/g1 min0.5)−0.014
I5.49
Table 9. Cd2+ adsorption thermodynamics on ArS.
Table 9. Cd2+ adsorption thermodynamics on ArS.
TΔG° (KJ·mol−1)ΔH° (KJ·mol−1)ΔS° (J/mol·K)R2
298.152.2620.2660.360.913
308.151.43
323.151.24
338.15−0.43
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

Abahdou, F.-Z.; Benbouzid, M.; Bouiti, K.; Nasrellah, H.; Bensemlali, M.; Labjar, N.; El Hajjaji, S. Enhanced Cadmium Removal by Raw Argan Shell Adsorbent: Experimental and Theoretical Investigations for Ecological Applications. Physchem 2025, 5, 13. https://doi.org/10.3390/physchem5010013

AMA Style

Abahdou F-Z, Benbouzid M, Bouiti K, Nasrellah H, Bensemlali M, Labjar N, El Hajjaji S. Enhanced Cadmium Removal by Raw Argan Shell Adsorbent: Experimental and Theoretical Investigations for Ecological Applications. Physchem. 2025; 5(1):13. https://doi.org/10.3390/physchem5010013

Chicago/Turabian Style

Abahdou, Fatima-Zahra, Maria Benbouzid, Khalid Bouiti, Hamid Nasrellah, Meryem Bensemlali, Najoua Labjar, and Souad El Hajjaji. 2025. "Enhanced Cadmium Removal by Raw Argan Shell Adsorbent: Experimental and Theoretical Investigations for Ecological Applications" Physchem 5, no. 1: 13. https://doi.org/10.3390/physchem5010013

APA Style

Abahdou, F.-Z., Benbouzid, M., Bouiti, K., Nasrellah, H., Bensemlali, M., Labjar, N., & El Hajjaji, S. (2025). Enhanced Cadmium Removal by Raw Argan Shell Adsorbent: Experimental and Theoretical Investigations for Ecological Applications. Physchem, 5(1), 13. https://doi.org/10.3390/physchem5010013

Article Metrics

Back to TopTop