Next Article in Journal
Unmodified Plant and Waste Oils as Functional Additives in PU Flooring Adhesives: A Comparative Study
Previous Article in Journal
Chemical Composition and Bioactivity of Essential Oils from Magnolia pugana, an Endemic Mexican Magnoliaceae Species
Previous Article in Special Issue
Cold Plasma-Induced Changes in Polyethylene Particles and Their Binding Affinity to Selected Pharmaceuticals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Removal of Oxyanions and Trace Metals from River Water Samples Using Magnetic Biopolymer/Halloysite Nanocomposites

by
Nyeleti Bridget Mabaso
1,2,
Philiswa Nosizo Nomngongo
1,2 and
Luthando Nyaba
1,2,*
1
Department of Chemical Sciences, University of Johannesburg, Doornfontein Campus, Johannesburg 2028, South Africa
2
Department of Science and Innovation-National Research Foundation South African Research Chair Initiative (DSI-NRF SARChI): Nanotechnology for Water, University of Johannesburg, Doornfontein 2028, South Africa
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(18), 3777; https://doi.org/10.3390/molecules30183777
Submission received: 12 June 2025 / Revised: 3 September 2025 / Accepted: 9 September 2025 / Published: 17 September 2025
(This article belongs to the Special Issue Green Chemistry Approaches to Analysis and Environmental Remediation)

Abstract

The presence of metallic pollutants presents a significant risk to human health, making their removal crucial. Magnetic halloysite nanotube (HNT@Fe3O4) nanocomposite was synthesised via co-precipitation, and then magnetic hydrogel (Fe3O4@HNT-SA and Fe3O4@HNT-CTS) nanocomposites were prepared using chitosan (CTS) and sodium alginate (SA) biopolymers. The structural, morphological, crystalline, surface, and thermal properties of the hydrogels were determined. The favourable adsorption performance of Fe3O4@HNT-SA and Fe3O4@HNT-CTS hydrogels towards As, Cd, Cr, Mo, Pb, Sb and V was established by optimising the factors affecting the sorption process. The results indicated that Fe3O4@HNT-CTS was suitable for the adsorption of As, Cr, Mo, Sb and V, while Fe3O4@HNT-SA had high adsorption affinity for Cd and Pb. The data for the adsorption of target analytes onto the hydrogels were mostly explained by both the Langmuir isotherm model and the pseudo-second order model. The maximum adsorption capacities of Fe3O4@HNT-SA hydrogel for Cd and Pb were 52.2 mg/g and 57.7 mg/g, respectively. On the other hand, the maximum capacities of the Fe3O4@HNT-CTS hydrogel for As, Cr, Mo, Sb, and V were 30.3 mg/g, 28.4 mg/g, 22.2 mg/g, 24.7 mg/g, and 19.9 mg/g, respectively. The Fe3O4@HNT-SA and Fe3O4@HNT-CTS hydrogels effectively removed the respective target analytes from river water samples.

Graphical Abstract

1. Introduction

The introduction of metal/metalloid contaminants poses significant risks to both human and ecological health. These elements concentrate due to human activities, such as industrial operations, mining, and agricultural practices, leading to water contamination [1]. Elements such as As, Cr, Mo, Sb, Se and V, collectively known as oxyanions, are highly water-soluble, increasing the potential for accumulation in vulnerable organisms and ecological systems, and they lead to serious public health effects [2]. While trace metals, such as Cd and Pb, are recognised for their toxicity and carcinogenic properties [3]. Therefore, developing a method to remove these metals and metalloids from the environment is essential.
Some of the methods used for removing metals include chemical precipitation [4], ionic exchange [5], reversed osmosis [6], and adsorption [7]. Adsorption is regarded as one of the most effective methods for removing metallic pollutants, offering simple operation, high efficiency, and low cost [8]. As a result, it is commonly used to treat metallic pollutants in water. Metal removal efficiency depends on the properties of the adsorbent, including regenerability, accessibility, surface area and adsorption capacity [9]. Various adsorbents are employed to remove metallic contaminants from aqueous solutions. These include Zr-MOF [10], ZnO-GO-NiO [11], biochars [12], GO-Fe3O4 [13], nanobentonite [14], and MnO2 [15]. However, many of these adsorbents present challenges such as high costs, toxicity, low surface area, and difficulties in regeneration. Natural clays and biopolymers are particularly interesting due to their availability, low cost, and lower toxicity than other materials [16].
Sodium alginate (SA) is a naturally occurring biopolymer extracted from brown seaweed, composed of β-D-mannuronic acid and guluronic acid as its monomeric components, as detailed by [17]. It is characterised by its carboxyl and hydroxyl functional groups, which offer distinct adsorption properties, alongside its notable biocompatibility and biodegradability, as highlighted by [18]. Due to its renewable nature and availability at an economical price point, alginate presents a sustainable option. It can undergo modifications, oxidation, and esterification to enhance its properties. Nonetheless, its application is somewhat hindered by its poor mechanical and chemical stability [19].
Whereas chitosan (CTS), a polymer obtained by deacetylating chitin present in the shells of crustaceans, is distinguished by its unique qualities, making it invaluable in various fields [20]. Its ability to remove heavy metals via ion exchange and binding processes is due to its structure’s amino and hydroxyl functional groups [21]. Its appeal is further brought out by its affordability and its low-toxicity nature [22]. Nevertheless, it is prone to breakdown in acidic settings, poor solubility, and a minimal surface area limit its broader usage [23]. Various modifications of chitosan have been investigated to overcome these challenges. These include modifying metal–organic frameworks with chitosan and citric acid for Cr (VI) removal [24]. Chitosan, a polyelectrolyte, together with carboxymethylcellulose, was synthesised to eliminate dyes and heavy metals from water [25].
Furthermore, incorporating halloysite nanotubes (HNT) into CTS and SA enhances their mechanical properties and thermal stability while adding new functionalities. The tubular structure of HNTs will capture and hold ions inside and on their surface, which is beneficial for ion adsorption. Thus, this research seeks to create Fe3O4@HNT-SA and Fe3O4@HNT-CTS hydrogel nanocomposites for As, Cd, Cr, Mo, Pb, Sb and V adsorption. The central composite design was used to determine the key factors influencing the adsorption efficiency of the hydrogels. The incorporation of CTS and SA to Fe3O4@HNT revealed the change in surface characteristics of the magnetic hydrogels that led to different adsorption affinities toward the target analytes. Adsorption kinetic and isotherm models, such as Langmuir, Freundlich, pseudo-first order and pseudo-second order, were used to elucidate the adsorption mechanism. Furthermore, the possibility of using Fe3O4@HNT-SA and Fe3O4@HNT-CTS in the removal of As, Cd, Cr, Mo, Pb, Sb and V from river water was investigated.

2. Results and Discussion

2.1. Characterisation of the Hydrogel Materials

2.1.1. Fourier-Transform Infrared Spectroscopy (FTIR)

Figure 1 shows FTIR spectra of SA, HNT, CTS, Fe3O4, Fe3O4@HNT-SA and Fe3O4@HNT-CTS. In the spectrum of HNT represented by Figure 1, the peaks at 3699 and 3619 cm−1 resulted from the internal aluminium surface (Al-OH). The peak of Si-O is at 1033 and 1025 cm−1 for HNT. The peaks at 912 and 914 cm−1 could be attributed to silica, corresponding to the asymmetrical stretching vibration of Si-O-Si. Peaks at 518 cm−1 correspond to Si-O-Al. The peaks of HNT identified are similar to the peaks identified by [26]. The FTIR spectrum of CTS is represented in Figure 1A. The peaks at 3093–3345 cm−1 correspond to N-H and O-H stretching. The peak at 2871 cm−1 can be attributed to C-H asymmetry; similar peaks of CTS were also observed by [27]. The 1638 and 1039 cm−1 peaks are attributed to -CO and C-O; they were also observed by [28]. SA spectrum is indicated in Figure 1B. The broad peak at 3404 cm−1 corresponds to stretching vibrations of hydroxyl (-OH) groups, the peak at 1601 cm−1 corresponds to C=O, and the peak at 1419 cm−1 corresponds to the symmetric stretching of the carboxylate groups (-COO), as observed also by [29]. The Fe3O4 spectrum is represented in Figure 1. The peak at 584 cm−1 is attributed to Fe-O, while the peak at 3481 cm−1 is attributed to -OH. Figure 1A represents the spectrum of Fe3O4@HNT-CTS. In this spectrum, there is a peak at 3470 cm−1 due to the –OH/-NH stretch. An overlap in -OH/-NH stretching vibrations with HNT’s Si-O stretch indicates interaction between the two materials, and there are shifts in the typical band positions due to hydrogen bonding between HNT, Fe3O4, and CTS. The peak at 565 cm−1 results from Fe-O bond stretching, which results from Fe3O4 incorporation. Si-O is at 1044 cm−1, confirming the presence of HNT in the Fe3O4@HNT-CTS composite. The peak at 1643 cm−1 may be due to OH from Fe3O4. The spectrum of Fe3O4@HNT-SA is represented in Figure 1B. The presence of SA in combination with HNT likely shifts or overlaps the hydroxyl and carbonyl bands; this phenomenon can be seen at 3272–3711 cm−1. The peak at 1622 cm−1 (C=O stretching) overlapped with HNT peaks and OH from Fe3O4. The peak at 1032 cm−1 may be attributed to Si-O of HNT. The peak at 585 cm−1 may be attributed to the Fe-O bond stretching of Fe3O4, and the presence of peaks from Fe3O4, SA, and HNT confirms the nanocomposite formation.

2.1.2. Transmission Electron Microscopy (TEM)

Figure 2 shows the TEM images of SA, CTS, Fe3O4@HNT-CTS and Fe3O4@HNT-SA. Pure HNT, indicated by Figure S1, has a hollow, tubular shape, with tubes overlapping each other. while Figure S1B shows a spherical shape of magnetite. Figure 2A shows that CTS particles were clumped together. The nanocomposite Fe3O4@HNT-CTS in Figure 2B shows that the magnetic nanoparticles were attached to the exterior and the interior of HNT. Figure 2C reveals the selected area of electron diffraction of Fe3O4@HNT-CTS; the figure shows six rings corresponding to Miller indexes of (220), (311), (400), (422), (511) and (440), which agree with the XRD results and suggest that the nanocomposite was coated with Fe3O4. SA is shown in Figure 2D, which shows an irregularly shaped porous material. In the case of Fe3O4@HNT-SA demonstrated in Figure 2E, it is observed that the HNT is coated with Fe3O4, with most of the magnetic nanoparticles on the side of the tube forming a chain-like structure. Furthermore, some magnetic nanoparticles form inside the HNT. Figure 2F for SA-HNT@Fe3O4, the six rings correspond to Miller indexes (220), (311), (400), (422), (511) and (440), which also agrees with the XRD results that suggest the coating with Fe3O4.

2.1.3. Scanning Electron Microscopy (SEM)

Figure 3 shows the morphology of SA, CTS, Fe3O4@HNT-CTS and Fe3O4@HNT-SA, whereas the SEM images of HNT and Fe3O4 are displayed in Figure S2A,B. Figure 3A shows the morphology of SA. The nanocomposite of Fe3O4@HNT-SA shown in Figure 3B shows the spherical shape of SA and Fe3O4, which were embedded in the HNT tube. EDS of SA-HNT@Fe3O4 shown in Figure 3C reveals the presence of elements such as Si, Al, Fe, Cl, O, and C, corresponding to the elemental composition of HNT, SA and Fe3O4 precursors, thereby confirming the synthesis of the nanocomposite. Figure 3D shows the morphology of CTS. The SEM image of Fe3O4@HNT-CTS, as represented in Figure 3E, shows the nanotubes coated with spherical nanoparticles of CTS and Fe3O4. In Figure 3F, the EDS analysis of Fe3O4@HNT-CTS identifies the elemental compositions of Fe, Al, Si, N, and C from CTS, Fe3O4, and HNT, further confirming the formation of the composite.

2.1.4. Brunauer–Emmett–Teller (BET)

The N2 adsorption–desorption analysis was used to assess the surface area, pore volume, pore size distribution, and average pore size of SA, HNT, CTS, Fe3O4, Fe3O4@HNT-SA and Fe3O4@HNT-CTS. Figure 4 shows that the materials HNT exhibit type Iva with an H3 hysteresis loop, whereas Fe3O4, CTS, Fe3O4@HNT-SA and Fe3O4@HNT-CTS exhibit type Iva and H1 hysteresis loops [30], which are characteristics of the presence of mesopores. Fe3O4, at the same time, is an H1 loop. The surface area of the polymers without modification is small due to their polymeric nature (for CTS = 1.60, SA = 3.82), as shown in Table 1. Incorporating these polymers with HNT and Fe3O4 brings additional surface area and functional groups. The surface area of the nanocomposite was (Fe3O4@HNT-CTS = 58, Fe3O4@HNT-SA = 41). Also, the pore volume of Fe3O4@HNT-SA and Fe3O4@HNT-CTS increased while the average pore size decreased due to the formation of new pores.

2.1.5. X-Ray Diffraction Analysis (XRD)

X-Ray diffraction analysis (XRD) was used to identify HNT, CTS, Fe3O4, SA, Fe3O4@HNT-SA and Fe3O4@HNT-CTS nanocrystalline structures. It is evident from Figure 5A,B that the diffractogram for HNTs shows reflection peaks at various diffraction angles, including 11.89°, 19.82 °, 24.87 °, 36.54°, and 38.04°, which were indexed to 001, 110, 002, 200, and 131, respectively, agreeing with what was reported by [31]. The diffractogram of Fe3O4 in Figure 5A,B has peaks including 29.69°, 34.87°, 42.47°, 56,47°, and 62.96°, which are indexed to (220), (311), (400), (422), (511) and (440), respectively, and correspond with what was reported by [32]. The peak characteristics of CTS, as indicated by Figure 5A at 2θ, were 4.3°,10.16° (020), and 19.88° (200), which were consistent with the literature [33]. The characteristic peaks of SA, as shown in Figure 5B, were 13.25° and 23.62°, which are indexed to (100) and (200), respectively; similar peaks were observed by [34]. The diffractogram of Fe3O4@HNT-CTS is shown in Figure 5A. The spectra show the peaks at 30.03°, 43.07°,57.38° and 63.06° resulting from Fe3O4 with a slight shift and the peak at 36.54° of HNT was overlapped by the Fe3O4 peak at 36.54°, and this confirms successful incorporation. There was a peak at 26.49°, which confirms the presence of HNT in the Fe3O4@HNT-CTS. The diffractogram of Fe3O4@HNT-SA is displayed in Figure 5B. The figure illustrates that Fe3O4 overlapped with the weaker peaks from SA or HNT, especially at similar 2θ values, confirming the bonding between Fe3O4 and SA and HNT.

2.1.6. Zeta Potential

The point of zero charge (pHpzc) was used to determine the surface charge of the adsorbent, which influences the adsorption of Cd and Pb on Fe3O4@HNT-SA and the adsorption of As, Cr, Mo, Sb and V using Fe3O4@HNT-CTS. As shown in Figure 6A, the zeta potential of Fe3O4@HNT-SA indicates that the adsorbent remains negatively charged across a range of pH levels. This behaviour is associated with the deprotonation of hydroxyl and carboxyl groups in Fe3O4@HNT-SA. The deprotonation increases as the pH increases, resulting in a higher concentration of negatively charged carboxylate ions and hydroxyl groups, which leads to an overall negative zeta potential for Fe3O4@HNT-SA. Figure 6B shows that the zeta potential decreased with the increase in the pH of the solution. The zeta potential was positive at 2–6, and the point of zero charge was at a pH of 7. At lower pH, the protonation of NH2 and hydroxyl groups results in the charge of Fe3O4@HNT-CTS becoming positive [35]. These amine groups lose protons at a higher pH above 7, producing a neutral and negative charge.

2.2. Adsorbent Selection

The suitability of Fe3O4@HNT-SA and Fe3O4@HNT-CTS hydrogels was evaluated for the adsorptive removal of As, Cd, Cr, Mo, Pb, Sb, and V. As seen in Figure 7, the removal efficiency of Fe3O4@HNT-CTS performed better for the removal of As, Cr, Mo, Sb, and V compared to Fe3O4@HNT-SA. Consequently, Fe3O4@HNT-SA had higher adsorption efficiencies for Cd and Pb compared to Fe3O4@HNT-CTS. The results also show that the Fe3O4@HNT-SA has some affinity for Cr and V, but its adsorption performance does not surpass that of Fe3O4@HNT-CTS. The variation in the %RE of these Fe3O4@HNT-SA and Fe3O4@HNT-CTS hydrogels arises from their distinct surface properties. The pH of the solution also influenced their different performances. At the pH of 5.5, the speciation of the oxyanions is as follows: vanadate (H2VO4) [36], molybdate (HMoO4) [37], arsenous acid (H3AsO3) [38], and chromium (Cr3+) [39], Sb(OH)3. The functional groups (nitrogen, carboxyl and hydroxyl) on the Fe3O4@HNT-CTS surface at a pH of 5.5 can remove positively charged metals such as Cr3+ through complexation or chelation mechanisms. The species of V and Mo are negatively charged at pH 5.5 and, therefore, most likely to be adsorbed onto the positively charged surface of Fe3O4@HNT-CTS.
Meanwhile, the neutral species As and Sb interact with the protonated functional groups on the adsorbent through their oxygen lone pairs of electrons. The adsorption of Cd and Pb on Fe3O4@HNT-CTS was minimal due to the electrostatic repulsion between the positively charged absorbent and cations. The minimal adsorption observed might be because the surface area of Fe3O4@HNT-CTS was higher than that of Fe3O4@HNT-SA, resulting in higher surface activity. The negative surface of Fe3O4@HNT-SA repels (H2VO4) and HMoO4 and reduces their interaction with the material. In contrast, Cr demonstrated a higher removal percentage because it is positively charged around a pH of 5.5 and can interact with the deprotonated groups, hydroxyl and carboxyl groups. The two species (H3AsO3) and Sb(OH)3 are neutral species and lack the charge necessary for effective interaction with the deprotonated hydroxyl and carboxyl groups. The high concentration of hydroxyl and carboxyl groups can attract cations through electrostatic forces. This adsorbent was found to be more suitable for the adsorption of Cd and Pb, which benefited from the negative charge.

2.3. Optimisation of Batch Adsorption Experiments Using Central Composite Design

Central composite design (CCD) optimisation modelling was employed to identify the factors influencing the removal of Cd and Pb using Fe3O4@-SA-HNT, as well as the removal of As, Cr, Mo, Sb, and V using Fe3O4@CTS-HNT. The Analysis of variance (ANOVA) results are illustrated in the Pareto chart shown in Figure 8. This figure indicates that pH and MA were statistically significant for both adsorbents, as evidenced by their bars exceeding the vertical line. The pH is crucial due to its influence on the surface charge of the adsorbent and the chemical species of metals and metalloids present. MA is also essential since it determines the number of active sites available for reaction with the analytes.
Figure 9 presents the combined effect of sample pH and MA on Cd and Pb adsorption onto Fe3O4@HNT-SA adsorbent. Figure 9A illustrates that the adsorption of Cd and Pb onto the Fe3O4@HNT-SA surface increased with rising sample pH. The pH significantly influences the adsorption of Cd and Pb. At lower pH levels, the concentration of H+ is high, which leads to competition with the positively charged Cd and Pb ions, resulting in decreased removal efficiency. However, as the pH increases, %R also increases due to the higher availability of hydroxyl and carboxyl groups that facilitate the removal of Cd and Pb. Figure 9A also indicates that MA had no effect, but high MA increases the sites available for Cd and Pb adsorption [40].
Figure 9B presents the combined effect of sample pH and MA for the adsorption of As, Cr, Mo, Sb and V onto Fe3O4@HNT-CTS. The figure shows that when the pH is increased, the % R also increases. This is because the pH also determines the existence of As, Cr, Mo, Sb and V, and some exist as anionic species (H2VO4, HMoO4) at higher pHs, which are more easily adsorbed at acidic pH. The H+ decreases with increased pH, so there will be less competition between H+ and positively charged species (Cr3+). The -OH, -NH2 and Si-OH of Fe3O4@HNT-CTS deprotonate at higher pH, enabling electrostatic interaction between the ions and these functional groups. The adsorbent dosage did not affect the %R of As, Cr, Mo, Sb and V.
Figure 10 illustrates the evaluation of the adsorption of Cd, Pb, V, Cr, Mo, As, and Sb ions using desirability functions. The desirability function for Fe3O4@HNT-SA, with a score of 1, had a maximum removal of 77%, and the lowest removal (5.0%) was scored 0. The intermediate was 0.5, with a removal percentage of 40.9%. The desirability function of Fe3O4@HNT-CTS scored 1 with a maximum removal of 83%, and the lowest removal (23%) was scored 0, with the intermediate being 0.5, which had a removal percentage of 52.7. A desirability score of 1 was targeted to achieve the best removal conditions. In comparison, the desirability function of Fe3O4@HNT-SA achieved a maximum removal efficiency (%RE) of 76.8 at a score of 1. At the intermediate score of 0.5, the %RE was 40.89, while the lowest score of 0 corresponded to a %RE of 5.01. Therefore, a desirability score of 1 was selected to obtain the maximum %RE. The Experimental validation with six replicates was performed using optimum conditions (pH = 5.5, MA = 40 mg and ET = 30 min) for the removal of As, Cr, Mo, Sb and V using Fe3O4@HNT-CTS yielded 70.2–93.3%RE While the optimum conditions (pH = 6.5–7.5, MA = 42mg and ET = 30 min) for the removal of Cd and Pb yielded an average %RE of 90.5–99.7%, demonstrating the surface methodology’s accuracy in determining optimal conditions.

2.4. Adsorption Studies

2.4.1. Adsorption Isotherms Studies

The equilibrium isotherm studies investigated the interaction between the Fe3O4@HNT-SA and Fe3O4@HNT-CTS adsorbents and their respective adsorbates. The relationship between Cd, Pb and the Fe3O4@HNT-SA is illustrated in Figure 11A. In contrast, the relationship between As, Cr, Mo, Sb and V with the Fe3O4@HNT-CTS is shown in Figure 11B. As depicted in both figures, the adsorption capacity of the analytes increases with higher initial metal ion concentrations. The presence of active sites on the surface of the Fe3O4@HNT and Fe3O4@HNT-CTS increased the amount adsorbed (qe) at lower initial concentrations. However, the active sites reach equilibrium as the initial concentrations rise due to the saturation of the binding sites. The data were analysed using Langmuir and the Freundlich isotherm models, and their parameters are summarised in Table 2.
The regression correlation coefficients and isotherm model parameters were calculated using linearised Langmuir and Freundlich expressions from Figures S3 and S4, and the parameters are shown in Table 2.
Langmuir mode1:
C e q e = 1 K L q m a x + C e q m a x
where qmax is the maximum adsorption capacity (mg/g), and RL is the Langmuir constant (L/mg).
R L = 1 1 + K L c o
Freundlich model:
l n q e = l n k f + 1 n l n c e
Kf (mg/g) is the Freundlich constant, and 1/n is the heterogeneity factor.
The R2 value was used to determine which model provided a better fit. The R2 values for As, Cr, Sb, Mo, and V were high for the Langmuir model, ranging from 0.96 to 0.99. This suggests that the adsorption processes occurred on homogeneous sites of Fe3O4@HNT-CTS [3]. In contrast, Cr showed high R2 (0.94) for the Freundlich model instead, indicating multilayered adsorption [41]. The adsorption of Cd by Fe3O4@HNT-SA exhibited a high R2 (0.98) for the Langmuir model, indicating its adsorption occurred on a uniform surface. in contrast, Pb displayed a similarly high R2 value (0.98) for the Freundlich model, suggesting that the adsorption happened on a heterogeneous surface [41].
The maximum capacities of As, Cd, Mo, Pb, Sb and V were closer to the experimental e, suggesting that the Langmuir model explained the adsorption data better. The parameter RL was used to evaluate the nature of the adsorption process, categorising it as unfavourable (RL > 1), linear (RL = 1), favourable (0 < RL < 1), or irreversible (RL = 0) [42]. All analytes measured had RL values that were less than one but greater than zero, indicating that their adsorption was favourable. The Freundlich constant (KF) reflects the degree of affinity between the analytes and the adsorbents, with higher KF values signifying a stronger affinity [43]. The order of affinity towards Fe3O4@HNT-CTS was found to be: Cr > Sb > As > V > Mo. In contrast, Pb exhibited a greater affinity for Fe3O4@HNT-SA compared to Cd.

2.4.2. Adsorption Kinetics Studies

Adsorption kinetics is a key measurement that evaluates the adsorbent’s effectiveness in the adsorption rate. This rate is significantly influenced by the diffusion process and the contact time of the adsorbate at the solid-solution interface [44]. Figure 12 presents the relationship between adsorption capacity and contact time. Figure 12A shows that between 0 and 40 min, Fe3O4@HNT-CTS provided sufficient active sites for interactions, facilitating rapid adsorption of As, Cr, Mo, Sb, and V. Then, after 40 min, the adsorptive sites were saturated, and equilibrium was reached. In Figure 12B, it is observed that Cd and Pb reached equilibrium within 40 min as well.
The pseudo-first-order and pseudo-second-order models were utilised, and the findings are displayed in Table 3. The parameters were calculated from Figures S6 and S5.
Pseudo-first order:
ln q e q t = ln ( K 1 q e ) K 1 t
Pseudo-second order:
t q t = 1 K 2 q e 2 + 1 q e
According to Table 3, all the analytes had the highest correlation coefficient (R2) for the pseudo-second-order model, and As fitted the pseudo-first-order model better. The pseudo-second-order model calculated qe values for As, Cd, Cr, Mo, Pb, Sb, and V were closer to the experimental qe values.
Also, similarly, As and Pb adsorption can be described by the pseudo-second-order model, as the R2 values are greater than 0.99.

2.4.3. Thermodynamics

The effect of temperature on the removal efficiency of Cd and Pb using Fe3O4@HNT-SA, and As, Cr, Mo, Sb, and V using Fe3O4@HNT-CTS, was investigated. The thermodynamic parameters were evaluated under optimum adsorption conditions, considering both equilibrium concentration and equilibrium time. The temperature range studied was 290–313 K. Thermodynamic parameters, including Gibbs free energy (ΔG°), enthalpy (ΔH°), and entropy (ΔS°), were calculated using Equations (6) and (7).
G ° = R T l n K
l n K = S ° / R D H ° / R T
K = Q e / C e
The universal gas constant is (8.314 J mol−1 K−1), while T refers to the absolute temperature in kelvins (K), and Kc denotes the equilibrium constant associated with the adsorption process.
The enthalpy change (ΔH°) for Pb, As, Cr, Mo, and V with their respective adsorbents, as shown in Table 4, was positive, indicating that the adsorption process was endothermic. Therefore, increasing the temperature enhanced the adsorption of these analytes. In contrast, Cd and possibly V adsorption was exothermic (ΔH° < 0), meaning lower temperatures favoured their adsorption. The entropy change (ΔS°) for Pb, As, Cr, Mo, and V was positive, suggesting increased randomness at the solid–solution interface and their spontaneous adsorption processes. In contrast, Cd and V exhibited negative ΔS° values, implying a decrease in randomness and less favourable adsorption behaviour, and the reaction needed energy for initiation [44]. The values of ΔG° for all analytes were less than 20 kJ/mol, showing that the analytes were physically adsorbed [45].

2.5. Interference Studies

The interference studies were conducted on Cd and Pb using Fe3O4@HNT-SA, and on As, Cr, Mo, Sb, and V using Fe3O4@HNT-CTS, using optimum conditions. The goal was to evaluate whether the coexisting species commonly found in water could influence the adsorption of the target analytes. To carry out the experiments, 20 mg/L concentrations of each interfering ion were spiked or added to the standard solutions of the target ions mentioned above. In Table S1, it is shown that the other ions (Mg, Na, Cr) have no noticeable impact (above 5%) on the removal of Cd and Pb. However, when zinc was added, there was a significant decrease in the percentage removal of these metals, dropping from 95.89% to 84.97%. This decline is attributed to Zn, which has a comparable ionic radius and also has a good affinity for the OH group on the surface of the Fe3O4@HNT-SA, which, as a result, competes with the target analytes for the active sites [46]. The interference study for As, Cr, Mo, Sb, and V onto Fe3O4@HNT-CTS was carried out with interference ions Zn, Mg, Na, Pb, Cd, and Fe using the optimum conditions. The interference ions of concentration of 20 ppm were each added to standard solutions containing the target ions. The increase in the concentrations of these interfering ions had a slight impact on the percentage removal of the target analytes. It was found that the investigated interferences did not have a significant effect, as the difference was less than 5% as indicated by Table S2. These finding was further supported by the analysis of real river water samples, where the removal percentage was not affected by the presence of possible interference studied above.

2.6. Reusability Studies

The efficacy of the two adsorbents was evaluated for reuse to determine if there was a significant change in the removal percentage after multiple uses. As shown in Figure 13A, Fe3O4@HNT-SA was reused five times for the removal process without a significant change in effectiveness. In contrast, Figure 13B indicates that Fe3O4@HNT-CTS had a decrease in %removal for As, Mo and V, with a difference of less than 5%. For chromium, the removal efficiency declined from 91.59% in the first cycle to 76.1% in the last cycle, showing a significant decrease. Similarly, antimony also experienced a noticeable decline, dropping from 74.65% to 65.26% by the fifth cycle. This decline may be attributed to the adsorbent containing a small portion of these two analytes after each adsorption/desorption cycle, which likely led to the blockage of active sites after each successive cycle. Therefore, it can be concluded that both Fe3O4@HNT-SA and Fe3O4@HNT-CTS can be effectively used for up to five cycles while achieving acceptable removal percentages. This finding also suggests that the materials are cost-effective.

2.7. Application in Real Water Samples

The applicability of the prepared adsorbents Fe3O4@HNT-CTS and Fe3O4@HNT-SA was assessed by analysing five samples of water from the river spiked with known concentrations of the target analytes. The results showed that the removal efficiencies of Fe3O4@HNT-CTS and Fe3O4@HNT-SA effectively removed the target analytes. The concentration in the final solution was not detected, suggesting that the Fe3O4@HNT-CTS and Fe3O4@HNT-SA are suitable adsorbents for treating contaminated surface water. The concentrations were reduced to levels that are below the World Health Organisation (WHO) permissible limits for various contaminants in drinking water [47,48,49,50,51,52].

2.8. Comparison Studies

Table 5 summarises a comparison of Fe3O4@HNT-CTS and Fe3O4@HNT-SA with other adsorbents for the removal of metallic elements. As shown in the table, these two adsorbents were evaluated against various other adsorbents. The Fe3O4@HNT-SA exhibited a higher adsorption capacity compared to polyethyleneimine (PEI) cryogels, which had capacities ranging from 19.88 to 24.39 mg/g [53], and nano-sized SiO2, with capacities of 42.2 mg/g and 34.2 mg/g for the adsorptive removal of Cd and Pb [54]. Additionally, modified activated carbon for V achieved a capacity of 19.45 mg/g [55], which is smaller than the adsorption capacity of 24.7 mg/g achieved with Fe3O4@HNT-CTS. However, adsorbents such as a manganese-residues/serpentine-based composite reached 98.05 mg/g for Cd and 565.81 mg/g for Pb were significantly higher [56]. Metal–organic frameworks (MOFs) also attained a very high adsorption capacity of 188.12 mg/g for Cr and 349.09 mg/g for Pb [57]. Although the synthesised adsorbents have lower adsorption capacities compared to most of those listed in the table, their eco-friendliness and cost-effectiveness remain strong selling points, especially when considering that they are derived from natural sources.

3. Experimental

3.1. Materials and Chemicals

Halloysite nano clay (H2O9Si2·2H2O) (1.26–1.34 mL/g), iron(II) sulphate (ACS reagent ≥ 99%, FeSO4·7H2O) and iron(III) chloride hexahydrate (FeCl3·6H2O), ammonium hydroxide (~25%, NH4OH), and ultrapure nitric acid (65–70% w/w, HNO3, ultra-trace) were purchased from Merck (Johannesburg, South Africa). Sodium hydroxide (NaOH), ultrapure hydrochloric acid (37% w/w, HCl. 1000 mg L−1 standards of Cd, Cr, Pb, and Sb were bought from Sigma (South Africa, Gauteng). Chitosan (C12H24N2O9) and alginate were purchased from Sigma. The Mo, V, As and Sb standards were bought from Spectroscan Teknolab, Norway.

3.2. Instrumentation

Functional groups of Fe3O4@HNT-SA and Fe3O4@HNT-CTS were characterised using a PerkinElmer Fourier transform infrared spectroscopy (FTIR) system (KBr) with scans ranging from 400 to 4000 cm−1, located in Waltham, MA, USA. The scanning electron microscopy (SEM) coupled with energy dispersive spectrometry (EDS) SEM, TESCAN VEGA 3 XMU, LMH instrument from Bruno, Czech Republic, determined the morphology and elemental composition of the adsorbents. Transmission electron microscopy (TEM) was employed to delve deeper into the morphology of Fe3O4@HNT-SA and Fe3O4@HNT-CTS, using a TEM JEM-2100 obtained from JOEL Ltd, Tokyo, Japan. A SciTech ultrasonic bath system, which operates at 50 Hz and 150 W, was provided by Labotec in Midrand, South Africa, for the adsorption processes for As, Cd, Cr, Mo, Pb, Sb and V. The analyte residuals were analysed by inductively coupled plasma-optical emission spectrometer (ICP-OES) (model iCAP 6500 Duo from Thermo Scientific, Birmingham, UK). The crystal structures of Fe3O4@HNT-SA and Fe3O4@HNT-CTS were explored through X-ray powder diffraction (XRD) techniques using instruments from PANalytical (Almelo, The Netherlands). Surface area and pore size measurements were conducted employing the Brunauer–Emmett–Teller (BET) method with an ASAP2020 V3.00H unit from Micromeritics Instrument Corporation in Norcross, GA, USA. The zetasizer Instrument from Malvern (UK) was used to measure the zeta potential of Fe3O4@HNT-SA and Fe3O4@HNT-CTS. The experimental data were analysed with STATISTICA software version 14.

3.3. Synthesis of Magnetic Halloysite Nanotube (HNT@Fe3O4)

The magnetic halloysite nanotube (HNT@Fe3O4) was synthesised by modification of the previously reported method [61]. Approximately 4.8 g of FeCl3·6H2O and 2.4 g of FeSO4·7H2O were dissolved in 50 mL of water. Then, on the side, HNT powder (1 g) was dispersed in 100 mL of water by sonication for 30 min. The two mixtures were combined and stirred, then purged with nitrogen gas for 10 min at 60 °C. Then (~25%) of the ammonium hydroxide was added dropwise into the mixture, and the black precipitate (HNT@Fe3O4) formed. The mixture was left to age for 4 h at 70 °C. An external magnet was used to separate the black residue of HNT@Fe3O4, which was then washed several times with water. The prepared HNT@Fe3O4 was subsequently dried overnight in an oven at 60 °C.

3.4. Synthesis of HNT@Fe3O4-CTS

The HNT@Fe3O4-CTS was synthesised as follows: 2 g of HNT@Fe3O4 was added to 100 mL of ultrapure water and stirred at 500 rpm. In a separate beaker, 2.0 g of CTS was dissolved in 100 mL of 1% acetic acid solution. Once the CTS was fully dissolved, it was added drop by drop to the HNT@Fe3O4 solution and stirred for 24 h to obtain the HNT@Fe3O4-CTS composite. Then, acetic acid was removed by washing HNT@Fe3O4-CTS with water several times [62].

3.5. Preparation of Fe3O4@HNT-SA

The Fe3O4@HNT-SA nanocomposite was prepared using the co-precipitation method. First, 1.0 g of SA polymer was dissolved in 150 mL of ultrapure water. In a separate beaker, 1.0 g of HNT was dispersed in 50 mL of water and mixed with the SA solution. Next, the solution was mixed with 50 mL of FeCl3·6H2O (4.8 g) and FeSO4·7H2O (2.4 g) (forming a brown gel). The mixture was then heated at 70 °C and protected with nitrogen gas for 30 min. Then, 50 mL (~28%) of ammonium hydroxide was added, forming a black precipitate (Fe3O4@HNT-SA). The mixture was left to age for 2 h at 70 °C. The particles were then collected with an external magnet, washed multiple times with water, and dried overnight at 60 °C.

3.6. Batch Adsorption Experiments

The adsorption of Cd and Pb using Fe3O4@HNT-SA and As, Cr, Mo, Sb, and V by Fe3O4@HNT-CTS was conducted using batch experiments. The method was carried out by adding 30 mL of a solution (pH of 1.26–9.74, adjusted with 0.5 mol/L NaOH/HCl) containing mixtures of As, Cr, Mo, Sb, and V or Cd and Pb at 1.0 mg/L of Cd and Pb into 50 mL centrifuge tubes containing 5.78–54.2 mg of either Fe3O4@HNT-CTS or Fe3O4@HNT-SA hydrogel. The mixture was sonicated for 30 min, and the adsorbent was separated by magnetic decantation. The aqueous solution was filtered using 0.22 µm PVDF syringe filter membranes, and the residual elemental concentrations were analysed with ICP-OES. Each experiment was performed three times.
The suitability of Fe3O4@HNT-CTS and Fe3O4@HNT-SA hydrogels for the simultaneous removal of the analytes of interest was evaluated using the method mentioned above. The experimental conditions were fixed as follows: 30 mL, 1.0 mg/L, 5.5, 50 mg, and 30 min for sample volume, initial concentration, sample pH, mass of adsorbent, and contact time, respectively.
Equation (1) was used to calculate the percentage removal efficiency, where Ci and Ce were elemental concentrations before and after adsorption.
% R = C 0 C e C e × 100
The central composite design (CCD) was used to investigate the most significant factors for the removal of As, Cd, Cr, Mo, Pb, Sb and V using respective hydrogel material. The independent factors (mass of adsorbent (MA) and sample pH) were examined at five levels denoted as −α, −1, 0, +1, and +α, representing the chosen values from the lowest to the highest. The factor levels and their experimental domains are presented in Table 6.

3.7. Adsorption Kinetics and Isotherm Studies

The kinetics studies were investigated using two commonly used linear expressions, such as pseudo-first-order and pseudo-second-order, to explain the adsorption of As, Cd, Cr, Mo, Pb, Sb and V onto Fe3O4@HNT-CTS and Fe3O4@HNT-SA hydrogels. The batch adsorption experiments were conducted at optimum conditions to investigate the effect of contact time on the adsorption capacities of Fe3O4@HNT-CTS and Fe3O4@HNT-SA hydrogels for the target analytes. The adsorption process was performed for 5–60 min using the optimised method. The adsorption of As, Cr, Mo, Sb and V onto Fe3O4@HNT-CTS hydrogel was carried out using sample pH, mass of adsorbent, and sample volume of 5.5, 40 mg and 300 mL, respectively. For the adsorption of Cd and Pb onto Fe3O4@HNT-SA hydrogel, a sample pH ranging from 6.5 to 7.0 was used.
Linear equations of Langmuir and the Freundlich isotherms were used to describe the equilibrium data. Therefore, the effect of elemental concentrations on the adsorption capacities of Fe3O4@HNT-CTS and Fe3O4@HNT-SA hydrogels was investigated in the range of 1–9 mg/L (As, Cr, Mo, Sb and V) and 1–25 mg/L (Cd and Pb). It is important to note that the elemental concentrations in the aqueous solutions were determined by ICP-OES. The adsorption data for isotherms and kinetics were calculated using Equations (10) and (11).
q e = C o C e m × V
q t = C o C t m × V
where qe, qt, V, m, Ce and Ct were the equilibrium adsorption capacity (mg/g), equilibrium capacity (mg/g) at time t, sample volume (L), mass of adsorbent (g), elemental concentrations (mg/L) at equilibrium and time t, respectively.

3.8. Reusability

A total of 42 mg of Fe3O4@HNT-SA was added to the centrifuge. Following that, 30 mL of a sample solution containing Cd and Pb was also added to the centrifuge. The mixture was sonicated for 30 min. After sonication, the supernatant was separated, filtered, and analysed using ICP-OES. The recovered adsorbent was then washed with 2M HNO3 to eliminate any adsorbed analytes, followed by washing with NaOH and then with water. This process allowed the adsorbent to be reused up to five times. In a separate centrifuge tube, 40 mg of Fe3O4@HNT-CTS was added, followed by the addition of 30 mL of a sample solution containing analytes (As, Cr, Mo, Sb, and V). The mixture was sonicated for 30 min, after which the supernatant was filtered and analysed using ICP-OES. The adsorbed analytes were then removed with 10 mL of a 1:1 HCl/HNO3 solution and subsequently neutralised with 1 mL of 1 M NaOH. Finally, the mixture was washed several times with water.

3.9. Application to Real Water Samples

The applicability of the hydrogels was evaluated as an adsorbent for the extraction of specific analytes from river water samples taken from various villages in the Eastern Cape Province of South Africa. Optimised adsorption processes were employed for this purpose. To remove As, Cr, Mo, Sb, and V using Fe3O4@HNT-CTS adsorbent, the pH of the river samples was adjusted to 5.5. In contrast, the water samples were not adjusted because their pH was between 6.5 and 7.5 during the removal of Cd and Pb. The samples were then separated using an external magnet, filtered, and analysed using ICP-OES.

4. Conclusions

This study successfully utilised naturally occurring materials to create low-cost and environmentally friendly adsorbents with different functionalities to remove two sets of metallic pollutants: oxyanions (As, Cr, Mo, Sb and V) and metals (Cd and Pb). The Fe3O4@HNT-biopolymer demonstrated versatility in removing toxic ions. The Fe3O4@HNT-SA demonstrated its maximum removal efficiency at a pH of 7.5, while Fe3O4@HNT-CTS showed optimal performance at pH 5.5. The adsorption of this study was strongly dependent on pH, showing the interaction between metal ion speciation and protonation/deprotonation of the surface groups of the adsorbents. The adsorption of most analytes was governed by Langmuir models, suggesting that the reaction happens on a homogeneous surface of the adsorbent.
Additionally, Cr and Pb fit well within the Freundlich model. The two adsorbents also demonstrated in the interference studies that they can remove the target analyte in the presence of other or competing ions, and their application in the river water highlighted the applicability of this method. Integrating Halloysite nanoclay with biopolymers offers a sustainable and environmentally friendly approach to remediating toxic metals from the environment.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules30183777/s1, Figure S1: TEM images of (A) HNT and (B) Fe3O4. Figure S2: SEM images of (A) HNT (B) Fe3O4. Figure S3: (A) Langmuir isotherm plot for the adsorption of Cd and Pb using Fe3O4@HNT-SA, (B) Freundlich isotherm plot for Cd and Pb using Fe3O4@HNT-SA. Figure S4: (A) Langmuir isotherm plot for the adsorption of As, Cr, Mo, Sb and V using Fe3O4@HNT-CTS (B) Freundlich isotherm plot for As, Cr, Mo, Sb and V using Fe3O4@HNT-CTS. Figure S5: (A) Pseudo first order kinetics and (B) pseudo second order kinetics of Cd and Pb using Fe3O4@HNTSA adsorbent. Figure S6: (A) Pseudo first order kinetics and (B) pseudo second order kinetics of As, Cr, Mo, Sb and V using Fe3O4@HNT-CTS. Table S1. The effect of interfering studies on the adsorption of Cd and Pb using Fe3O4@HNT-SA. Table S2. The effect of interfering studies on the adsorption of As, Cr, Mo, Sb and V onto Fe3O4@HNT-CTS.

Author Contributions

N.B.M.: Conceptualisation, visualisation, methodology, investigation, validation, formal analysis, data curation, writing—original draft. P.N.N.: Conceptualisation, supervision, funding acquisition, validation, visualisation, resources, software, formal analysis, writing, review and editing. L.N.: Methodology, writing—review and editing, funding acquisition, supervision, formal analysis, conceptualisation. All authors have read and agreed to the published version of the manuscript.

Funding

The authors express their gratitude to the Department of Chemical Sciences at the University of Johannesburg for granting access to the laboratory facilities and instruments essential for this project. They also acknowledge the financial support received from the Department of Science and Innovation–National Research Foundation South African Research Chair Initiative (DSI-NRF SARChI, Grant No. 91230), the NRF Thuthuka Programme (Grant No. 138375), and the NRF Freestanding, Innovation, and Scarce Skills Scholarship for Doctoral Students (Grant No. 130347).

Acknowledgments

The authors wish to express their gratitude to the Department of Chemical Sciences at the University of Johannesburg (DFC Campus) for providing a supportive environment, as well as the essential instruments and laboratory facilities required for this project. We also extend our thanks to the Department of Science and Innovation, the National Research Foundation, and the South African Research Chair Initiative (DSI-NRF SARChI) for their support.

Conflicts of Interest

The authors confirm that they do not have any known conflicting financial interests or personal relationships that could have impacted the outcomes presented in this study.

Abbreviations

CCDCentral composite design
CTSChitosan
Fe3O4@HNT-CTSMagnetic halloysite nanoclay chitosan
HNTHalloysite nanotubes
ICP-OESInductively coupled plasma optical emission spectroscopy
Fe3O4@HNT-SAMagnetic halloysite nanoclay alginate
NIPNon-imprinted polymer
SAAlginate

References

  1. Padhiary, M.; Kumar, R. Assessing the Environmental Impacts of Agriculture, Industrial Operations, and Mining on Agro-Ecosystems. In Smart Internet of Things for Environment and Healthcare; Springer Nature: Cham, Switzerland, 2024; pp. 107–126. [Google Scholar]
  2. Wang, C.; Jiang, A.; Liu, X.; Yuen Koh, K.; Yang, Y.; Chen, J.P.; Li, K. Amorphous Metal-Organic Framework UiO-66-NO2 for Removal of Oxyanion Pollutants: Towards Improved Performance and Effective Reusability. Sep. Purif. Technol. 2022, 295, 121014. [Google Scholar] [CrossRef]
  3. Çelebi, H.; Gök, G.; Gök, O. Adsorption Capability of Brewed Tea Waste in Waters Containing Toxic Lead(II), Cadmium (II), Nickel (II), and Zinc(II) Heavy Metal Ions. Sci. Rep. 2020, 10, 17570. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, Y.; Duan, X. Chemical Precipitation of Heavy Metals from Wastewater by Using the Synthetical Magnesium Hydroxy Carbonate. Water Sci. Technol. 2020, 81, 1130–1136. [Google Scholar] [CrossRef] [PubMed]
  5. Li, Y.; Liu, Y.; Liu, C.; Feng, L.; Yang, S.; Shan, Y.; Xiao, F. Quantitatively Ion-Exchange between Mg(II) and Pb(II)/Cd(II) during the Highly Efficient Adsorption by MgO-Loaded Lotus Stem Biochar. J. Taiwan Inst. Chem. Eng. 2023, 144, 104736. [Google Scholar] [CrossRef]
  6. Harharah, R.H.; Abdalla, G.M.T.; Elkhaleefa, A.; Shigidi, I.; Harharah, H.N. A Study of Copper (II) Ions Removal by Reverse Osmosis under Various Operating Conditions. Separations 2022, 9, 155. [Google Scholar] [CrossRef]
  7. Elboughdiri, N. The Use of Natural Zeolite to Remove Heavy Metals Cu (II), Pb (II) and Cd (II), from Industrial Wastewater. Cogent Eng. 2020, 7, 1782623. [Google Scholar] [CrossRef]
  8. Ayouch, I.; Kassem, I.; Kassab, Z.; Barrak, I.; Barhoun, A.; Jacquemin, J.; Draoui, K.; Achaby, M. El Crosslinked Carboxymethyl Cellulose-Hydroxyethyl Cellulose Hydrogel Films for Adsorption of Cadmium and Methylene Blue from Aqueous Solutions. Surf. Interfaces 2021, 24, 101124. [Google Scholar] [CrossRef]
  9. Hnamte, M.; Pulikkal, A.K. Clay-Polymer Nanocomposites for Water and Wastewater Treatment: A Comprehensive Review. Chemosphere 2022, 307, 135869. [Google Scholar] [CrossRef]
  10. Gul, S.; Ahmad, Z.; Asma, M.; Ahmad, M.; Rehan, K.; Munir, M.; Bazmi, A.A.; Ali, H.M.; Mazroua, Y.; Salem, M.A.; et al. Effective Adsorption of Cadmium and Lead Using SO3H-Functionalized Zr-MOFs in Aqueous Medium. Chemosphere 2022, 307, 135633. [Google Scholar] [CrossRef]
  11. Maqbool, A.; Shahid, A.; Jahan, Z.; Bilal Khan Niazi, M.; Ali Inam, M.; Tawfeek, A.M.; M Kamel, E.; Saeed Akhtar, M. Development of ZnO-GO-NiO Membrane for Removal of Lead and Cadmium Heavy Metal Ions from Wastewater. Chemosphere 2023, 338, 139622. [Google Scholar] [CrossRef]
  12. Kończyk, J.; Kluziak, K.; Kołodyńska, D. Adsorption of Vanadium (V) Ions from the Aqueous Solutions on Different Biomass-Derived Biochars. J. Environ. Manag. 2022, 313, 114958. [Google Scholar] [CrossRef]
  13. Moges, A.; Nkambule, T.T.I.; Fito, J. The Application of GO-Fe3O4 Nanocomposite for Chromium Adsorption from Tannery Industry Wastewater. J. Environ. Manag. 2022, 305, 114369. [Google Scholar] [CrossRef]
  14. Mutar, R.F.; Saleh, M.A. Optimization of Arsenic Ions Adsorption and Removal from Hospitals Wastewater by Nano-Bentonite Using Central Composite Design. Mater. Today Proc. 2022, 60, 1248–1256. [Google Scholar] [CrossRef]
  15. Nie, J.; Yao, Z.; Shao, P.; Jing, Y.; Bai, L.; Xing, D.; Yi, G.; Li, D.; Liu, Y.; Yang, L.; et al. Revisiting the Adsorption of Antimony on Manganese Dioxide: The Overlooked Dissolution of Manganese. Chem. Eng. J. 2022, 429, 132468. [Google Scholar] [CrossRef]
  16. del Mar Orta, M.; Martín, J.; Santos, J.L.; Aparicio, I.; Medina-Carrasco, S.; Alonso, E. Biopolymer-Clay Nanocomposites as Novel and Ecofriendly Adsorbents for Environmental Remediation. Appl. Clay Sci. 2020, 198, 105838. [Google Scholar] [CrossRef]
  17. Alver, E.; Metin, A.Ü.; Brouers, F. Methylene Blue Adsorption on Magnetic Alginate/Rice Husk Bio-Composite. Int. J. Biol. Macromol. 2020, 154, 104–113. [Google Scholar] [CrossRef] [PubMed]
  18. Zeng, H.; Sun, S.; Xu, K.; Zhao, W.; Hao, R.; Zhang, J.; Li, D. Iron-Loaded Magnetic Alginate-Chitosan Double-Gel Interpenetrated Porous Beads for Phosphate Removal from Water: Preparation, Adsorption Behavior and PH Stability. React. Funct. Polym. 2022, 177, 105328. [Google Scholar] [CrossRef]
  19. ALSamman, M.T.; Sánchez, J. Recent Advances on Hydrogels Based on Chitosan and Alginate for the Adsorption of Dyes and Metal Ions from Water. Arab. J. Chem. 2021, 14, 103455. [Google Scholar] [CrossRef]
  20. Majiya, H.; Clegg, F.; Sammon, C. Bentonite-Chitosan Composites or Beads for Lead (Pb) Adsorption: Design, Preparation, and Characterisation. Appl. Clay Sci. 2023, 246, 107180. [Google Scholar] [CrossRef]
  21. Zhang, F.; Wang, B.; Jie, P.; Zhu, J.; Cheng, F. Preparation of Chitosan/Lignosulfonate for Effectively Removing Pb(II) in Water. Polymer 2021, 228, 123878. [Google Scholar] [CrossRef]
  22. Rahaman, M.H.; Islam, M.A.; Islam, M.M.; Rahman, M.A.; Alam, S.M.N. Biodegradable Composite Adsorbent of Modified Cellulose and Chitosan to Remove Heavy Metal Ions from Aqueous Solution. Curr. Res. Green Sustain. Chem. 2021, 4, 100119. [Google Scholar] [CrossRef]
  23. Zia, Q.; Tabassum, M.; Meng, J.; Xin, Z.; Gong, H.; Li, J. Polydopamine-Assisted Grafting of Chitosan on Porous Poly (L-Lactic Acid) Electrospun Membranes for Adsorption of Heavy Metal Ions. Int. J. Biol. Macromol. 2021, 167, 1479–1490. [Google Scholar] [CrossRef]
  24. Alsuhaibani, A.M.; Alayyafi, A.A.A.; Albedair, L.A.; El-Desouky, M.G.; El-Bindary, A.A. Efficient Fabrication of a Composite Sponge for Cr(VI) Removal via Citric Acid Cross-Linking of Metal-Organic Framework and Chitosan: Adsorption Isotherm, Kinetic Studies, and Optimization Using Box-Behnken Design. Mater. Today Sustain. 2024, 26, 100732. [Google Scholar] [CrossRef]
  25. Ferreira, D.C.M.; dos Santos, T.C.; dos Reis Coimbra, J.S.; de Oliveira, E.B. Chitosan/Carboxymethylcellulose Polyelectrolyte Complexes (PECs) Are an Effective Material for Dye and Heavy Metal Adsorption from Water. Carbohydr. Polym. 2023, 315, 120977. [Google Scholar] [CrossRef] [PubMed]
  26. Prabhu, R.; Shetty, K.; Jeevananda, T.; Ananda Murthy, H.C.; Sillanpaa, M.; Nhat, T. Novel Polyaniline–Halloysite Nanoclay Hybrid Composites: Synthesis, Physico-Chemical, Thermal and Electrical Properties. Inorg. Chem. Commun. 2023, 148, 110328. [Google Scholar] [CrossRef]
  27. Queiroz, M.F.; Melo, K.R.T.; Sabry, D.A.; Sassaki, G.L.; Rocha, H.A.O. Does the Use of Chitosan Contribute to Oxalate Kidney Stone Formation? Mar. Drugs 2015, 13, 141. [Google Scholar] [CrossRef]
  28. Drabczyk, A.; Kudłacik-Kramarczyk, S.; Głab, M.; Kedzierska, M.; Jaromin, A.; Mierzwiński, D.; Tyliszczak, B. Physicochemical Investigations of Chitosan-Based Hydrogels Containing Aloe Vera Designed for Biomedical Use. Materials 2020, 13, 3073. [Google Scholar] [CrossRef]
  29. Liu, Q.; Li, Q.; Xu, S.; Zheng, Q.; Cao, X. Preparation and Properties of 3D Printed Alginate-Chitosan Polyion Complex Hydrogels for Tissue Engineering. Polymers 2018, 10, 664. [Google Scholar] [CrossRef] [PubMed]
  30. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S.W. Physisorption of Gases, with Special Reference to the Evaluation of Surface Area and Pore Size Distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef]
  31. Abdel-Fadeel, M.A.; Aljohani, N.S.; Al-Mhyawi, S.R.; Halawani, R.F.; Aljuhani, E.H.; Salam, M.A. A Simple Method for Removal of Toxic Dyes Such as Brilliant Green and Acid Red from the Aquatic Environment Using Halloysite Nanoclay. J. Saudi Chem. Soc. 2022, 26, 101475. [Google Scholar] [CrossRef]
  32. Faaliyan, K.; Abdoos, H.; Borhani, E.; Afghahi, S.S.S. Magnetite-Silica Nanoparticles with Core-Shell Structure: Single-Step Synthesis, Characterization and Magnetic Behavior. J. Sol-Gel Sci. Technol. 2018, 88, 609–617. [Google Scholar] [CrossRef]
  33. Podgorbunskikh, E.; Kuskov, T.; Rychkov, D.; Lomovskii, O.; Bychkov, A. Mechanical Amorphization of Chitosan with Different Molecular Weights. Polymers 2022, 14, 4438. [Google Scholar] [CrossRef]
  34. Shameem, A.; Devendran, P.; Siva, V.; Venkatesh, K.S.; Manikandan, A.; Bahadur, S.A.; Nallamuthu, N. Dielectric Investigation of NaLiS Nanoparticles Loaded on Alginate Polymer Matrix Synthesized by Single Pot Microwave Irradiation. J. Inorg. Organomet. Polym. Mater. 2018, 28, 671–678. [Google Scholar] [CrossRef]
  35. Ramanery, F.P.; Mansur, A.A.P.; Mansur, H.S. One-Step Colloidal Synthesis of Biocompatible Water-Soluble ZnS Quantum Dot/Chitosan Nanoconjugates. Nanoscale Res. Lett. 2013, 8, 512. [Google Scholar] [CrossRef]
  36. Nargis, A.; Habib, A.; Rashid, H.O.; Harun, H.B.; Islam Sarker, M.S.; Jin, R.; Liu, G.; Liu, W.; Al-Razee, A.N.M.; Chen, K.; et al. Status of Multielement in Water of the River Buriganga, Bangladesh: Aquatic Chemistry of Metal Ions in Polluted River Water. Emerg. Contam. 2021, 7, 99–115. [Google Scholar] [CrossRef]
  37. Cabrerizo, A.; Bulteel, D.; Waligora, J.; Landrot, G.; Fonda, E.; Olard, F. Chemical, Mineralogical, and Environmental Characterization of Tunnel Boring Muds for Their Valorization in Road Construction: A Focus on Molybdenum Characterization. Environ. Sci. Pollut. Res. 2020, 27, 44314–44324. [Google Scholar] [CrossRef]
  38. Lim, M.; Han, G.C.; Ahn, J.W.; You, K.S.; Kim, H.S. Leachability of Arsenic and Heavy Metals from Mine Tailings of Abandoned Metal Mines. Int. J. Environ. Res. Public Health 2009, 6, 2865. [Google Scholar] [CrossRef] [PubMed]
  39. Yuan, X.; Li, J.; Luo, L.; Zhong, Z.; Xie, X. Advances in Sorptive Removal of Hexavalent Chromium (Cr(VI)) in Aqueous Solutions Using Polymeric Materials. Polymers 2023, 15, 388. [Google Scholar] [CrossRef] [PubMed]
  40. Baskaran, P.; Abraham, M. Adsorption of Cadmium (Cd) and Lead (Pb) Using Powdered Activated Carbon Derived from Cocos Nucifera Waste: A Kinetics and Equilibrium Study for Long-Term Sustainability. Sustain. Energy Technol. Assess. 2022, 53, 102709. [Google Scholar] [CrossRef]
  41. Ramutshatsha-Makhwedzha, D.; Mbaya, R.; Mavhungu, M.L. Application of Activated Carbon Banana Peel Coated with Al2O3-Chitosan for the Adsorptive Removal of Lead and Cadmium from Wastewater. Materials 2022, 15, 860. [Google Scholar] [CrossRef]
  42. Kochar, C.; Taneja, L.; Yadav, P.K.; Tripathy, S.S. RSM-CCD Approach for Optimization Study on Effective Remediation of Lead and Cadmium from Water Using Surface-Modified Water Caltrop Peel Biochar. Biomass Convers. Biorefinery 2024, 14, 17985–18003. [Google Scholar] [CrossRef]
  43. Karrat, A.; Palacios-Santander, J.M.; Amine, A.; Cubillana-Aguilera, L. A Novel Magnetic Molecularly Imprinted Polymer for Selective Extraction and Determination of Quercetin in Plant Samples. Anal. Chim. Acta 2022, 1203, 339709. [Google Scholar] [CrossRef]
  44. Fu, C.; Zhu, X.; Dong, X.; Zhao, P.; Wang, Z. Study of Adsorption Property and Mechanism of Lead(II) and Cadmium(II) onto Sulfhydryl Modified Attapulgite. Arab. J. Chem. 2021, 14, 102960. [Google Scholar] [CrossRef]
  45. Yasin, S.R.; Jasim, S.S.; Al-Baldawi, I.A. Removal of Lead, Cadmium, and Copper from Wastewater Using Cinnamon Bark Waste to Introduce It as a Value-Added Product: Removal, Kinetics and Thermodynamics Study. Desalination Water Treat. 2025, 323, 101345. [Google Scholar] [CrossRef]
  46. Saleem, M.; Rasheed, A.; Haider, F.; Raja, S.S. Evaluation of Cadmium Remediation Studies by Cobalt-Doped Magnetic Biochar Composite (CoFe2O4@MBC) Derived from Corn Stalk. Int. J. Environ. Sci. Technol. 2025, 3, 1–14. [Google Scholar] [CrossRef]
  47. Semysim, F.A.; Ridha, R.K.; Azooz, E.A.; Snigur, D. Switchable Hydrophilicity Solvent-Assisted Solidified Floating Organic Drop Microextraction for Separation and Determination of Arsenic in Water and Fish Samples. Talanta 2024, 272, 125782. [Google Scholar] [CrossRef]
  48. Ghorbani, Y.A.; Ghoreishi, S.M.; Ghani, M. Derived N-Doped Carbon through Core-Shell Structured Metal-Organic Frameworks as a Novel Sorbent for Dispersive Solid Phase Extraction of Cr(III) and Pb(II) from Water Samples Followed by Quantitation through Flame Atomic Absorption Spectrometry. Microchem. J. 2020, 155, 104786. [Google Scholar] [CrossRef]
  49. Zeng, W.; Hu, Z.; Luo, J.; Hou, X.; Jiang, X. Highly Sensitive Determination of Trace Antimony in Water Samples by Cobalt Ion Enhanced Photochemical Vapor Generation Coupled with Atomic Fluorescence Spectrometry or ICP-MS. Anal. Chim. Acta 2022, 1191, 339361. [Google Scholar] [CrossRef] [PubMed]
  50. Tuzen, M.; Altunay, N.; Hazer, B.; Mogaddam, M.R.A. Synthesis of Polystyrene-Polyricinoleic Acid Copolymer Containing Silver Nano Particles for Dispersive Solid Phase Microextraction of Molybdenum in Water and Food Samples. Food Chem. 2022, 369, 130973. [Google Scholar] [CrossRef] [PubMed]
  51. Vasseghian, Y.; Sadeghi Rad, S.; Vilas-Boas, J.A.; Khataee, A. A Global Systematic Review, Meta-Analysis, and Risk Assessment of the Concentration of Vanadium in Drinking Water Resources. Chemosphere 2021, 267, 128904. [Google Scholar] [CrossRef]
  52. Wang, Y.; Peng, C.; Padilla-Ortega, E.; Robledo-Cabrera, A.; López-Valdivieso, A. Cr(VI) Adsorption on Activated Carbon: Mechanisms, Modeling and Limitations in Water Treatment. J. Environ. Chem. Eng. 2020, 8, 104031. [Google Scholar] [CrossRef]
  53. Bagdat, S.; Tokay, F.; Demirci, S.; Yilmaz, S.; Sahiner, N. Removal of Cd(II), Co(II), Cr(III), Ni(II), Pb(II) and Zn(II) Ions from Wastewater Using Polyethyleneimine (PEI) Cryogels. J. Environ. Manag. 2023, 329, 117002. [Google Scholar] [CrossRef]
  54. El-Feky, H.H.; Behiry, M.S.; Amin, A.S.; Nassar, M.Y. Facile Fabrication of Nano-Sized SiO2 by an Improved Sol–Gel Route: As an Adsorbent for Enhanced Removal of Cd(II) and Pb(II) Ions. J. Inorg. Organomet. Polym. Mater. 2022, 32, 1129–1141. [Google Scholar] [CrossRef]
  55. Ali, M.M.S.; Imam, D.M.; El-Nadi, Y.A. Vanadium(V) Removal and Recovery by Adsorption onto Modified Activated Carbon Derived from Natural Hydroxyapatite. J. Iran. Chem. Soc. 2021, 18, 2771–2784. [Google Scholar] [CrossRef]
  56. Ma, M.Y.; Ke, X.; Liu, Y.C.; Ha, Z.H.; Wang, T.; Li, J.; Zhang, F.; Zhang, T.C. A Novel Electrolytic-Manganese-Residues-and-Serpentine-Based Composite (S-EMR) for Enhanced Cd(II) and Pb(II) Adsorption in Aquatic Environment. Rare Metals 2023, 42, 346–358. [Google Scholar] [CrossRef]
  57. Wang, H.; Wang, S.; Wang, S.; Fu, L.; Zhang, L. The One-Step Synthesis of a Novel Metal–Organic Frameworks for Efficient and Selective Removal of Cr(VI) and Pb(II) from Wastewater: Kinetics, Thermodynamics and Adsorption Mechanisms. J. Colloid. Interface Sci. 2023, 640, 230–245. [Google Scholar] [CrossRef] [PubMed]
  58. Wang, N.; Feng, J.; Yan, W.; Zhang, L.; Liu, Y.; Mu, R. Dual-Functional Sites for Synergistic Adsorption of Cr(VI) and Sb(V) by Polyaniline-TiO2 Hydrate: Adsorption Behaviors, Sites and Mechanisms. Front. Environ. Sci. Eng. 2022, 16, 105. [Google Scholar] [CrossRef]
  59. Xu, R.; Li, Q.; Nan, X.; Jiang, G.; Wang, L.; Xiong, J.; Yang, Y.; Xu, B.; Jiang, T. Simultaneous Removal of Antimony(III/V) and Arsenic(III/V) from Aqueous Solution by Bacteria–Mediated Kaolin@Fe–Mn Binary (Hydr)Oxides Composites. Appl. Clay Sci. 2022, 217, 106392. [Google Scholar] [CrossRef]
  60. Wang, Z.; Tian, T.; Xu, K.; Jia, Y.; Zhang, C.; Li, J.; Wang, Z. Removal of Antimony(III) by Magnetic MIL-101(Cr)-NH2 Loaded with SiO2: Optimization Based on Response Surface Methodology and Adsorption Properties. Chem. Pap. 2022, 76, 2733–2745. [Google Scholar] [CrossRef]
  61. Li, L.; Wang, F.; Lv, Y.; Liu, J.; Zhang, D.; Shao, Z. Halloysite Nanotubes and Fe3O4 Nanoparticles Enhanced Adsorption Removal of Heavy Metal Using Electrospun Membranes. Appl. Clay Sci. 2018, 161, 225–234. [Google Scholar] [CrossRef]
  62. Türkeş, E.; Sağ Açıkel, Y. Synthesis and Characterization of Magnetic Halloysite–Chitosan Nanocomposites: Use in the Removal of Methylene Blue in Wastewaters. Int. J. Environ. Sci. Technol. 2020, 17, 1281–1294. [Google Scholar] [CrossRef]
Figure 1. FTIR spectra of (A) Fe3O4@HNT-CTS (B) Fe3O4@HNT-SA.
Figure 1. FTIR spectra of (A) Fe3O4@HNT-CTS (B) Fe3O4@HNT-SA.
Molecules 30 03777 g001
Figure 2. TEM images of (A) CTS (B) Fe3O4@HNT-CTS (C) Fe3O4@HNT-CTS (D) SA (E) SA-HNT@Fe3O4 (F) SA-HNT@Fe3O4.
Figure 2. TEM images of (A) CTS (B) Fe3O4@HNT-CTS (C) Fe3O4@HNT-CTS (D) SA (E) SA-HNT@Fe3O4 (F) SA-HNT@Fe3O4.
Molecules 30 03777 g002
Figure 3. SEM (A) SA (B) Fe3O4@HNT-SA (C) Fe3O4@HNT-SA (EDS) (D) CTS (E) Fe3O4@HNT-CTS (F) Fe3O4@HNT-CTS (EDS).
Figure 3. SEM (A) SA (B) Fe3O4@HNT-SA (C) Fe3O4@HNT-SA (EDS) (D) CTS (E) Fe3O4@HNT-CTS (F) Fe3O4@HNT-CTS (EDS).
Molecules 30 03777 g003
Figure 4. BET adsorption–desorption of isotherms of CTS, SA, Fe3O4, HNT, Fe3O4@HNT-CTS and Fe3O4@HNT-SA.
Figure 4. BET adsorption–desorption of isotherms of CTS, SA, Fe3O4, HNT, Fe3O4@HNT-CTS and Fe3O4@HNT-SA.
Molecules 30 03777 g004
Figure 5. The X-Ray diffractograms (A) Fe3O4@HNT-CTS (B) Fe3O4@HNT-SA.
Figure 5. The X-Ray diffractograms (A) Fe3O4@HNT-CTS (B) Fe3O4@HNT-SA.
Molecules 30 03777 g005
Figure 6. (A) Surface charge of Fe3O4@HNT-SA (B) Surface charge of Fe3O4@HNT-CTS.
Figure 6. (A) Surface charge of Fe3O4@HNT-SA (B) Surface charge of Fe3O4@HNT-CTS.
Molecules 30 03777 g006
Figure 7. Selection of adsorbents for the adsorption of As, Cd, Cr, Mo, Pb, Sb, and V onto Fe3O4@HNT-SA and Fe3O4@HNT-CTS hydrogels. Experimental conditions: Sample volume: 30 mL; initial concentration: 1.0 mg/L; sample pH: 5.5; mass of adsorbent: 50 mg; contact time: 30 min.
Figure 7. Selection of adsorbents for the adsorption of As, Cd, Cr, Mo, Pb, Sb, and V onto Fe3O4@HNT-SA and Fe3O4@HNT-CTS hydrogels. Experimental conditions: Sample volume: 30 mL; initial concentration: 1.0 mg/L; sample pH: 5.5; mass of adsorbent: 50 mg; contact time: 30 min.
Molecules 30 03777 g007
Figure 8. Pareto chart of (A) Fe3O4@HNT-SA (B) Fe3O4@HNT-CTS.
Figure 8. Pareto chart of (A) Fe3O4@HNT-SA (B) Fe3O4@HNT-CTS.
Molecules 30 03777 g008
Figure 9. Response surfaces for combined effects of pH and MA for the removal of (A) Cd and Pb using Fe3O4@HNT-SA (B) As, Cr, Mo, Sb and V using Fe3O4@HNT-CTS.
Figure 9. Response surfaces for combined effects of pH and MA for the removal of (A) Cd and Pb using Fe3O4@HNT-SA (B) As, Cr, Mo, Sb and V using Fe3O4@HNT-CTS.
Molecules 30 03777 g009
Figure 10. Predicted values and desirability function for removal of (A) As, Cr, Mo, Sb and V using Fe3O4@HNT-CTS (B) Cd and Pb using Fe3O4@HNT-SA.
Figure 10. Predicted values and desirability function for removal of (A) As, Cr, Mo, Sb and V using Fe3O4@HNT-CTS (B) Cd and Pb using Fe3O4@HNT-SA.
Molecules 30 03777 g010
Figure 11. Equilibrium data for adsorption of (A) Cd, Pb, experimental conditions pH = 7.5, MA = 42mg and ET = 30 min, (B) As, Cr, Mo, Sb and V with experimental conditions: pH = 5.5, MA = 40mg and ET = 30 min.
Figure 11. Equilibrium data for adsorption of (A) Cd, Pb, experimental conditions pH = 7.5, MA = 42mg and ET = 30 min, (B) As, Cr, Mo, Sb and V with experimental conditions: pH = 5.5, MA = 40mg and ET = 30 min.
Molecules 30 03777 g011
Figure 12. Equilibrium data for adsorption of (A) As, Cr, Mo, Sb and V with experimental conditions: pH = 5.5, MA = 40 mg and initial concentration = 9 mg/L and (B) Cd, Pb, experimental conditions pH = 7.5, MA = 42 mg and initial concentration = 17 mg/L.
Figure 12. Equilibrium data for adsorption of (A) As, Cr, Mo, Sb and V with experimental conditions: pH = 5.5, MA = 40 mg and initial concentration = 9 mg/L and (B) Cd, Pb, experimental conditions pH = 7.5, MA = 42 mg and initial concentration = 17 mg/L.
Molecules 30 03777 g012
Figure 13. Recyclability study of Fe3O4@HNT-SA for (A) Cd and Pb removal and (B) Fe3O4@HNT-CTS for As, Cr, Mo, Sb, and V removal.
Figure 13. Recyclability study of Fe3O4@HNT-SA for (A) Cd and Pb removal and (B) Fe3O4@HNT-CTS for As, Cr, Mo, Sb, and V removal.
Molecules 30 03777 g013
Table 1. Textual properties of SA, HNT, CTS, Fe3O4, Fe3O4@HNT-SA and Fe3O4@HNT-CTS.
Table 1. Textual properties of SA, HNT, CTS, Fe3O4, Fe3O4@HNT-SA and Fe3O4@HNT-CTS.
AdsorbentSurface Area (m2/g)Pore Volume (cm3/g)Average Pore Size (nm)
CTS1.600.004998.57
SA3.820.002317.46
HNT60.40.34513.1
Fe3O486.00.30212.5
Fe3O4@HNT-CTS58.40.1799.85
Fe3O4@HNT-SA40.70.14411.2
Table 2. Linearised isotherm model parameters for the adsorption of As, Cr, Mo, Sb and V onto Fe3O4@HNT-CTS and the adsorption of Cd and Pb onto Fe3O4@HNT-SA.
Table 2. Linearised isotherm model parameters for the adsorption of As, Cr, Mo, Sb and V onto Fe3O4@HNT-CTS and the adsorption of Cd and Pb onto Fe3O4@HNT-SA.
ModelsParameters AsVCrMoSbCdPb
Fe3O4@HNT-CTS HydrogelFe3O4@HNT-SA Hydrogel
qe (expt) (mg/g)30.319.928.422.224.752.257.7
Langmuirqmax (mg/g)33.6027.269.927.031.756.269.4
KL(L/mg)2.480.740.2461.461.191.2259.9
RL0.04640.1660.3750.09220.1090.04720.0735
R20.99510.96930.66670.99200.97200.98610.9589
FreundlichKF (mg/g)11.611.121.910.613.5016.426.0
N0.3310.4290.3850.3780.3670.3040.301
R20.94590.95740.94410.9420.92540.97740.9788
Table 3. Linearised kinetic model parameters for the adsorption of Cd and Pb onto the Fe3O4@HNT-SA adsorbent and the adsorption of As, Cr, Mo, Sb and V using Fe3O4@HNT-CTS.
Table 3. Linearised kinetic model parameters for the adsorption of Cd and Pb onto the Fe3O4@HNT-SA adsorbent and the adsorption of As, Cr, Mo, Sb and V using Fe3O4@HNT-CTS.
ModelsParametersCdPbAsVCrMoSb
qe, expt (mg/g)52.257.730.319.928.422.224.7
Pseudo-first-orderk1 (1/min)0.07060.05700.03020.08340.08430.01150.0429
qe,calc.34.1436.3420.221.933.314.020.3
R20.87990.99350.97090.93100.92540.88730.9259
Pseudo-second-orderk2 (g/mg∙min)0.002980.02200.00005750.004260.002190.008810.00243
qe (mg/g)57.462.532.6823.2634.8416.226.2
R20.99370.99530.99460.99080.98570.99590.9598
Table 4. Thermodynamic parameters of adsorption of Cd and Pb onto the Fe3O4@HNT-SA adsorbent and the adsorption of As, Cr, Mo, Sb and V using Fe3O4@HNT-CTS.
Table 4. Thermodynamic parameters of adsorption of Cd and Pb onto the Fe3O4@HNT-SA adsorbent and the adsorption of As, Cr, Mo, Sb and V using Fe3O4@HNT-CTS.
Analytes ΔG° (kj, mol−1) ΔH°
(kj, mol−1)
ΔS°
(j, mol−1k−1)
298 K308 K313 K
Fe3O4@HNT-Cd −4.30−1.95−0.768−74.6−236
Fe3O4@HNT-SA-Pb−5.32−6.36−6.8825.7104
Fe3O4@HNT-CTS-As−4.46−4.85−5.057.2939.4
Fe3O4@HNT-CTS-Cr−1.78−2.18−2.39133451
Fe3O4@HNT-CTS-Mo−2.85−4.51−5.4448.6173
Fe3O4@HNT-CTS-Sb−4.62−2.52−1.48−67.1−210
Fe3O4@HNT-CTS-V−3.43−5.04−5.5844.6161
Table 5. Comparison of Fe3O4@HNT-CTS and Fe3O4@HNT-SA with other adsorbents for the removal of metals/metalloids.
Table 5. Comparison of Fe3O4@HNT-CTS and Fe3O4@HNT-SA with other adsorbents for the removal of metals/metalloids.
AdsorbentAnalyteMaximum Uptake (mg/g)References
Nano-sized silicon dioxide Cd and Pb42.2 and 34.2[54]
Polyethyleneimine (PEI) cryogelsCd, Co, Cr, Ni, Pb, and Zn 19.88–24.39 [53]
manganese-residues-and-serpentine-based compositeCd and Pb98.05 and 565.81[56]
Polyaniline-TiO2 hydrate:Cr and Sb394 and 48.5[58]
Bacteria–mediated kaolin@Fe–Mn binary hydroxides Sb(III), Sb(V), As(III) and As(V) 177.19, 56.26, 62.92 and 42.18 [59]
metal–organic frameworksCr and Pb188.12 and 349.09 [57]
Sb30.26, 86.35[60]
modified activated carbonV19.45 [55,58]
Fe3O4@HNT-CTS As, Cr, Mo, Sb, and V 30.3, 19.9, 28.4, 22.2, 24.7This study
Fe3O4@HNT-SA Cd and Pb52.2 and 57.7This study
Table 6. Factors and levels employed in central composite design for the removal of As, Cd, Cr, Mo, Pb, Sb and V.
Table 6. Factors and levels employed in central composite design for the removal of As, Cd, Cr, Mo, Pb, Sb and V.
Factors−α−10+
Mass adsorbent (MA, mg)5.7810305054.2
Sample pH1.262.05.59.09.74
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

Mabaso, N.B.; Nomngongo, P.N.; Nyaba, L. Removal of Oxyanions and Trace Metals from River Water Samples Using Magnetic Biopolymer/Halloysite Nanocomposites. Molecules 2025, 30, 3777. https://doi.org/10.3390/molecules30183777

AMA Style

Mabaso NB, Nomngongo PN, Nyaba L. Removal of Oxyanions and Trace Metals from River Water Samples Using Magnetic Biopolymer/Halloysite Nanocomposites. Molecules. 2025; 30(18):3777. https://doi.org/10.3390/molecules30183777

Chicago/Turabian Style

Mabaso, Nyeleti Bridget, Philiswa Nosizo Nomngongo, and Luthando Nyaba. 2025. "Removal of Oxyanions and Trace Metals from River Water Samples Using Magnetic Biopolymer/Halloysite Nanocomposites" Molecules 30, no. 18: 3777. https://doi.org/10.3390/molecules30183777

APA Style

Mabaso, N. B., Nomngongo, P. N., & Nyaba, L. (2025). Removal of Oxyanions and Trace Metals from River Water Samples Using Magnetic Biopolymer/Halloysite Nanocomposites. Molecules, 30(18), 3777. https://doi.org/10.3390/molecules30183777

Article Metrics

Back to TopTop