Entsar H. Taha*a,
Adel A. El-Zahhar
b,
Majed M. Alghamdib,
Ahmed M. Masoud
c,
Mohamed F. Kamel
c and
Mohamed H. Taha
c
aDepartment of Plant Protection, Faculty of Agriculture, Ain Shams University, Egypt, Nuclear Materials Authority, P. O. Box 530, El Maddi, Cairo, Egypt. E-mail: tahaentesar214@gmail.com
bDepartment of Chemistry, Faculty of Science, King Khalid University, Abha 9004, Saudi Arabia
cNuclear Materials Authority, P. O. Box 530, El Maddi, Cairo, Egypt
First published on 25th September 2025
Despite the growing interest in biofunctionalized magnetic nanocomposites, there is limited research on integrating nickel with gum Arabic (GA) to enhance selectivity and stability for heavy metal removal in complex aqueous systems. A green, biofunctional magnetite–nickel nanocomposite (GA-NiMMPs) was synthesized via a modified co-precipitation method using GA as a natural stabilizer and surface modifier. Cadmium and copper were selected as target contaminants because they are among the most hazardous and prevalent heavy metals in industrial effluents, frequently co-existing in electroplating, mining, and metal-finishing wastewaters. GA-NiMMPs exhibited a surface area of 29.32 m2 g−1, a pore diameter of 1.91 nm, and a zeta potential of −22.5 mV, favoring the adsorption of divalent metal ions and enabling their simultaneous removal. Batch adsorption experiments were conducted using 50 mg per L initial concentrations, varying pH, dosage, and contact time. High removal efficiencies of 93.5% for Cd(II) and 89.0% for Cu(II), with maximum capacities of 30.4 and 27.7 mg g−1, respectively, were obtained with optimal performance at pH 6.0, 4.0 g per L dosage, and 240 min contact time. Kinetics follow a pseudo-second-order model, while isotherms fit Langmuir and Sips models, indicating monolayer chemisorption as described by the Langmuir model, alongside surface heterogeneity captured by the Sips model. The material showed higher affinity for Cd(II), achieving 64% selectivity in binary and 58% in multicomponent systems. These results underscore GA-NiMMPs as a promising bio-derived adsorbent for sustainable Cd(II) and Cu(II) remediation in complex waters.
Traditional heavy metal removal technologies—including chemical precipitation,6 ion exchange,7 membrane filtration,8 and solvent extraction9—have been widely applied but often suffer from drawbacks such as high operational costs, energy demand, sludge generation, and limited metal selectivity. For example, chemical precipitation produces hazardous sludge requiring costly post-treatment, membrane systems are prone to fouling and demand frequent maintenance, ion exchange resins are efficient but expensive, and solvent extraction may introduce organic solvent residues that cause secondary pollution.1–4 These limitations underscore the need for alternative solutions that are both environmentally benign and economically viable.
As a result, there is a growing demand for cost-effective, environmentally sustainable, and highly efficient materials for heavy metal removal. Adsorption has emerged as a promising approach due to its simplicity, affordability, and ability to remove trace-level contaminants with high precision.1–4 Recent research has explored a wide range of adsorbents, including natural sorbents, synthetic polymers, and magnetic nanoparticles. These materials can be grouped by type, such as: (i) carbon-based biochars (e.g., from willow wood and cattle manure,10 and Bael fruit shell-derived carbon2), (ii) biopolymer-based composites (e.g., chitosan-vermiculite,11 Ag-MOF/chitosan sponge12), (iii) inorganic frameworks (e.g., zeolites13), and (iv) advanced materials like sulfidized nanoscale zero-valent iron (S-nZVI) and sulfur-functionalized activated carbons.14–16 While informative, long listings of such sorbents often obscure comparative insight; here we focus on materials that combine high affinity with reusability and recoverability.
The development of advanced adsorbents with high surface area, enhanced binding affinity, and superior selectivity is crucial for optimizing adsorption efficiency. Magnetite-based nanoparticles have gained significant attention due to their unique magnetic properties, ease of recovery, and potential for functionalization.17–19 The use of plant-derived polysaccharides—such as gum Arabic (GA)—in the synthesis of these materials is a compelling strategy for biomass valorization. Gum Arabic is a naturally abundant, biodegradable exudate obtained from Acacia species, composed primarily of arabinogalactan polysaccharides with ample hydroxyl and carboxyl groups suitable for metal ion binding.20–22 It functions as a stabilizing agent, functionalizing ligand, and dispersant, enabling uniform particle distribution and preventing agglomeration.20–22 Recent studies have demonstrated the effectiveness of GA-modified sorbents in toxic elements uptake. GA-magnetite nanoparticles (GA/MNPs) efficiently adsorb Pb(II) (50.5 mg g−1),23 while GA-functionalized magnetic nanoparticles (GA-MNPs) achieve a 96.3% removal rate of ciprofloxacin (CIP).22 GA-poly(vinyl alcohol) (PVA) hydrogels exhibit strong affinities for Pb(II), Cu(II), Co(II), and Cd(II), making them promising for industrial wastewater treatment.21 Additionally, GA-alginate beads, with and without magnetite, show enhanced adsorption of Cu(II), Cd(II), and Pb(II).24 Moreover, composites of GA with TiO2, Al2O3, or MgO nanoparticles exhibit remarkable performance in multi-metal adsorption, benefiting from high surface areas and active binding sites.20 Nickel incorporation into the magnetite lattice introduces lattice distortions and enhances surface redox properties, thereby increasing the number and diversity of active sites available for metal ion binding. This modification facilitates improved adsorption capacity and selectivity, particularly for divalent ions such as Cd(II) and Cu(II).1,25–28 Recent studies have demonstrated that nickel-based ferrites and Ni-functionalized oxides exhibit enhanced sorption performance due to strong complexation with oxygen-donor groups and favorable electronic configurations.25,26
Motivated by these findings, the present study introduces a novel nanoparticles (GA-NiMMPs) synthesized via a green, low-temperature co-precipitation method, combining the biofunctionality of gum Arabic with the magnetic responsiveness of nickel-doped magnetite. To the best of our knowledge, this is one of the first studies to integrate gum Arabic and nickel into a mesoporous magnetite platform and validate its performance in both synthetic and real wastewater matrices. The nanoparticles was extensively characterized to confirm its structure, morphology, and functional surface chemistry. Its adsorption performance was evaluated for Cd(II) and Cu(II) ions in single-ion systems, under varying operational parameters, including pH, contact time, adsorbent dosage, and initial ion concentration. Kinetic, isotherm, and thermodynamic models were applied to describe the adsorption mechanisms. The real wastewater matrix used in this study contained multiple heavy metal ions; although dye residues were present, their removal was not quantitatively addressed. This work provides a green synthesis route for a multifunctional, biomass-derived adsorbent, with validated performance in both controlled and real-world treatment scenarios.
Fourier-transform infrared spectroscopy (FTIR) was employed to identify the surface functional groups and bonding interactions. Spectra were recorded in the range of 400–4000 cm−1 using a Bruker ALPHA spectrometer in attenuated total reflectance (ATR) mode, revealing characteristic vibrational bands corresponding to hydroxyl (–OH), carboxyl (–COOH), and metal–oxygen bonds (Fe–O and Ni–O). The crystallographic structure of the nanoparticles was investigated via X-ray diffraction (XRD) using a PANalytical X'Pert PRO diffractometer equipped with Cu Kα radiation (λ = 1.5406 Å), scanned over a 2θ range of 10–80°. The resulting diffraction patterns confirmed the presence of magnetite (Fe3O4), nickel ferrite (NiFe2O4), and minor NiO phases, indicating successful formation of a mixed-phase spinel structure.
Textural properties, including specific surface area, pore size, and pore volume, were analyzed by nitrogen adsorption–desorption isotherms measured at 77 K using a Micromeritics ASAP 2020 analyzer. The Brunauer–Emmett–Teller (BET) method was used to calculate the surface area, while the Barrett–Joyner–Halenda (BJH) method determined pore size distribution and mesoporosity. Samples were degassed at 150 °C for 12 hours prior to measurement to eliminate moisture and surface contaminants. Additionally, the hydrodynamic particle size and surface charge (zeta potential) of the GA-NiMMPs dispersed in deionized water were determined by dynamic light scattering (DLS) and electrophoretic light scattering using a Malvern Zetasizer Nano ZS. These measurements provided insight into the colloidal stability and electrostatic behavior of the GA-NiMMPs in aqueous environments, which are critical for adsorption performance.
Adsorption kinetics were studied by varying shaking times (5–600 minutes), while isotherm performance was evaluated at initial metal concentrations of 20–200 mg L−1. The Lagergren pseudo-first-order, pseudo-second-order, and intraparticle diffusion (Weber and Morris) models were used to analyze kinetics. The Langmuir, Freundlich, Temkin, and Sips models were applied to describe adsorption isotherms. The non-linear forms of these kinetic and isotherm models are provided in Table S1.29–34 The fitting of the models was evaluated based on the Chi-square (x2) and the coefficient of determination (R2), as described in Table S1.33,34 Thermodynamic parameters, including standard Gibbs free energy change (ΔG°), standard enthalpy change (ΔH°), and standard entropy change (ΔS°), were calculated at temperatures ranging from 25 to 50 ± 1 °C using the equations listed in Table S1.35,36 These parameters provided insights into the spontaneity and feasibility of the adsorption process.
In all experiments, the initial (C0) and final residual (Ce) concentrations of Cu(II) and Cd(II) were measured using an Atomic Absorption Spectrometer (GBC 932 AA, UK). Samples were filtered through 0.22 μm filters before analysis. Experiments were conducted in triplicate, and results with a relative error of ≤5% were considered acceptable. The adsorption efficiency (E%), adsorption capacity (qe, mg g−1), and distribution coefficient (Kd) were calculated using the following equations:
![]() | (1) |
![]() | (2) |
![]() | (3) |
The co-precipitation mechanism involves the hydrolysis of Fe3+, Fe2+, and Ni2+ ions in the presence of ammonium hydroxide, leading to the in situ formation of mixed metal hydroxides. Upon heating and aging, these hydroxides transform into a bimetallic spinel structure composed of magnetite (Fe3O4) and nickel ferrite (NiFe2O4). The alkaline pH (∼10) facilitates the formation of Ni(OH)2, which subsequently integrates into the Fe–O lattice during the growth process. This mechanism is consistent with the XRD results, confirming crystalline phases of both Fe3O4 and NiFe2O4. GA therefore plays a dual role—acting both as a reducing and stabilizing agent during synthesis and as a functional surface modifier post-synthesis. It introduces hydroxyl (–OH) and carboxyl (–COOH) functional groups, which are key for complexation and electrostatic interaction with divalent metal ions.25,36 This surface functionalization also facilitates water dispersibility and colloidal stability, crucial attributes for sorbent performance in aqueous environments. The elemental composition measured by EDX (Fe ∼64.23%, Ni ∼6.41%, O ∼21.61%, and C ∼8.2 wt%) thus validates the successful incorporation of GA and its functional role in modifying the composite surface.
Fourier transform infrared (FTIR) spectra (Fig. 1-II) provide clear evidence of the chemical functionalities present on the GA-NiMMPs surface. The distinct absorption band observed between 550–600 cm−1 corresponds to Fe–O stretching vibrations, confirming the formation of magnetite and nickel ferrite phases.22–24 The broad band centered at ∼3400 cm−1 is attributed to O–H stretching vibrations of hydroxyl groups and adsorbed water molecules,22–24 which are crucial for surface complexation and hydrogen bonding during adsorption. Well-defined peaks at ∼1700 cm−1 (CO stretching) and ∼1400 cm−1 (COO− symmetric stretching) confirm the incorporation of carboxyl functionalities derived from gum Arabic, serving as strong binding sites for Cd(II) and Cu(II). A weaker shoulder at ∼1250 cm−1 (C–O stretching), together with the feature around 1375 cm−1 (C–H bending), further demonstrates the retention of polysaccharide groups from GA on the composite surface.22–24 These functional moieties enable multiple adsorption pathways including electrostatic attraction, ion exchange, and hydrogen bonding, enhancing the versatility and binding specificity of the GA-NiMMPs.
X-ray diffraction (XRD) analysis (Fig. 1-III) revealed sharp diffraction peaks at 2θ = 30.1°, 35.5°, 43.2°, 53.5°, and 62.7°, corresponding to the (220), (311), (400), (422), and (511) planes of Fe3O4 (JCPDS 19-0629), confirming magnetite as the predominant phase.22–24 Superimposed reflections at 37.2° and 57.5° correspond to Ni-ferrite (Ni0.4Fe2.6O4) (JCPDS 10-0325), validating nickel integration within the spinel structure.25,26 Additionally, minor peaks corresponding to NiO (JCPDS 47-1049) suggest partial segregation of Ni as a discrete oxide phase, likely contributing to enhanced electron transport and sorption dynamics.25,26 Broadening of the XRD peaks, particularly at 35.5° and 43.2°, indicates nanoscale crystallite domains and possible lattice strain resulting from metal ion substitution and oxygen vacancy formation. These crystalline and sub-crystalline features enhance the number of active surface sites and increase the reactivity of the sorbent.25,26 Partial oxidation of Fe3O4 to maghemite may have occurred during synthesis or post-processing, further influencing the electronic configuration of the surface. Collectively, the hybrid crystalline structure comprising Fe3O4, Ni-ferrite, and NiO contributes to structural robustness, magnetic responsiveness, and enhanced metal ion affinity.
![]() | ||
Fig. 2 (I) SEM image of GA–NiMMPs, (II) DLS particle size distribution, and (III) nitrogen adsorption–desorption isotherm. |
Dynamic light scattering (DLS) analysis (Fig. 2-II and Table 1) revealed an average hydrodynamic diameter of 688.3 nm, with 75% of the particles measuring below 828.5 nm. The presence of a broad size distribution suggests that while most particles are in the sub-micron range, some degree of agglomeration may have occurred due to the nature of bio-assisted synthesis. Nevertheless, the average particle size supports a favorable surface area-to-volume ratio, which is critical for adsorption kinetics.22–24 The particle size results align with SEM observations (Fig. 2-II) and confirm that the GA-NiMNPs exhibit a well-dispersed, moderately polydisperse morphology. The relatively small size ensures efficient diffusion and interaction with metal ions in aqueous environments. Moreover, the size distribution may be attributed to lattice distortion induced by Ni doping and surface functionalization, highlighting the influence of synthesis parameters on final particle characteristics.
DLS analysis, nm | Zeta potential, mV |
---|---|
688.3 | −21.4 |
Zeta potential measurements indicated a surface charge of −21.32 mV (Table 1), reflecting moderate colloidal stability and a net negative surface potential. This negative charge results from the deprotonation of carboxyl and hydroxyl groups contributed by GA, as confirmed by FTIR analysis (Fig. 1-II). It facilitates electrostatic attraction of positively charged heavy metal ions, enhancing the adsorption capacity of the GA-NiMMPs.22–24 This moderately negative zeta potential also helps prevent aggregation by promoting electrostatic repulsion among particles in suspension, preserving a high degree of dispersion in aqueous environments.22–24 Together with the porous structure and small particle size, this surface potential supports efficient ion accessibility and high sorption efficiency, further validating the functional role of GA in enhancing material stability and performance.
Nitrogen adsorption–desorption isotherms (Fig. 2-III) exhibited a Type IV pattern, according to the IUPAC classification with a pronounced H3 hysteresis loop at relative pressures (P/P0) > 0.4, characteristic of mesoporous materials.27 This profile suggests the existence of slit-shaped pores and indicates well-defined mesostructures, often formed by the stacking of plate-like particles or irregular pore assemblies. The BET surface area of 29.32 m2 g−1 demonstrates a moderately high external surface area, sufficient to support rapid metal ion uptake. The C constant value of 7.48 reflects favorable interactions between nitrogen molecules and the composite surface, implying a high density of adsorption sites. This value correlates with the presence of oxygen-containing functional groups and the surface roughness observed in SEM images. The porosity, in conjunction with surface chemistry, creates an ideal microenvironment for adsorption processes, especially in aqueous-phase applications. BJH pore analysis (Fig. S2) confirmed the mesoporous nature of the GA-NiMNPs, showing an average pore radius of 1.91 nm and a total pore volume of 0.0577 cm3 g−1. The cumulative pore volume, as determined by DFT, was 0.0624 cm3 g−1, with a pore width distribution primarily in the 2–4 nm range. These properties are summarized in Table 2. These results affirm the hierarchical mesoporous framework, which enables effective diffusion and retention of metal ions. Such a pore structure ensures that the material maintains a balance between surface area and pore accessibility, essential for adsorption under dynamic flow conditions. The mesopores, supported by GA-induced particle spacing and nickel-enhanced lattice structuring, contribute to the GA-NiMMPs overall adsorption kinetics and equilibrium behavior. This configuration promotes not only rapid metal capture but also ease of regeneration, enhancing the reusability of the biomass-derived sorbent for environmental applications.
SBET (m2 g−1) | VP (cm3 g−1) | DH (nm) |
---|---|---|
29.32 | 0.0624 | 1.91 |
Solution pH plays a pivotal role in adsorption behavior by modulating both the ionization state of surface functional groups on the sorbent and the speciation of metal ions in solution.11–14 As shown in Fig. 3-I, the adsorption efficiency of GA-NiMMPs for Cd(II) and Cu(II) increased progressively with rising pH and reached a maximum at pH 5–6, where removal efficiencies were 90.4% for Cd(II) and 85.1% for Cu(II). This trend reflects the combined influence of surface charge characteristics and metal ion speciation. At strongly acidic pH values (2–3), extensive protonation of carboxyl and hydroxyl groups on gum Arabic suppresses their ability to coordinate with Cd(II) and Cu(II), while the zeta potential measurement (−21.32 mV) further indicates reduced electrostatic attraction under such proton-rich conditions. At near-neutral pH (5–6), deprotonation of –COOH and –OH groups increases the density of negatively charged sites, which enhances electrostatic attraction toward the divalent cations (Cd2+, Cu2+). In addition, surface Ni–O and Fe–O sites originating from the spinel lattice of Ni-ferrite and magnetite play a direct role by acting as electron-donor centers, enabling inner-sphere complexation with the metal ions. The synergistic effect of GA-derived functional groups and lattice metal–oxygen sites explains the high adsorption performance observed in this range. This interpretation is fully consistent with FTIR spectra (Fig. 1-II), which highlight the presence of carboxylate and hydroxyl functionalities, as well as with XRD evidence confirming Ni incorporation into the Fe3O4 structure. At alkaline pH values above 7, the efficiency declined, which can be attributed to the onset of metal hydroxide precipitation (Cd(OH)2 and Cu(OH)2). Under these conditions, the removal is no longer governed solely by adsorption but also by precipitation processes, making the measured uptake less representative of surface binding.15–18 Therefore, the optimal adsorption window for GA-NiMMPs lies between pH 5 and 6, where the combined contribution of GA functional groups and Ni/Fe oxide lattice sites maximizes electrostatic attraction and coordination with Cd(II) and Cu(II) ions.20–22
Adsorption kinetics showed a rapid uptake during the initial 120 minutes, where removal efficiencies reached 84.0% for Cd(II) and 77.0% for Cu(II), and gradually approached equilibrium after 600 minutes, stabilizing at 91.0% and 84.9%, respectively (Fig. 3-II). The fast initial phase reflects the abundance of unoccupied active sites and the accessibility provided by the mesoporous structure.10–12 As equilibrium is approached, the adsorption rate slows due to site saturation,14–16 which is supported by the material's high surface area and mesoporous structure (BET surface area = 29.32 m2 g−1; Table 2), providing accessible pathways for initial rapid uptake followed by diffusion-limited behavior.
The adsorption efficiency improved significantly with increasing GA-NiMNPs dosage, reaching a plateau at 5.0 g L−1, with Cd(II) and Cu(II) removal efficiencies of 96.4% and 92.1%, respectively (Fig. 3-III). The increase in removal corresponds to the greater availability of active binding sites at higher dosages.10–12 Beyond 4.0 g L−1, only marginal gains were observed, suggesting near-saturation of the system.14–16 This saturation was further validated through changes in the partition coefficient (logKd). At low dosage (1.0 g L−1), the log
Kd values were negative, reflecting poor adsorption due to limited sites.14–16 However, values turned positive at 4.0 g L−1 (0.44 for Cd(II) and 0.14 for Cu(II)), confirming enhanced metal uptake (Fig. 3-IV).14–16 The difference in Cd(II) and Cu(II) partitioning reflects their respective physicochemical behaviors and affinity for the sorbent.
Initial concentrations ranging from 20 to 200 mg L−1 revealed that GA-NiMMPs perform exceptionally well at low to moderate concentrations, achieving removal efficiencies of 93.5% for Cd(II) and 89.0% for Cu(II) at 20 mg L−1 (Fig. 3-V). At higher concentrations, efficiency declined due to site saturation, with values decreasing to 57.0% (Cd) and 50.0% (Cu) at 200 mg L−1. This trend reflects a capacity-limited system in which the abundance of available sites at 20–50 mg L−1 ensures high binding, while saturation dominates at higher loadings.10–12 Correspondingly, the adsorption capacity (qe) increased steadily with initial concentration: from 4.7 mg per g (Cd) and 4.5 mg per g (Cu) at 20 mg L−1, to 11.4 mg per g (Cd) and 10.6 mg per g (Cu) at 50 mg L−1. Beyond this point, qe continued to rise with increasing concentration, reaching 28.5 mg per g (Cd) and 25.0 mg per g (Cu) at 200 mg L−1, although with lower removal efficiency. Therefore, 50 mg L−1 was selected as the optimum initial concentration, since it simultaneously provided high removal efficiencies (>90%) and the maximum adsorption capacity observed under the studied conditions (11.4 mg g−1 for Cd and 10.6 mg g−1 for Cu). The structural features of GA-NiMMPs—including mesoporosity and abundant surface functional groups—aid in sustaining performance under moderate loading, while capacity-driven binding dominates at higher concentrations.
Temperature-dependent adsorption tests showed a marked increase in efficiency with rising temperature (25–50 °C), confirming the endothermic nature of the process.10–12 For Cu(II), efficiency rose from 84.3% to 92.1%, and for Cd(II), from 90.8% to 96.4% (Fig. 3-VI). This enhancement is attributed to increased kinetic energy and reduced hydration shell size, facilitating stronger sorbate–sorbent interactions.10–12 FTIR and XRD analysis (Fig. 1-II and III respectively) support that higher temperatures enhance functional group reactivity and promote stronger complexation.14–16 These results highlight the suitability of GA-NiMNPs for thermally variable environments and industrial effluents.
Collectively, these findings confirm that GA-NiMNPs are highly efficient, biomass-derived adsorbents for Cd(II) and Cu(II), with performance directly tied to their structural, surface, and textural attributes. The synergy between porosity, surface charge, and bio-functionalization facilitates multi-modal interactions—electrostatic, coordinative, and diffusional—offering a robust platform for sustainable water purification. Similar adsorption performance for copper(II) and cadmium(II) ions from aqueous solutions has been observed using various adsorbents, including CuO-modified ceramic membranes,4 chitosan–vermiculite composites,11 Ag-MOF/chitosan composite sponges,12 exhausted copper slag-supported sulfidized nanoscale zerovalent iron,14 dithiocarbamate-functionalized activated carbon,15 Cystoseira sedoide algae,16 and ethylenediamine-functionalized chelating resin 28. These findings provide valuable insights into optimizing process conditions and highlight the potential of this adsorbent for practical applications in wastewater treatment and environmental remediation, particularly in complex aqueous matrices.
Cd(II) | Cu(II) | |
---|---|---|
Pseudo first-order model | ||
q1 (mg g−1) | 10.5 | 10.0 |
k1 (min−1) | 0.103 | 0.069 |
R2 | 0.91 | 0.95 |
X2 | 0.40 | 0.49 |
![]() |
||
Pseudo second-order model | ||
q2 (mg g−1) | 11.1 | 10.4 |
k2 (min−1) | 0.013 | 0.011 |
h (mol g−1 h−1) | 1.7 | 1.2 |
t1/2 (h) | 6.7 | 8.9 |
R2 | 0.99 | 0.99 |
X2 | 0.02 | 0.05 |
![]() |
||
Elovich model | ||
a (mg g−1 min−1) | 0.80 | 0.65 |
β (g mg−1) | 29.08 | 4.32 |
R2 | 0.90 | 0.91 |
X2 | 0.6 | 0.8 |
The pseudo-first-order model yielded relatively lower correlation coefficients (R2 = 0.91 for Cd(II) and 0.95 for Cu(II)) and higher error values (χ2 = 0.40 for Cd(II) and 0.49 for Cu(II)) (Table 3), indicating limited applicability for these systems. In contrast, the pseudo-second-order model exhibited an excellent fit, with R2 values of 0.99 for both Cd(II) and Cu(II) and significantly lower error values (χ2 = 0.02 for Cd(II) and 0.05 for Cu(II)). The superior fit of this model suggests that chemisorption governs the adsorption process, involving strong chemical interactions between the metal ions and the functional groups on the GA-NiMMPs surface.11–15 These kinetic trends align well with findings from FTIR analysis (Fig. 1-II), which confirms the presence of –OH, –COO−, and –CO moieties on the GA-NiMMPs surface. This validation reinforces the chemisorptive nature of the adsorption system, highlighting the role of specific functional groups in the binding process. Furthermore, the pseudo-second-order model also revealed higher adsorption capacities (q2 = 11.1 mg g−1 for Cd(II) and 10.4 mg g−1 for Cu(II)), consistent with the experimental data. The rate constants (k2 = 0.013 min−1 for Cd(II) and 0.011 min−1 for Cu(II)), the initial sorption rates (h = 1.7 mol g−1 h−1 for Cd(II) and 1.2 mol g−1 h−1 for Cu(II)), and half-life values (t1/2 = 6.7 h for Cd(II) and 8.9 h for Cu(II)) suggest a relatively fast adsorption process, with Cd(II) being adsorbed slightly faster than Cu(II). This difference is likely due to the smaller ionic radius and lower hydration energy of Cd(II), which facilitates faster diffusion and stronger binding.10,12
The Elovich model, which describes chemisorption on highly heterogeneous surfaces, also provided a good fit to the experimental data, with R2 values of 0.90 for Cd(II) and 0.91 for Cu(II). The high initial adsorption rate (α = 0.80 mg g−1 min−1 for Cd(II) and 0.65 mg g−1 min−1 for Cu(II)) and the desorption constant (β = 29.08 g mg−1 for Cd(II) and 4.32 g mg−1 for Cu(II)) suggest that as evidenced by the variation in pore size and surface morphology seen in SEM and BET data. Such physical heterogeneity is consistent with Elovich assumptions of chemisorption across energetically diverse sites.10,12,38 This matches well with the high surface roughness and pore diversity observed via SEM and BET analysis (Fig. 2-I, and Table 2).
The Weber–Morris model indicates that adsorption follows a multi-step pathway, beginning with external surface adsorption, followed by intraparticle diffusion, and finally reaching equilibrium.11–15 The initial stage, characterized by a steep slope, signifies the rapid adsorption of metal ions onto the external surface of GA-NiMMPs. In the second stage, metal ions gradually diffuse into the interior pores of the adsorbent. The final stage represents equilibrium, at which point the adsorption rate slows due to site saturation.11–15 The Weber–Morris model parameters (Table 4) reveal that the intraparticle diffusion rate constants (ki = 0.16 mg g−1 min−1/2 for Cd(II) and Cu(II)) and the intercept values (C = 7.5 for Cd(II) and 6.8 for Cu(II)) suggest that intraparticle diffusion is not the only rate-limiting step. These findings, when combined with the high correlation of the pseudo-second-order model and the nonlinear qt vs. t1/2 plot (Fig. 5), confirm that adsorption occurs via a dual mechanism involving both surface adsorption and intraparticle diffusion. The high R2 values (0.98) affirm the model's applicability, and the multi-linearity of the qt vs. t1/2 plots (Fig. 5) supports this interpretation. These results underscore that the adsorption mechanism is primarily chemisorptive, supplemented by diffusion and governed by the sorbent's heterogeneous mesoporous network.
Cd(II) | Cu(II) | ||
---|---|---|---|
Weber and Morris model | ki (mg g−1 min−1/2) | 0.16 | 0.16 |
C | 7.5 | 6.8 | |
R2 | 0.98 | 0.98 |
![]() | ||
Fig. 5 W–M model illustrate for uptake process (solution pH 6.01; ligand dose of 4.0 g L−1; original concentration of 50 mg L−1; temperature of 25 ± 1 °C). |
Overall, the kinetic evaluation underscores the effectiveness of GA-NiMNPs in enabling rapid and efficient metal ion removal. The fast initial adsorption, governed by abundant functional sites, is complemented by a robust porous framework that facilitates prolonged interaction and diffusion. These dual characteristics—chemical specificity through GA-derived functional groups and mesoporosity through nanoarchitectural control—demonstrate a highly synergistic adsorption process. This mechanistic clarity enhances the material's application potential, particularly for dynamic water treatment systems where contact time and efficiency are critical performance parameters. Comparable kinetic patterns have been documented for copper(II) and cadmium(II) ion sequestration across diverse adsorbent systems, including CuO-modified ceramic membranes,4 chitosan–vermiculite composites,11 Ag-MOF/chitosan composite sponges,12 exhausted copper slag-supported sulfidized nanoscale zerovalent iron,14 dithiocarbamate-functionalized activated carbon,15 Cystoseira sedoide algae,16 and ethylenediamine-functionalized chelating resin 28.
![]() | ||
Fig. 6 Isotherm profile for the sorption of (I) Cd(II) and (II) Cu(II) ions from aqueous onto GA–NiMMPs (time is 240 min; ligand dose of 4.0 g L−1; pH 5.9; temperature of 25 ± 1 °C). |
Cd(II) | Cu(II) | |
---|---|---|
Langmuir model | ||
qm (mg g−1) | 30.4 | 27.7 |
KL (L mg−1) | 0.139 | 0.092 |
R2 | 0.99 | 0.99 |
X2 | 0.04 | 0.06 |
![]() |
||
Freundlich model | ||
1/nF | 0.3 | 0.3 |
kF (mg g−1) (mg L−1) | 7.9 | 5.8 |
R2 | 0.93 | 0.92 |
X2 | 3.2 | 2.5 |
![]() |
||
Temkin model | ||
bT (J mol−1) | 414.3 | 428.9 |
AT (L g−1) | 1.7 | 1.0 |
R2 | 0.99 | 0.99 |
X2 | 0.4 | 0.3 |
![]() |
||
Sips model | ||
qS (mg g−1) | 30.8 | 27.5 |
kS (L mg−1) | 0.135 | 0.094 |
mS | 0.99 | 1.04 |
R2 | 0.99 | 0.99 |
X2 | 0.04 | 0.04 |
The Langmuir model, which assumes monolayer adsorption on a homogeneous surface with a finite number of identical sites, provided an excellent fit to the experimental data, with R2 values of 0.99 for both Cd(II) and Cu(II). The maximum monolayer adsorption capacities (qm) were 30.4 mg g−1 for Cd(II) and 27.7 mg g−1 for Cu(II), consistent with the high density of accessible binding sites inferred from BET and SEM analysis. The Langmuir constant (kL), which reflects the adsorption affinity, was higher for Cd(II) (0.139 L mg−1) compared to Cu(II) (0.092 L mg−1), suggesting stronger interactions between Cd(II) ions and the adsorbent surface.10,15,16,36 These results validate the homogenous nature of binding sites provided by GA-functionalized magnetic nanoparticles and support strong electrostatic and complexation interactions. Comparable isotherm behavior, characterized by homogeneous and monolayer adsorption processes, was observed for the removal of copper(II), and cadmium(II) from aqueous solutions using various adsorbents. These include CuO-modified ceramic membranes,4 chitosan–vermiculite composites,11 Ag-MOF/chitosan composite sponges,12 exhausted copper slag-supported sulfidized nanoscale zerovalent iron,14 dithiocarbamate-functionalized activated carbon,15 Cystoseira sedoide algae,16 and ethylenediamine-functionalized chelating resin.28 The nature of the sorption system can be assessed using the dimensionless equilibrium parameter (RL), defined by the equation RL = 1/(1 + KLC0). This variable classifies sorption behavior into distinct categories: linear (RL = 1), irreversible (RL = 0), favorable (0 < RL < 1), or unfavorable (RL > 1).33,34 As demonstrated in Fig. S1, the obtained Rl values confirm that the sorption process is favorable. Specifically, Cd(II) exhibited RL values ranging from 0.03 to 0.26, while Cu(II) showed values between 0.05 and 0.35 across all tested initial concentrations.
The Freundlich model, which describes adsorption on heterogeneous surfaces with non-uniform energy distribution, also provided a good fit to the experimental data, with R2 values of 0.93 for Cd(II) and 0.92 for Cu(II). The Freundlich constant (kF), which indicates adsorption capacity, was higher for Cd(II) (7.9 mg g−1) compared to Cu(II) (5.8 mg g−1), further supporting the stronger affinity of GA-NiMMPs for Cd(II) ions. The heterogeneity factor (1/nF) was 0.3 for both metal ions, indicating favorable adsorption conditions and a high degree of surface heterogeneity.31–33 This matches the Elovich and Weber–Morris model results and is supported by the presence of diverse surface functionalities.10,15,16,36 The Temkin model, which accounts for the effects of adsorbate–adsorbent interactions on adsorption heat, provided a strong fit to the experimental data, with R2 values of 0.99 for both Cd(II) and Cu(II). The Temkin heat of adsorption (bT = 414.3 J mol−1 for Cd(II), 428.9 J mol−1 for Cu(II)) and equilibrium binding constants (AT = 1.7 L g−1 for Cd(II), 1.0 L g−1 for Cu(II)) suggest moderate physisorption alongside chemisorption, particularly at higher concentrations where multilayer interactions are more likely.12,15,16,29
Lastly, Sips model, which combines the features of the Langmuir and Freundlich models, provided an excellent fit to the experimental data, with R2 values of 0.99 for both Cd(II) and Cu(II). The maximum adsorption capacities (qS) were 30.8 mg g−1 for Cd(II) and 27.5 mg g−1 for Cu(II), aligned with Langmuir values. The Sips constant (kS) was higher for Cd(II) (0.135 L mg−1) compared to Cu(II) (0.094 L mg−1), further supporting the stronger affinity of GA-NiMMPs for Cd(II) ions. The heterogeneity parameter (mS) was close to 1.0 for both metal ions, affirmed a predominantly monolayer mechanism, consistent with the Langmuir model.31,32 These findings collectively demonstrate that adsorption occurs via both homogeneous and heterogeneous interactions, consistent with the material's structural and surface features.
Overall, the isotherm modeling results provide a comprehensive understanding of the adsorption mechanism, validating the structural-functional synergy of GA-NiMMPs. The strong agreement with the Langmuir model highlights the dominance of monolayer coverage on well-defined active sites, while the satisfactory fit with Freundlich and Sips models acknowledges the surface heterogeneity introduced by gum Arabic functionalization. Additionally, Temkin analysis supports moderate interaction energies, suggesting the involvement of both physical and chemical sorption processes. This multipronged adsorption behavior aligns with the hybrid composition of the sorbent, where the natural polymer framework, transition metal oxide phases, and mesoporous architecture collectively influence performance. These findings are further reinforced by the Weber–Morris kinetic model, which reveals a complex interplay of binding modes. The material's effectiveness can be attributed to its mesoporous structure, high surface area, and functionalized chemistry—particularly the presence of key functional groups (–COOH, –OH, and –CO) that facilitate chemisorption, a negatively charged surface promoting electrostatic interactions, and a porous structure providing extensive binding sites. Together, these properties establish GA-NiMMPs as an efficient and robust adsorbent for heavy metal removal, offering valuable insights for the development of advanced, sustainable water treatment materials.
The higher adsorption tendency of GA-NiMMPs for Cd(II) over Cu(II) arises from a combination of physicochemical properties and coordination chemistry principles. Several key factors, including ionic radius, hydration energy, electronegativity, charge density, and the nature of functional groups on the adsorbent surface, play a crucial role in determining the adsorption affinity of metal ions.4,10,12,33,39–43 These factors collectively influence the stability and strength of interactions between metal ions and the GA-NiMMPs, ultimately dictating the preferential adsorption of Cd(II).
One significant parameter contributing to this selectivity is the hydrated ionic radius. Cd(II) has a larger ionic radius (0.426 nm) compared to Cu(II) (0.295 nm),4,39 which facilitates stronger and more effective coordination with the functional groups present on the GA-NiMMPs. This larger ionic radius allows Cd(II) to form more stable interactions with carboxyl (–COO−), and hydroxyl (–OH) groups, which have been confirmed through FTIR analysis (Fig. 1-II). Moreover, Cd(II) generally exhibits a higher coordination number (typically 6) than Cu(II), which often forms square-planar or distorted octahedral complexes with a coordination number of 4–6.4,39 The ability of Cd(II) to engage with a greater number of binding sites on the adsorbent further enhances its adsorption capacity. Hydration energy plays a critical role in the adsorption process by determining how easily metal ions shed their hydration shells to interact with the adsorbent surface. Cd(II) has a lower hydration energy (−1807 kJ mol−1) than Cu(II) (−2100 kJ mol−1), meaning that Cd(II) ions are less strongly solvated in aqueous solutions.33,41,42 The weaker solvation of Cd(II) allows for easier dehydration and subsequent interaction with the adsorbent's active sites. In contrast, the higher hydration energy of Cu(II) results in stronger solvation, making it less available for adsorption and limiting its ability to effectively bind with the functional groups on GA-NiMMPs.
Electronegativity and charge density also influence adsorption behavior. Cd(II) has a lower electronegativity (1.69 on the Pauling scale) than Cu(II) (1.90), which contributes to reduced electrostatic repulsion between Cd(II) ions and the electron-rich functional groups on the GA-NiMMPs surface.4,21,33,42 This lower repulsion allows Cd(II) to establish stronger and more stable complexes with the adsorbent. Additionally, Cd(II) possesses a lower charge density (59C mm−3) than Cu(II) (71C mm−3), further reducing electrostatic repulsion and promoting adsorption. The relatively lower charge density of Cd(II) facilitates stronger binding with the adsorbent's functional groups, whereas the higher charge density of Cu(II) results in stronger electrostatic interactions with surrounding water molecules, decreasing its adsorption efficiency.43 From a coordination chemistry perspective, the Hard-Soft Acid-Base (HSAB) theory provides further insight into the preferential adsorption of Cd(II). Cadmium(II) is classified as a softer Lewis acid, whereas Cu(II) is a borderline Lewis acid.33,44 The functional groups present on GA-NiMMPs, such as hydroxyl (–OH), and carboxyl (–COO−), behave as intermediate to soft bases. As a result, Cd(II), being a softer acid, exhibits a stronger affinity for these functional groups due to more favorable orbital overlap and bonding interactions. In contrast, Cu(II), with its borderline acid character, does not interact as effectively with these groups, leading to reduced adsorption.
The structural and chemical properties of GA-NiMMPs further enhance the selective adsorption of Cd(II). The presence of multiple active functional groups provides diverse binding sites, enabling complexation, electrostatic attraction, and hydrogen bonding with metal ions. The larger ionic radius and lower hydration energy of Cd(II) allow it to interact more effectively with these functional groups, forming stable complexes that improve adsorption efficiency. On the other hand, Cu(II), with its smaller ionic radius and higher hydration energy, encounters steric and solvation-related challenges that hinder its ability to form similarly stable interactions.
Overall, the superior adsorption of Cd(II) onto GA-NiMMPs can be attributed to its larger ionic radius, lower hydration energy, reduced charge density, and favorable interactions with functional groups based on HSAB theory. These factors collectively enhance the selectivity of GA-NiMMPs for Cd(II) over Cu(II), making it an effective adsorbent for cadmium removal in environmental remediation applications. This preferential adsorption of Cd(II) over Cu(II) from aqueous solutions has also been observed with other sorbents, including Ag-MOF/chitosan composite sponges,12 Fly Ash-Based Linde F(K) Zeolite,13 exhausted copper slag-supported sulfidized nanoscale zerovalent iron,14 dithiocarbamate-functionalized activated carbon,15 Cystoseira sedoide algae,16 and ethylenediamine-functionalized chelating resin.28 These findings provide valuable insights for optimizing adsorbents for selective heavy metal removal, particularly in wastewater treatment processes where cadmium contamination is a concern.
The comparative evaluation of various sorbents for Cu(II) and Cd(II) removal from aqueous solutions provides valuable insights into the performance of the newly synthesized GA-MMPs and GA-NiMMPs. As shown in Table 6, the adsorption data exhibit clear trends in removal efficiencies, highlighting the enhanced performance of GA-NiMMPs relative to both GA-MMPs and several benchmark materials reported in the literature. For Cu(II) adsorption, GA-NiMMPs achieved a capacity of 27.7 mg g−1, surpassing conventional sorbents such as biochar derivatives (12.2–14.7 mg g−1),10 natural biomass-based materials like pata-de-vaca leaves (15.1 mg g−1),38 and several zeolite-based adsorbents.13,39 In the case of Cd(II), GA-NiMMPs reached 30.4 mg g−1, outperforming engineered materials including CuO-modified ceramic membranes (12.4 mg g−1)4 and Zn0.6Mn0.4Fe2O4/PUF nanoparticles (11.1 mg g−1).18
Sorbent type | Co, mg L−1 | Temp. | pH | Time, min | qe, mg g−1 | R | |
---|---|---|---|---|---|---|---|
Cu(II) | Cd(II) | ||||||
CuO-modified ceramic membrane | 6.3–44.5 | 25 | 4.0 | 120 | 13.04 | 12.4 | 4 |
Willow wood biochar (WWB) | 0.05–160 | 25 | 5.2 | 1440 | 12.2 | 35.2 | 10 |
Cattle manure biochar (CMB) | 14.7 | 31.3 | |||||
Chitosan-vermiculite composite | — | 25 | 8.0 | 360 | 116.2 | 147.6 | 11 |
Ag-MOF/CSC composite sponge | — | 25 | 5.0 | 60 | 94.7 | 193.3 | 12 |
Fly ash-based Linde F(K) zeolite | — | 25 | 6.0 | 480 | 18.5 | 21.6 | 13 |
Sulfidized nanoscale zero-valent iron | 50–400 | 25 | 6.0 | 2880 | 85.1 | 126.9 | 14 |
Cystoseira sedoide alga | 25–150 | 25 | 5.0 | 180 | 14.66 | 23.78 | 16 |
Magnetite–chitosan composite | 0–200 | 25 | 5.0 | 10 | 26.28 | 18.67 | 17 |
Zn0.6Mn0.4Fe2O4/PUF nanocomposite | 0.2–300 | 25 | 6.5 | 60 | 20.4 | 11.1 | 18 |
Magnetic iron-modified calcium silicate hydrate | — | 25 | 4.0 | 300 | 26.12 | 16.39 | 19 |
Pectin hydrogel/Fe3O4/bentonite | 100–300 | 25 | 4.7 | 45 | 36.8 | 35.5 | 36 |
Pata-de-vaca leaves | 1–400 | 25 | 5.0 | 240 | 15.1 | 12.7 | 38 |
NaP1 zeolite | 0–223.5 | 22 | 5.6 | 180 | 20.9 | 26.9 | 39 |
Nanosized titanate composites (NaxFeyMgiTiOz) | 300–450 | 20 | — | 420 | 13.8 | 8.8 | 45 |
Commercial active carbon (Purolite AC 20) | 50–200 | 25 | 5.0 | 360 | 8.47 | 12.6 | 46 |
Biochar | 24.9 | 33.9 | |||||
GA-NiMMPs | 20–200 | 25 | 6.0 | 240 | 27.7 | 30.4 | PW |
The superior performance of GA-NiMMPs can be attributed to several synergistic factors. The incorporation of nickel likely introduces additional active sites, enhancing metal ion binding capacity. Moreover, the combination of gum Arabic—a natural polymer rich in functional groups—with the magnetite–nickel matrix may facilitate multiple adsorption mechanisms, including electrostatic attraction, surface complexation, and ion exchange. The mesoporous architecture further contributes by offering a high surface area, thereby increasing the number of accessible binding sites. It is also noteworthy that both GA-based sorbents exhibited higher affinity for Cd(II) over Cu(II), a trend consistent with other reported adsorbents. This selectivity may be influenced by intrinsic ionic properties such as hydration energy, ionic radius, and electronegativity differences between the two metal ions.
The experimental conditions employed (pH 6.0, 25 °C, and a contact time of 240 minutes) were carefully selected to simulate practical application scenarios, offering a balance between adsorption efficiency and operational feasibility. Additionally, the initial metal concentration range (20–200 mg L−1) reflects typical contamination levels found in industrial effluents. In conclusion, the newly developed GA-NiMMPs exhibit promising potential for the remediation of heavy metal-contaminated water, particularly for the removal of Cd(II) and Cu(II). Their adsorption capacities compare favorably with many established sorbents, positioning them as effective, eco-friendly materials for future water treatment technologies.
![]() | ||
Fig. 7 Thermodynamic profile for the capture process (initial concentration of 60 ppm; pH 5.9; ligand dose of 4.0 g L−1; shaking time is 240 min). |
ΔG (kJ mol−1) | ΔH | ΔS | ||||
---|---|---|---|---|---|---|
20 °C | 30 °C | 40 °C | 50 °C | (kJ mol−1) | (J mol−1 K−1) | |
Cu(II) | −17.8 | −18.3 | −19.7 | −21.4 | 25.0 | 144.0 |
Cd(II) | −19.3 | −20.0 | −21.6 | −23.6 | 31.8 | 171.9 |
The entropy change (ΔS° = 144.03 J mol−1 K−1 for Cu(II), 171.94 J mol−1 K−1 for Cd(II)) suggests increased randomness at the solid–solution interface, attributed to water molecule displacement and enhanced molecular freedom upon metal binding.14–16 The higher ΔS for Cd(II) compared to Cu(II) suggests a more significant structural reorganization upon adsorption, which is consistent with its higher affinity for GA-NiMMPs. Research on the adsorption of cadmium(II), and copper(II) ions onto various materials—including chitosan–vermiculite composite,11 Ag-MOF/chitosan composite sponges,12 exhausted copper slag-supported sulfidized nanoscale zerovalent iron,14 dithiocarbamate-functionalized activated carbon,15 Cystoseira sedoide algae,16 and ethylenediamine-functionalized chelating resin28 has shown the process to be spontaneous, thermodynamically favorable, and endothermic.
Overall, the thermodynamic evaluation highlights the fundamental energetic feasibility and thermal responsiveness of the adsorption process. The negative ΔG° values across the tested temperature range confirm that the adsorption of Cd(II) and Cu(II) onto GA-NiMNPs occurs spontaneously, a desirable attribute for practical wastewater treatment systems. The positive enthalpy values indicate that the process is endothermic, gaining in efficiency with increasing temperature, which is particularly beneficial for real-world applications involving warm industrial effluents. Meanwhile, the positive entropy changes reflect increased molecular disorder at the interface, attributed to the release of water molecules and enhanced freedom of adsorbed ions upon complexation. These thermodynamic parameters reinforce the findings of kinetic and isotherm studies, suggesting a mixed-mode adsorption involving both physical and chemical interactions. The compatibility of GA-NiMNPs with varying thermal conditions, along with their high binding affinity and adaptable surface chemistry, underscores their robustness and versatility as a sustainable adsorbent in dynamic environmental systems.
The removal of Cu(II) and Cd(II) from aqueous solutions proceeds through a hierarchical sequence of mechanistic pathways, beginning with surface interactions and progressing to specific chemical binding processes. Initially, electrostatic attraction serves as the primary driving force, drawing positively charged metal ions from bulk solution toward negatively charged adsorbent surfaces.15,18 This physical adsorption phase is particularly influenced by solution pH, which determines both the surface charge of the adsorbent (point of zero charge) and the speciation of metal ions.10 Following surface contact, multiple chemical binding mechanisms operate simultaneously with varying contributions. Coordination chemistry dominates through oxygen-containing functional groups—carboxyl (–COOH), hydroxyl (–OH), and phenolic (R-OH)—which undergo deprotonation to form stable metal complexes such as –(COO)2M, –COOM+, –(O)2M, and –OM+.46 Importantly, lattice Ni–O and Fe–O sites within the GA–NiMMPs framework contribute significantly to surface complexation, providing additional coordination centers for Cd(II) and Cu(II). Nitrogen-containing groups (–NH2) demonstrate effective coordination with both metals, while interestingly, hydroxyl groups show preferential binding to Cd(II) over Cu(II).28 Ion exchange represents another significant mechanism, particularly in silicate-based adsorbents where structural cations (often Ca2+) are displaced by incoming Cu(II) and Cd(II) ions.19 In more advanced materials, surface complexation and chelation create highly stable metal–ligand structures,15 while π–π interactions with aromatic structures in carbonaceous materials offer additional binding sites.12 Material-specific mechanisms include precipitation processes, such as the formation of metal hydroxides catalyzed by silanol groups in silicates,19 and redox reactions, exemplified by the reduction of Cu(II) to Cu0 on sulfurized nanoscale zero-valent iron surfaces.14 The synergistic integration of these mechanisms—electrostatic attraction, coordination bonding, ion exchange, precipitation, and redox transformations—varies with adsorbent composition and target metal, explaining the observed differences in removal efficiencies. This mechanistic understanding highlights the importance of rational adsorbent design that strategically incorporates functional groups with optimal affinity and selectivity for Cu(II) and Cd(II),4 enabling the development of more effective water purification technologies.
The adsorption of Cd(II) and Cu(II) ions onto GA-NiMMPs is a multifaceted process driven by the material's unique structural and functional properties, as revealed by comprehensive characterization and adsorption studies. The mesoporous structure and high surface area (29.32 m2 g−1) of GA-NiMMPs, confirmed by BET and SEM analyses (Table 2 and Fig. 2-I respectively), provide an extensive network of accessible active sites, facilitating efficient metal ion binding. The presence of functional groups such as hydroxyl (–OH), and carboxyl (–COO−), identified through FTIR analysis (Fig. 1-II), plays a pivotal role in enhancing the material's adsorption capacity through multiple mechanisms, including coordination bonding, electrostatic attraction, hydrogen bonding, and surface complexation. These functional groups act as electron donors, forming strong coordination bonds with Cd(II) and Cu(II) ions, which act as Lewis acids. The carboxyl groups, in particular, deprotonate at moderate pH values (pH ∼6.0) to form carboxylate anions (–COO−), enabling chelation and electrostatic interactions with the metal ions. The negative zeta potential of −21.32 mV (Table 1) further supports the electrostatic attraction mechanism, as the negatively charged adsorbent surface attracts positively charged Cd2+ and Cu2+ ions, enhancing adsorption efficiency.
Kinetic studies reveal that the adsorption process is dominated by chemisorption, as evidenced by the excellent fit of the pseudo-second-order model, which indicates that chemical interactions are the rate-limiting step. The initial rapid adsorption phase is attributed to the high density of active sites on the adsorbent surface, while the slower phase involves intraparticle diffusion, as confirmed by the Weber–Morris model. The mesoporous structure of GA-NiMMPs, with an average pore size of 688.3 nm (Table 1), facilitates efficient diffusion of metal ions to the interior active sites, ensuring rapid and effective adsorption. The Elovich model further supports the heterogeneous nature of the adsorbent surface, highlighting the role of multiple binding sites in the adsorption process.
Isotherm studies demonstrate that the adsorption of Cd(II) and Cu(II) follows a monolayer coverage mechanism, as indicated by the excellent fit of the Langmuir model. The maximum adsorption capacities (qm) of 30.4 mg g−1 for Cd(II) and 27.7 mg g−1 for Cu(II) (Table 5) reflect the material's high affinity for both metal ions. The Freundlich and Temkin models further underscore the heterogeneous nature of the adsorbent surface and the role of adsorbate–adsorbent interactions in enhancing adsorption performance. The Sips model, which combines features of the Langmuir and Freundlich models, confirmed that the adsorption process is predominantly monolayer in nature, consistent with the homogeneous distribution of functional groups on the adsorbent surface. Thermodynamic investigations reveal that the adsorption process is spontaneous, endothermic, and entropy-driven, as evidenced by the negative Gibbs free energy change (ΔG), positive enthalpy change (ΔH), and positive entropy change (ΔS) (Table 7). The occurrence of physisorption during the elimination of cadmium(II) and copper(II) is also confirmed by thermodynamic and Temkin isotherm analyses, which indicate ΔH° values of 25.0 and 31.8 kJ mol−1 (Table 7) and heats of adsorption of 414.3 and 428.9 (Table 5) for copper and cadmium ions respectively. These findings suggest that physisorption, via weak van der Waals interactions, plays a role in the removal of metal cations.14,15
Structural characterization by EDX (Fig. 1-I), and XRD (Fig. 1-III) further supports the adsorption mechanism, revealing the successful incorporation of Ni within the magnetite framework, which enhances the density of active sites and contributes to improved metal binding. The morphological features observed in SEM images indicate a rough and porous surface that maximizes contact with metal ions, optimizing the adsorption process. The thermodynamic findings confirm that adsorption is spontaneous, with increasingly negative Gibbs free energy (ΔG) values at higher temperatures, further validating the favorability of the process. The superior adsorption of Cd(II) compared to Cu(II) is attributed to its more effective interaction with GA-NiMMPs due to its softer Lewis acid nature, allowing for stronger chelation with the functional groups present on the sorbent.
Overall, the adsorption of Cd(II) and Cu(II) onto GA-NiMMPs is a complex but highly efficient process, integrating multiple interaction mechanisms-electrostatic attraction, coordination/complexation (via GA and Ni/Fe–O sites), ion exchange, hydrogen bonding, and van der Waals interactions-that are strongly supported by kinetic, isotherm, thermodynamic, and structural characterization insights. This comprehensive understanding of the adsorption behavior highlights the potential of GA-NiMMPs as a highly effective and selective material for heavy metal remediation in environmental applications. The main interaction pathways are summarized in Fig. 8, illustrating electrostatic attraction, coordination/complexation, ion exchange, hydrogen bonding, and physical adsorption (van der Waals), which act synergistically to capture Cd(II) and Cu(II) ions.
![]() | ||
Fig. 8 Schematic representation of the main proposed adsorption mechanisms for Cd(II) and Cu(II) onto GA-NiMMPs. |
To evaluate the real-world applicability of GA-NiMNPs beyond controlled laboratory environments, the material was tested on a chemically complex waste raffinate solution containing a mixture of heavy metals, and alkali metals. The sample was sourced from the Central Analytical Laboratory of the Nuclear Materials Authority (NMA), Cairo, Egypt, where multi-element chemical analyses and metal dissolution procedures generate composite aqueous effluents. The tested solution included toxic and trace-level metals such as Cd(II), Cu(II), As(III/V), Co(II), Ni(II), Cr(III), and Fe(III), as well as background electrolytes such as Na+ and K+, and trace amounts of organic dyes (e.g., methylene blue) used in analytical quality control. Initial concentrations of the target contaminants ranged from 6.0 to 65.8 mg L−1 (see Fig. S3).
Adsorption experiments were conducted under previously optimized conditions: GA-NiMNPs dosage of 4.0 g L−1, solution pH ∼6.0, contact time of 240 minutes, and ambient temperature. Following treatment, the removal efficiencies (E%) of the major contaminants were assessed and visually summarized in Fig. S3. GA-NiMNPs exhibited the highest removal for Cd(II) and Cu(II), achieving efficiencies of 40.1% and 25.3%, respectively—demonstrating strong affinity toward these priority pollutants and consistent with batch and isotherm findings. Other metal ions, including Co(II), Ni(II), Cr(III), and Fe(III), showed moderate removal (15–23%), which can be attributed to electrostatic interactions and coordination with carboxyl and hydroxyl groups on the gum Arabic-functionalized surface. Arsenic (present as As(III/V)) exhibited limited uptake (∼10%), likely due to the weak affinity of its oxyanionic forms under near-neutral pH. Common alkali metals Na+ and K+ were minimally removed (<5%) due to their low charge density and limited interaction potential.
Although adsorption performance in this multicomponent matrix was lower than in synthetic single-ion systems, GA-NiMNPs maintained structural integrity and functional activity under highly competitive conditions. The moderate reduction in performance is attributed to matrix interference effects. Specifically, the presence of organic dyes such as methylene blue may interfere by forming weak metal-dye complexes that hinder the availability of free Cu2+ and Cd2+ ions. Additionally, the presence of competing cations (e.g., Cr3+, Fe3+, Ni2+) and the high ionic strength of the solution may cause electrostatic shielding and partial occupation of available adsorption sites, thereby reducing overall uptake. Nevertheless, the sorbent preserved a consistent preferential binding for Cd(II) over Cu(II), aligning with earlier observations and affinity trends based on ion size, hydration energy, and softness.
These results underscore two key advantages of GA-NiMNPs: (1) their chemical robustness in diverse and ion-rich matrices, and (2) their maintained selectivity in the presence of multiple interfering species. From a practical standpoint, the ability to capture cadmium and copper from real, multi-element wastewaters highlights the material's value for sustainable water treatment applications. This case study supports the broader applicability of GA-NiMNPs in environmental remediation settings and affirms their role in circular economy strategies for biomass-derived sorbents.
The SI file includes adsorption kinetic and isotherm model equations with corresponding fitting parameters, detailed thermodynamic calculations, and supporting figures and tables (Fig. S1–S3 and Tables S1 and S2) that complement and extend the results presented in the main article. See DOI: https://doi.org/10.1039/d5ra04152j.
This journal is © The Royal Society of Chemistry 2025 |