Previous Article in Journal
Bituminous Coal-Derived Carbon Anode: Molten Salt-Assisted Synthesis and Enhanced Performance in Sodium-Ion Battery
Previous Article in Special Issue
NaOH-Modified Activated Carbon Materials for Hydrogen Sulfide Removal
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanistic Evaluation of Pb(II) Adsorption on Magnetic Activated Carbon/Fe3O4 Composites: Influence of Hydrothermal and Ultrasonic Synthesis Routes

by
Gaukhar Smagulova
1,2,
Aigerim Imash
1,2,3,*,
Akniyet Baltabay
1,2,
Aruzhan Keneshbekova
4,
Alisher Abdisattar
1,2,
Ramazan Kazhdanbekov
3,
Aidos Lesbayev
1 and
Zulkhair Mansurov
2,3
1
Department of General Physics, Intistute of Energy and Mechanical Engineering Named After A. Burkitbayev, Satbayev University, Almaty 050013, Kazakhstan
2
Institute of Combustion Problems, Almaty 050012, Kazakhstan
3
Faculty of Chemistry and Chemical Technology, Al Farabi Kazakh National University, 71 al-Farabi Ave., Almaty 050040, Kazakhstan
4
International Chinese-Belorussian Scientiffc Laboratory on Vacuum Plasma Technology, Nanjing University of Science and Technology, 200 Xiaolingwei Str., Nanjing 210094, China
*
Author to whom correspondence should be addressed.
Submission received: 26 July 2025 / Revised: 19 October 2025 / Accepted: 1 November 2025 / Published: 4 November 2025
(This article belongs to the Special Issue Carbons for Health and Environmental Protection (2nd Edition))

Abstract

This study presents a comparative analysis of two synthesis approaches for fabricating magnetic sorbents based on activated carbon (AC) incorporated with magnetite (Fe3O4) nanoparticles: hydrothermal synthesis and ultrasonic treatment. The results demonstrate that ultrasonic-assisted synthesis yields a magnetically responsive composite, us-AC/Fe3O4, exhibiting a Pb2+ removal efficiency of 92.84%, which is comparable to that of pristine activated carbon (99.0%). A key advantage of the synthesized composite lies in its facile recovery via magnetic separation following adsorption, rendering it a promising candidate for the remediation of heavy metal-contaminated water. Kinetic modeling suggests a dual adsorption mechanism: initial stages are governed by physisorption, while chemisorption dominates in the later phases. Adsorption isotherm modeling demonstrated that the Langmuir model provided the best description of Pb2+ adsorption on AC and us-AC/Fe3O4, with the highest sorption capacities observed for pristine activated carbon, followed by the ultrasonically modified composite, and comparatively lower values for the hydrothermally treated material. These findings underscore the potential of ultrasonic processing as an effective route for developing magnetically separable sorbents with high performance in aqueous heavy metal removal.

1. Introduction

Contamination of water resources with heavy metals remains one of the most pressing environmental issues, largely driven by intensive anthropogenic activities such as mining, battery manufacturing, electronics, metallurgy, and the chemical industry [1,2,3]. Heavy metal ions such as Pb2+, Cd2+, As3+/As5+, and Hg2+ are characterized by high toxicity, resistance to biodegradation, and a strong tendency to bioaccumulate in living organisms. These properties make them extremely hazardous to both the environment and human health, leading to neurotoxic, carcinogenic, reproductive, and mutagenic effects [4,5,6].
Particular concern is raised by lead (Pb2+), which exhibits high neurotoxicity and accumulates in bone tissue. Its effects are especially dangerous for children, as it impairs the development and function of the central nervous system, potentially leading to irreversible cognitive and behavioral disorders [7]. In aquatic systems, lead may exist in various forms, including free ions (Pb2+), hydroxyl complexes, carbonate and chloride complexes, as well as poorly soluble precipitates and colloidal particles [8,9,10,11]. The dominant forms depend on pH, redox conditions, and the presence of other ions in the medium. This complex chemical nature complicates the removal of lead, necessitating the development of universal and highly efficient sorbent materials. Various methods are employed for the removal of heavy metal ions from water, including chemical precipitation, membrane processes, ion exchange, electrochemical techniques, and adsorption [12]. Among these, adsorption is considered one of the most promising due to its simplicity, availability of raw materials, low cost, high efficiency, and the potential for sorbent reuse [13,14].
Activated carbon (AC) plays a particularly important role among adsorbents due to its high specific surface area (up to 3000 m2/g), well-developed meso- and microporous structure, significant pore volume, and the presence of oxygen-containing functional groups (carboxyl, phenolic, hydroxyl), which facilitate metal ion complexation [15,16,17]. Additionally, the use of renewable biomass sources for AC production enhances its sustainability and cost-effectiveness [18]. However, despite its outstanding adsorption properties, AC has several limitations, including difficulties in separating it from water after adsorption, as well as limited selectivity and adsorption rate for target ions. One effective strategy to overcome these drawbacks is the modification of AC with metal oxide nanoparticles, particularly magnetite (Fe3O4) [19,20]. Magnetite is a spinel-structured iron oxide with a high saturation magnetization, high adsorption capacity, and surface functionality that can be tailored to enhance selectivity toward specific ions [21,22,23]. These properties make Fe3O4 nanoparticles widely used as modifiers that impart magnetic properties to adsorbents, enabling rapid and complete separation from water using an external magnetic field, thereby preventing secondary pollution [24].
A key factor determining the effectiveness of such composites is the interaction between Fe3O4 nanoparticles and the porous structure of AC. Despite the extensive research on Fe3O4/AC-based sorbents, the effect of the nanoparticle incorporation method on the morphology, distribution of active sites, and adsorption properties of the material remains insufficiently explored. In particular, comparative analysis of different modification methods—such as ultrasonic treatment and hydrothermal synthesis—is of significant relevance, as these approaches lead to varying degrees of nanoparticle penetration into meso- and macropores, as well as differences in surface defect density and functionalization. The integration of magnetic nanoparticles into the macro- and mesopores of the carbon matrix not only prevents their aggregation and loss of activity in aqueous environments, but also significantly increases the active surface area and the number of available adsorption sites [25]. The porous AC framework serves as a stabilizing platform, ensuring uniform nanoparticle distribution, protection against oxidation and aggregation, and enhanced hydrophilicity of the material. At the same time, Fe3O4 nanoparticles embedded in the adsorbent structure enhance electrostatic interactions with metal ions and promote faster mass transfer through magnetic gradients [26].
Recent studies have shown that the synergistic effect resulting from the combination of activated carbon with Fe3O4 nanoparticles significantly enhances the sorption performance of the composite. This includes high Pb2+ ion removal efficiency, improved sorption kinetics, regeneration stability, and the potential for repeated use without a notable decline in activity [27,28]. However, the key factor determining the functionality of such composites remains their microstructural organization—particularly the degree of nanoparticle incorporation into the porous carbon matrix, their distribution within the pores, defect formation, and changes in surface chemistry. The structural and chemical integration of Fe3O4 nanoparticles not only affects the accessibility of active sites but may also induce textural rearrangement, formation of oxygen-containing functional groups, and the emergence of additional sorption sites due to defects and heterogeneous functionalization, as summarized in Table 1.
The relevance of the present study lies in addressing a clear gap in understanding how different modification methods affect the structure and performance of activated carbon/Fe3O4 composites. Although these materials have been extensively studied, most reports focus mainly on adsorption capacity, while the relationship between synthesis method, morphology, nanoparticle distribution, and functional properties remains insufficiently explored [35,36,37,38,39,40]. A comparative analysis of hydrothermal and ultrasonic modification is therefore of particular importance, as these approaches exert distinct effects on pore formation, nanoparticle dispersion and anchoring within the carbon matrix, and the density and accessibility of active sites [41,42,43,44,45]. Such insights are essential for identifying optimal modification conditions to improve sorbent efficiency. Furthermore, the sorption of Pb2+ proceeds through a combined mechanism: initial stages dominated by physical adsorption and intraparticle diffusion, followed by chemisorption involving carbon functional groups and Fe3O4 surface sites, as supported by recent studies [46,47].
Thus, the aim of the present study is to conduct a comparative investigation of the sorption properties of magnetic nanocomposites based on activated carbon and Fe3O4 nanoparticles synthesized via hydrothermal and ultrasonic methods. Special attention is given to the impact of synthesis conditions on adsorption efficiency and the feasibility of magnetic separation of the sorbent from aqueous solutions. Comprehensive physicochemical characterization and kinetic modeling were carried out to identify the key mechanisms underlying the sorption behavior of these materials and to propose rational design strategies for the development of next-generation sorbents for heavy metal ion removal from water.

2. Materials and Methods

The following chemical reagents were used in this study: rice husk (Kyzylorda region, Kazakhstan), potassium hydroxide (KOH), FeSO4·7H2O, FeCl3·6H2O, 25% aqueous ammonia solution, and Pb(NO3)2—all of analytical grade and purchased from LLP “Labor-Pharma (Almaty, Kazakhstan)” and LLP “Labhimprom(Almaty, Kazakhstan)”. Argon gas (Ar, 99.999%) was obtained from LLP “Ikhsan Technogaz (Almaty, Kazakhstan).”

2.1. Synthesis of Activated Carbon

Activated carbon was synthesized from rice husk using a two-step process consisting of carbonization followed by thermochemical activation. Prior to carbonization, the raw biomass was pretreated by thorough washing with hot water, rinsing with distilled water, and subsequent drying in a laboratory oven at 110 °C for 12 h to remove impurities and reduce moisture content. Carbonization was carried out at 550 °C for 100 min in a vertical tubular furnace under a constant nitrogen flow of 150 SCCM to ensure an inert atmosphere. The resulting biochar was then chemically activated using potassium hydroxide in a 1:4 mass ratio (carbon:KOH). The activation process was conducted in a stainless-steel reactor (AISI 321 grade), where the mixture was heated to 850 °C at a ramp rate of 7 °C/min under a continuous argon flow of 150 SCCM. The target temperature was maintained for 120 min to complete the activation. Passive cooling to ambient temperature was employed after both the carbonization and activation steps. A detailed description of the synthesis process can be found in [48].

2.2. Synthesis of Fe3O4 Nanoparticles

Magnetite nanoparticles were synthesized by a chemical co-precipitation method [49]. An aqueous solution of ferric chloride (FeCl3) with a concentration of 0.32 mol/L and a ferrous sulfate (FeSO4) solution with a concentration of 0.20 mol/L were prepared separately. Equal volumes of the two solutions (325 mL each) were transferred into a heat-resistant flask, and 200 mL of 25% aqueous ammonia solution was added dropwise at a rate of one drop per second under continuous stirring using a magnetic stirrer. The reaction temperature was maintained at 50 °C throughout the process. A black precipitate of Fe3O4 formed, which was separated by filtration and washed repeatedly with distilled water by decantation until a neutral pH was reached. The resulting precipitate was dried in an oven at 60 °C for 24 h to remove residual moisture.

2.3. Synthesis of AC/Fe3O4 Composites

2.3.1. Ultrasonication-Assisted Synthesis of us-AC/Fe3O4 Composite

The composite AC/Fe3O4 was synthesized via an ultrasonication-assisted method. To prepare the composite, activated carbon and the pre-synthesized Fe3O4 nanoparticles were mixed at a mass ratio of 2:1 and dispersed in 20 mL of distilled water. The resulting suspension was subjected to ultrasonication for 1 h in an ice bath using a Bandelin ultrasonic bath (Germany) operating at a frequency of 35 kHz, with a high-frequency power of 80 W and a peak ultrasonic power of 320 W. After sonication, the product was filtered, washed with distilled water, and dried at 60 °C. The resulting composite was designated as us-AC/Fe3O4, where “us” refers to ultrasonication.

2.3.2. Hydrothermal Synthesis of h-AC/Fe3O4 Composite

Magnetite nanoparticles supported on activated carbon were synthesized via a co-precipitation method followed by hydrothermal treatment. Initially, 100 mL of a 0.052 M FeSO4 solution and 150 mL of a 0.069 M FeCl3 solution were prepared and mixed under vigorous stirring at 70 °C. Subsequently, 6.4 mL of 25% aqueous ammonia was added dropwise until the pH of the mixture reached approximately 11. The appearance of a dark brown color indicated the formation of magnetite, after which 2.4 g of activated carbon was introduced into the reaction mixture. These conditions yielded a composite with an activated carbon to magnetite ratio of 2:1. The suspension was continuously stirred at 70 °C for 2 h, then allowed to stand at room temperature for 24 h.
The resulting composite was magnetically separated, thoroughly washed several times with distilled water and ethanol, and dried in an oven at 90 °C for 12 h. The intermediate product, obtained in a yield of 3.4 ± 0.18 g, was then subjected to hydrothermal treatment. Specifically, 0.230 ± 0.002 g of the composite was dispersed in 30 mL of deionized water and ultrasonicated for 30 min to ensure uniform dispersion. The suspension was transferred into a Teflon-lined stainless-steel autoclave and hydrothermally treated at 150 °C for 24 h. After cooling to room temperature, the product was magnetically separated, thoroughly washed with distilled water and ethanol, and dried in an oven at 80 °C for 12 h.
The final material, obtained in a yield of 0.210 ± 0.005 g, was designated as h-AC/Fe3O4, where “h” denotes hydrothermal treatment. The chosen mass ratios ensured sufficient material for 3–4 repeated sorption experiments (50 mg of h-AC/Fe3O4 per sorption test).

2.4. Adsorption Performance Test

To evaluate the adsorption properties of the synthesized composites, the lead ion adsorption capacity was measured. In a typical test, 50 mg of activated carbon or composite (us-AC/Fe3O4 or h-AC/Fe3O4) was dispersed in 50 mL of an aqueous Pb(NO3)2 solution with a concentration of 50 mg/L (based on Pb2+ ions) and stirred on a laboratory shaker at 250 rpm at room temperature (25 °C). An initial Pb(II) concentration of 50 mg/L was selected for the adsorption experiments based on two key considerations: practical relevance and scientific comparability. Pb(II) levels in industrial effluents vary widely depending on the source; for example, wastewater from the wood-processing industry has been reported to contain Pb(II) concentrations in the range of 51.8–414.4 mg/L [50]. Moreover, a concentration of 50 mg/L is frequently employed in the literature as a benchmark for assessing the performance of novel adsorbents, thereby enabling meaningful comparisons with previously published studies [51,52]. To evaluate the adsorption kinetics and identify the equilibrium time, experiments were conducted at different contact times (2, 4, and 6 h). After adsorption, the suspension was filtered, and the residual concentration of Pb2+ ions was analyzed using a Shimadzu AA-6200 (Kyoto, Japan) atomic absorption spectrometer. The adsorption capacity (q, mg/g) was calculated using the following equation:
q e = ( C 0 C e ) · V m
where C0 and Ce are the initial and equilibrium concentrations of Pb2+ (mg/L), V is the volume of the solution (L), and m is the mass of the sorbent (g).
To investigate the kinetic models, sorption experiments were conducted at various contact times. For this purpose, 50 mg of activated carbon or composite (us-AC/Fe3O4 or h-AC/Fe3O4) was dispersed in 50 mL of an aqueous Pb(NO3)2 solution with a concentration of 50 mg/L (based on Pb2+ ions) and stirred on a laboratory shaker at 250 rpm at room temperature (25 °C). Aliquots were taken at 5, 10, 20, 40, 60, and 120 min, respectively, to determine the residual concentration of Pb2+.
For constructing adsorption isotherms, solutions with Pb2+ ion concentrations of 25, 50, 75, 100, 125, and 150 mg/L were prepared. To each flask containing 50 mL of solution, 50 mg of activated carbon or composites (us-AC/Fe3O4 or h-AC/Fe3O4) was added. The flask with the solution and sorbent was continuously agitated on an orbital shaker. The adsorption time for all sorbents was set at 4 h. After the adsorption process, the sorbent was separated by centrifugation for activated carbon and by magnetic separation for the us-AC/Fe3O4 and h-AC/Fe3O4 composites. Subsequently, an aliquot of the post-adsorption solution was taken, and the residual Pb2+ concentration was determined using an atomic absorption spectrometer (Shimadzu AA-6200, Japan).

2.5. Determination of the Point of Zero Charge

The point of zero charge (PZC) of the composites were determined using the pH-drift method [53]. A fixed amount of the sorbent (25 mg) was added into 25 mL of mixture in a series of glasses. The pH of the mixture was adjusted by using 0.1 N HCl and/or 0.1 N NaOH solution. The pH was measured with a pH meter. Potassium chlorate (KCl) with 0.1 N concentrations was made to be added to the mixture of adsorbent and water to obtain ionic strengths.
The initial pH values of the suspensions were adjusted to 3, 5, 6, 7, 8, 10, and 12 by adding small amounts of standard HCl or NaOH solutions of various concentrations to 25 mL glasses of deionized water. The suspensions were shaken on an orbital shaker for 1 h at 25 °C, and the initial pH values were recorded as pHᵢ (pHwater).
After equilibration, 1 mL of KCl solution (0.1 N) was added to each beaker to maintain constant ionic strength, yielding a final volume of 26 mL. The suspensions were further agitated for 30 min under the same conditions, and the final pH values (pHf) were measured. The difference ΔpH was calculated as:
ΔpH = pHf − pHi
The ΔpH values were plotted against the initial pH values (pHᵢ). The PZC was identified as the pH at which the curve intersects the line ΔpH = 0.

2.6. Characterization

The structural and morphological characteristics of the synthesized samples were comprehensively examined using a range of advanced analytical techniques. Field-emission scanning electron microscopy (SEM, Auriga Crossbeam 540, Carl Zeiss, Oberkochen, Germany), equipped with an energy-dispersive X-ray spectroscopy (EDS) system (Oxford Instruments, AZtec 6.0 software, UK), was employed to investigate the surface morphology and elemental composition of the materials. Additionally, a Helios CX5 system (Thermo Fisher Scientific, Waltham, MA, USA) was utilized for high-resolution morphological and structural analysis of the composites. All microscopic analyses were performed at Nazarbayev University (Astana, Kazakhstan). X-ray diffraction (XRD) measurements were conducted using a PANalytical X’Pert PRO MPD diffractometer (Netherlands) equipped with Cu Kα radiation (λ = 1.5406 Å), operating at 40 kV and 40 mA. Diffraction data were collected over a 2θ range of 5° to 90° to determine the crystalline structure and phase composition of the magnetite-containing samples. Raman spectroscopic analysis was performed using an NT-MDT NTegra Spectra system (Russia, Moscow) with laser excitation wavelengths of 473 nm and 633 nm. These measurements were conducted at the National Nanotechnology Laboratory of Open Type, al-Farabi Kazakh National University (Almaty, Kazakhstan). Nitrogen adsorption–desorption isotherms were measured using Static multi-purpose gas adsorption analyzer SSA-4000 (Henan Mingguan Scientific Instruments Co., Ltd., Zhengzhou, China), and the specific surface area and porosity characteristics were subsequently calculated using the Brunauer–Emmett–Teller (BET) method.

3. Results and Discussion

The structural and chemical transformation of activated carbon to impart magnetic properties—enabling magnetic separation while preserving its porous structure and active adsorption sites—is crucial for the development of novel, sustainable, and safe sorbents. This study focuses on evaluating the influence of different methods for incorporating magnetite nanoparticles into the carbon matrix of activated carbon on its surface structure, porosity, and surface chemistry. Particular attention is given to maintaining high adsorption performance, demonstrated through the removal of lead ions from aqueous solutions.
Activated carbon is a well-established material for water purification, effective against a wide range of contaminants including metal ions, organic pollutants, and others [54]. The adsorption performance of AC is influenced not only by the type of raw precursor used but also by the carbonization and post-treatment conditions, which may include various activation and functionalization methods [55]. These factors collectively affect the material’s structure and morphology, porosity, surface chemistry, defect density, and degree of graphitization—all of which directly impact its adsorption capacity. Figure 1 shows SEM images of activated carbon derived from rice husk, revealing its heterogeneous surface morphology
The material exhibits a rough, porous texture with micro- and mesopores, which is typical for carbonaceous adsorbents prepared via thermochemical activation. These pores enhance the specific surface area and provide numerous active sites for interaction with adsorbates. In addition to the porous structure, certain regions display layered arrangements, indicative of partial graphitization. The presence of these lamellar features may suggest the formation of graphitic carbon domains during high-temperature treatment [56]. Such structures can improve the material’s electrical conductivity and mechanical stability, and may contribute to π–π interactions with adsorbed metal ions or organics. EDS elemental analysis confirms that the sample is primarily composed of carbon (C, 78.88 ± 0.33 wt. %) and oxygen (O, 20.26 ± 0.54 wt. %), with trace amounts of silicon (Si, 0.35 ± 0.04 wt. %) and potassium (K, 0.51 ± 0.05 wt. %). Silicon likely originates from residual silica inherent in rice husk, while potassium is introduced during the activation process. These results suggest that the rice husk-derived activated carbon possesses a well-developed porous structure combined with partially graphitized carbon domains, making it a promising material for adsorption-based water purification [57].
The Raman spectrum of the rice husk-derived activated carbon exhibits two prominent peaks located at approximately 1346 cm−1 and 1581 cm−1, which correspond to the D–and G–bands, respectively (Figure 2). The G–band (~1581 cm−1) is attributed to the E2g vibration mode of sp2-hybridized carbon atoms in graphitic structures. It indicates the presence of ordered graphitic domains and is a signature of crystalline carbon networks. The D–band at ~1346 cm−1 (disordered-induced) associated with defects or disorders within the carbon lattice, such as edge planes, vacancies, and functional groups.
The intensity ratio ID/IG provides insight into the degree of disorder and graphitization. The intensity ratio ID/IG was calculated to be 0.95. A relatively high intensity of the D-band compared to the G-band suggests a significant amount of structural defects or amorphous carbon, which is typical for activated carbon materials derived from biomass. This value implies that while the material retains some graphitic character, a significant proportion of amorphous or defect-rich carbon is present [58,59]. Such structural disorder is beneficial for adsorption, as it enhances surface reactivity by providing more active sites for Pb(II) ion binding. Overall, the Raman spectrum confirms that the synthesized material contains both disordered carbon regions and partially graphitized structures, in agreement with the SEM observations of layered domains. This hybrid structure enhances both the surface reactivity and the mechanical stability of the adsorbent.
Magnetite nanoparticles obtained by the co-precipitation method were also characterized using SEM, XRD, and BET analyses. Analysis of the SEM images (Figure 3) revealed that the Fe3O4 nanoparticles possess a nearly spherical morphology with a relatively uniform size distribution in the range of 20 to 35 nm. Additionally, larger particles exceeding 50 nm were observed, which is attributed to agglomeration driven by the intrinsic magnetic interactions between particles. EDS elemental mapping confirmed the high purity of the synthesized material, consisting primarily of iron and oxygen. Minor signals of carbon, silicon, and gold detected in the spectrum are ascribed to the substrate and sample preparation process.
The X-ray diffraction (Figure 4) pattern of the magnetite nanoparticles exhibited six distinct diffraction peaks at 2θ values of 30.3°, 35.6°, 43.4°, 53.6°, 57.2°, and 62.9°, corresponding to the characteristic crystallographic planes of the magnetite phase: (220), (311), (400), (422), (511), and (440), respectively. A weak reflection at 2θ = 19.7° and a minor peak at 18.8° may be attributed to lattice disorder, which is commonly observed in nanoparticles, or to the presence of trace impurities. The average crystallite size (t) was estimated using the Scherrer equation (t = λ/βcos θ), where λ is the wavelength of the Cu Kα radiation, θ is the Bragg angle, and β is the full width at half maximum of the diffraction peak in radians. The calculated crystallite size was approximately 21.04 nm. In comparison, scanning electron microscopy analysis revealed an average particle size of approximately 31.41 nm, suggesting possible agglomeration or that individual particle might consist of multiple crystalline domains. Such discrepancies are common, as the XRD method determines the size of coherently diffracting domains, whereas SEM reflects the actual physical dimensions of the particles. According to the Brunauer–Emmett–Teller analysis, the specific surface area of the obtained magnetite nanoparticles was found to be 68.42 m2/g, indicating a high degree of dispersion and nanoscale morphology. These properties render them highly promising for applications in adsorption, catalysis, and magnetic separation.
The us-AC/Fe3O4 composite was synthesized by combining rice husk-derived activated carbon with magnetite nanoparticles prepared via the co-precipitation method, followed by ultrasonication treatment. The SEM images of the us-AC/Fe3O4 composite illustrate a textured and heterogeneous surface morphology (Figure 5). The activated carbon matrix retains its porous, irregular structure, which is favorable for adsorption. At higher magnifications (×50,000–×120,000), it can be observed that the magnetite nanoparticles tend to agglomerate [60]. The elemental composition includes carbon (C, 69.58 ± 0.17 wt%), oxygen (O, 25.04 ± 0.29 wt%), and iron (Fe, 4.94 ± 0.10 wt%), along with trace amounts of silicon (Si, 0.20 ± 0.02 wt%) and potassium (K, 0.24 ± 0.02 wt%). Silicon is present as a trace component in the activated carbon, originating from the rice husk precursor, whereas potassium may arise from residual activating agents. As a result of ultrasonication, magnetite nanoparticles are uniformly distributed across the surface of the activated carbon particles, thereby imparting the required magnetic properties to the material and enabling its subsequent magnetic separation after the sorption processes.
The XRD pattern of the us-AC/Fe3O4 composite (Figure 6) displays distinct diffraction peaks at 30.2°, 35.6°, 43.3°, 53.7°, 57.3°, and 62.9°, which can be indexed to the (220), (311), (400), (422), (511), and (440) planes of magnetite (Fe3O4) in accordance with JCPDS card no. 19-0629. These reflections confirm the characteristic cubic spinel structure of Fe3O4 [61,62], indicating that the crystalline framework of magnetite remains preserved after ultrasonic treatment. In addition, the broad diffraction band observed near 26.2° is attributed to the amorphous structure of activated carbon, whereas the sharp peaks correspond to well-crystallized Fe3O4 nanoparticles uniformly embedded within the carbon matrix [63,64].
The h-AC/Fe3O4 composite was synthesized by co-precipitating magnetite nanoparticles onto activated carbon, followed by hydrothermal treatment. The SEM images of the h-AC/Fe3O4 composite at various magnifications reveal a heterogeneous surface morphology typical of magnetically modified carbon materials (Figure 7).
Distributed across the activated carbon surface are numerous spherical and quasispherical nanoparticles of magnetite Fe3O4 formed during co-precipitation. At higher magnifications (×20,000–×120,000), these nanoparticles appear uniformly anchored to the carbon surface, forming tight clusters or aggregates.
The EDS mapping confirms the presence of key elements: carbon (C, 70.00 ± 0.18 wt%), oxygen (O, 25.01 ± 0.31 wt%), and iron (Fe, 4.88 ± 0.10 wt%), with minor traces of silicon (Si, 0.11 ± 0.02 wt%). The significant iron content confirms the successful incorporation of Fe3O4 into the composite. The detected silicon originates from residual silica inherently present in the rice husk precursor used for the production of activated carbon. The combined SEM and EDS results demonstrate that the hydrothermal method enables uniform deposition of Fe3O4 nanoparticles onto the carbon substrate, yielding a composite with structural stability, and magnetically responsive properties.
The XRD pattern of the h-AC/Fe3O4 composite (Figure 8) exhibits distinct reflections at 26.0°, 30.3°, 35.6°, 43.2°, 44.7°, 48.3°, 53.8°, 57.4°, and 63.0°. The broad peak at 26.0° is attributed to the (002) plane of graphitic carbon with sp2 hybridization [65]. The diffraction peaks at 30.3°, 35.6°, 43.2°, 53.8°, 57.4°, and 63.0° correspond to the characteristic reflections of magnetite (Fe3O4), confirming the preservation of its crystalline spinel structure. In addition, the weak peak observed at 44.7° may suggest a minor contribution from metallic Fe or Fe3C phases [66].
The nitrogen adsorption–desorption isotherms measured at 77 K (Figure 9a) correspond to type Ib with an H4 hysteresis loop, indicating the predominance of micropores [67,68], referred to in [69] as “silt-pores”, as well as a high specific surface area as determined by the Brunauer–Emmett–Teller (BET) method. The pronounced uptake at low p/p0 is attributed to rapid filling of molecular-scale micropores owing to strong adsorbent–adsorptive interactions [70]. The presence of the characteristic hysteresis loop confirms the existence of capillary pores and a complex porous morphology. At low relative pressures (p/p0 < 0.1), a steep adsorption uptake is observed, corresponding to micropore filling. At higher relative pressures (p/p0 > 0.8), a sharp rise in adsorption is evident, particularly for AC and more prominently for us-AC/Fe3O4, which is attributed to capillary condensation within pores. The specific surface areas calculated using the BET equation for AC, us-AC/Fe3O4, and h-AC/Fe3O4 were 1818.2, 1007.2, and 833.1 m2/g, respectively.
The unmodified activated carbon exhibited the highest nitrogen adsorption–desorption capacity, reflecting its well-developed microporous structure typical of plant-derived activated carbons. Following hydrothermal modification with Fe3O4 nanoparticles, a pronounced decrease in nitrogen uptake was observed, indicating a significant reduction in accessible surface area and/or pore volume (Figure 9a). This decrease can be attributed to the partial blockage of pores by Fe3O4 particles during hydrothermal treatment and precipitation, resulting in denser coverage of the carbon surface. In contrast, ultrasonic treatment facilitated a more homogeneous dispersion of Fe3O4 nanoparticles on the carbon framework, thereby reducing excessive pore blockage. Consequently, the surface area of us-AC/Fe3O4 remained higher than that of h-AC/Fe3O4, but lower than that of pristine AC.
The pore size distribution was obtained from the nitrogen desorption branch using the Horváth–Kawazoe model, which is suitable for the analysis of microporous structures under the assumption of slit-shaped pore geometry. This semi-empirical method is particularly useful for the comparative analysis of microporous materials [70]. The results revealed a narrow distribution predominantly in the range of 0.57–0.84 nm for all three samples (Figure 9b). Such pore-size distribution for the activated carbon derived from rice husk via KOH thermochemical activation is in good agreement with previously reported studies. In particular, in work [71], it was shown that for KOH-activated rice husk carbon the predominant fraction of micropores was also concentrated within the 0.5–0.9 nm range. In work [72], activated carbon was also obtained from oil-tea shell by KOH thermochemical activation, yielding a microporous structure with pore widths predominantly in the range of 0.6–0.8 nm.
At the same time, the pore volume is significantly reduced for the h-AC/Fe3O4 and us-AC/Fe3O4 composites compared with AC, which also indicates possible pore blocking, since the pore diameter remains in the range of 0.57–0.84 nm for all investigated sorbents. Pores of adsorbents fall within a suitable size range for the sorption of Pb2+ ions, whose effective ionic radius is ~120 pm (0.12 nm). This dimensional match allows the ions to penetrate the pores and promotes effective interaction with the active sorption sites. Considering these findings, the us-AC/Fe3O4 composite represents an intermediate case, combining a relatively high specific surface area with a uniform dispersion of nanoparticles and effective magnetic separation. This combination of properties makes it a promising candidate for adsorption applications, providing both enhanced performance and convenient recovery via magnetic separation.
The synthesized sorbents were evaluated for their efficiency in removing Pb2+ ions. To evaluate the contact time at which the maximum adsorption capacity is achieved, experiments were conducted at 2, 4, and 6 h (SI Table S1). It was found that for AC and the us-AC/Fe3O4 composite, the maximum adsorption capacity was reached at 4 h. In the case of the h-AC/Fe3O4 composite, the highest value was observed at 6 h. However, due to its comparatively lower adsorption capacity relative to the other materials, the data for this composite are also presented at 4 h in the main text. The results of lead adsorption capacity after 4 h of sorption at actual pH of initial solution equal to 5.7 ± 0.2 are presented in Figure 9. After the sorption processes, an increase in the solution pH was observed with decreasing residual concentrations of lead ions. Specifically, the pH value was 6.5 for activated carbon, 6.3 for the us-AC/Fe3O4 composite, and 5.9 for the h-AC/Fe3O4 composite. These results indicate that the efficiency of Pb2+ removal is directly reflected in the change in the solution’s acid–base environment: the higher the degree of cation removal, the closer the pH shifts toward neutral values. The points of zero charge (PZC) of AC and AC/Fe3O4 composites were determined using the pH-drift method. The ΔpH vs. pHᵢ plots clearly demonstrate the crossing of the curve with the ΔpH = 0 axis. It was determined that the point of zero charge (pHpzc) for the three investigated sorbents corresponded to the following pH values: 6.2 for AC, 6.8 for us-AC/Fe3O4, and 7.1 for h-AC/Fe3O4 (SI Figure S1). Determination of the PZC is critically important for understanding and optimizing the adsorption performance of AC/Fe3O4 in water treatment processes aimed at Pb2+ removal. The PZC defines the pH at which the adsorbent surface has no charge, at pH values below the PZC the surface is positively charged. At this pH value, competing adsorption of protons and positively charged lead ions uptake due to electrostatic repulsion can be observed. While at pH values moderately above the PZC the surface becomes negatively charged, thereby enhancing Pb2+ adsorption. However, increasing the pH of the solution by adding alkali to the solution leads to the precipitation of insoluble lead hydroxide, which distorts the true adsorption and increases the error [73]. Thus, the sorption characteristics of the sorbents were assessed at the actual pH value, without its correction. The results of lead adsorption capacity after 4 h of sorption at actual pH equal to 5.7 ± 0.2 are presented in Figure 10.
It was found that the adsorption efficiency and adsorption capacity of the studied materials after 4 h of contact time were as follows: AC—99.00%, 42.03 ± 1.82 mg/g, h-AC/Fe3O4—56.27%, 23.89 ± 1.19 mg/g, us-AC/Fe3O4—92.84%, 39.15 ± 1.64 mg/g. To evaluate the magnetic separation performance, both hydrothermally and ultrasonically synthesized AC/Fe3O4 composites were subjected to an external magnetic field. The complete separation of both magnetic composites upon application of an external magnetic field demonstrates the high efficiency of magnetic separation following the sorption process (Figure S2). The observed decrease in adsorption performance in the order AC > us-AC/Fe3O4 > h-AC/Fe3O4 is consistent with the specific surface area results obtained for the three composites, which showed a reduction from 1818.2 m2/g to 1007.2 and 833.1 m2/g, respectively. Thus, the decrease in active sorption surface area and partial pore blocking lead to reduced adsorption efficiency, which is most pronounced for the h-AC/Fe3O4 composite.
The adsorption kinetics of the investigated materials were analyzed by fitting the experimental data to the pseudo-first-order and pseudo-second-order models, which are expressed by Equations (3) and (4), respectively:
ln (qe − qt) = ln qe − k1 · t
t q = 1 k 2 · q e 2 + 1 q e · t
where qt (mg·g−1) is the amount of Pb(II) adsorbed at time t, qe (mg·g−1) is the adsorption capacity at equilibrium, k1 is the sorption rate constant for the pseudo-first order model, k2 is the sorption rate constant for the pseudo-second order model.
In addition, the sorption mechanisms were examined using the intraparticle diffusion model (Weber–Morris), Boyd’s film diffusion model, Bangham’s kinetic model, and the Elovich chemisorption model.
The results obtained from the adsorption kinetics (Figure 11) demonstrate that the adsorption rates vary among the three investigated sorbents. For pristine activated carbon, equilibrium concentration (Ce) is reached within the first 5 min, indicating rapid adsorption. In contrast, for the us-AC/Fe3O4 composite, the most significant decrease in lead concentration occurs within the initial 5 min, followed by a comparatively slower adsorption process, reaching equilibrium after approximately 120 min. This behavior may be attributed to the diffusion of Pb2+ ions into deeper pores of the sorbent. In the case of the h-AC/Fe3O4 composite, the sharpest decline in initial concentration is also observed within the first 5 min, with equilibrium being achieved after 40 min.
Extending the contact time beyond this point results in no significant change in the residual lead ion concentration. The results of the pseudo-first-order model calculations, including the rate constant k1 and the correlation coefficient R2, are presented in Table 2 for all three investigated sorbent composites. As shown in Figure 11a,c,e, a similar trend is observed in the pseudo-first-order plots for all samples. However, deviations from linearity become apparent after 10 min of adsorption for the h-AC/Fe3O4 and us-AC/Fe3O4 composites, and after 20 min for pristine AC. This suggests that the pseudo-first-order model is primarily applicable during the initial stages of the adsorption process [74]. It can be assumed that at the beginning, adsorption is governed by physical mechanisms such as van der Waals interactions or diffusion into readily accessible pores [75].
The adsorption profile shown in Figure 11 supports this behavior: for pristine AC, a rapid attainment of equilibrium concentration is observed within the first 5 min, with no significant change in residual concentration thereafter. In contrast, the adsorption curves for the composites exhibit two distinct stages: an initial rapid decline in lead ion concentration due to physisorption on the sorbent surface, followed by a slower adsorption phase leading to equilibrium. Accordingly, adsorption on AC is primarily governed by physisorption, whereas for the h-AC/Fe3O4 and us-AC/Fe3O4 composites, the process involves initial physisorption followed by chemisorption at later stages [76]. Physisorption may proceed via both surface diffusion and pore diffusion mechanisms. A comprehensive discussion of the processes represented by the pseudo-second-order kinetic model, along with its inherent limitations, is provided in [77].
The obtained results suggest that the incorporation of magnetite nanoparticles into the AC structure via hydrothermal treatment leads to the partial filling of micro- and mesopores with Fe3O4 nanoparticles, significantly reducing the adsorption capacity of the resulting composite [78]. In contrast, the ultrasonic-assisted deposition of magnetite nanoparticles promotes their attachment to the surface without compromising the porosity of the activated carbon. As a result, a composite is formed that is suitable for lead ion adsorption and exhibits magnetic properties [79], enabling efficient magnetic separation of the sorbent after the adsorption process.
The determination of mass transfer parameters is essential for clarifying the adsorption mechanism and for the rational design of sorption processes. As described in [76], the overall kinetics of mass transfer can be divided into three main stages. The first involves external diffusion, where adsorbate molecules migrate through the liquid film surrounding the adsorbent under the driving force of the concentration gradient between the bulk solution and the adsorbent surface. The second stage corresponds to intraparticle diffusion, during which the adsorbate penetrates into the pores and channels of the sorbent. The third stage is the interaction of the adsorbate with the active binding sites of the sorbent surface. Among these, the slowest step controls the overall sorption process and thus determines the rate-limiting stage. Identifying the rate-limiting step is essential for the effective design and optimization of sorption processes. Typically, adsorption is governed by two dominant mass transfer mechanisms: external film diffusion and intraparticle diffusion. To obtain deeper insights into the controlling mechanism, several kinetic models were applied in this study, namely the Weber–Morris intraparticle diffusion model, Boyd’s film diffusion model, Bangham’s kinetic model, and the Elovich chemisorption model.
The Boyd model was employed to illustrate the diffusion of adsorbate through a limiting liquid film. This model is employed to predict the actual rate-determining stage of the adsorption process. The Boyd model is employed to ascertain the governing mechanism of the adsorption process, whether it occurs due to intramolecular (or intradiffusion) diffusion or external diffusion (diffusion through the liquid-solid boundary) [80,81]. The Boyd model is expressed by Equation (5).
Bt = −0.4977 − ln(1 − F)
where F = qe/qt is the proportion of equilibrium sorption at time t, Bt is the Boyd function, qt is the adsorption at time t (mg/g), and qe is the equilibrium adsorption capacity (mg/g).
The Elovich model, also referred to as the Elovich equation, is an empirical kinetic model that has gained significant traction in the field of adsorption on heterogeneous surfaces, particularly in scenarios where the sorbent surface exhibits an uneven distribution of active centers [80,82]. The nondifferential form of the equation is expressed by Equation (6):
q t =   1 B log e α β +   1 β log e t
In this equation, qt denotes the amount of adsorbed substance at time t (mg/g), α is the initial rate of adsorption (mg/g·min), β is the desorption constant associated with energy activation (g/mg), and t is time (min).
Bangham’s kinetic model is employed to elucidate the mechanisms of adsorption, particularly in circumstances where pore diffusion and micropore accessibility are taken into account [83,84]. This model posits that intrapore diffusion functions as a limiting stage in the adsorption process. The model was utilized to ascertain whether diffusion in pores is the sole mechanism governing velocity.
The logarithmic form of the model looks like this (7):
log log C 0 C 0 q t   · m = log k 2.303   · V + α · log t
where C0 is the initial concentration of ions in the solution (mg/L), qt is the amount of adsorbed substance at a time (mg/g), m is the mass of the sorbent (g), V is the volume of the solution (mL), k is the Bangham constant (related to kinetics and porosity), α is the indicator characterizing the mechanism of pore diffusion, t is the time (min).
The Weber–Morris Model (intraparticle diffusion model) is used to analyze the mechanism of adsorption and, in particular, to find out whether intraparticle diffusion (pore diffusion) is involved in limiting the rate of the sorption process [85,86]. The linear form of the equation has the form (8):
q t =   k i d   · t 1 2 + C
where qt is the amount of adsorbed substance at time t (mg/g); kid is the intraparticle diffusion constant (mg·g−1·min−0․5), t1/2 is the square root of time, and C is the parameter reflecting the boundary layer thickness.
The calculation results for each of the models for the studied adsorbents are presented in Table 3.
The fitting of the experimental data to the Weber–Morris intraparticle diffusion model is presented in Figure 12, Figure 13 and Figure 14 for the investigated sorbents. As shown by the Weber–Morris calculations, the obtained results were multilinear (Figure 12a), suggesting that the adsorption process is controlled not only by intraparticle diffusion but also by two or more stages. The data are represented by two linear phases: the initial phase corresponds to the influence of the boundary layer with external mass transfer, during which lead ions are rapidly adsorbed. After 60 min, the adsorption rate decreased, leading to the second phase, which lasted up to 120 min. According to [81], this phase is associated with the diffusion of molecules to the internal adsorption sites of the sorbent. A comparison of the two diffusion constants of the Weber–Morris model (kid) shows that kid1 > kid2, as presented in Table 2. The diffusion of molecules inside the sorbent is thus the rate-determining step of the adsorption process, as confirmed by the lower kid2 value. However, intraparticle diffusion may not be the sole limiting step of the adsorption process, and other interaction mechanisms may act simultaneously.
The parameters of the initial adsorption rate, α and β, calculated using the Elovich model for the sorption of Pb ions on activated carbon, have excessively high values. This indicates that the description of the sorption process using this model is not applicable or incorrect (Figure S3). The parameters of the initial adsorption rate calculated using the Elovich model for the sorption of Pb2+ ions on h-AC/Fe3O4 are α = 58.01 mg/g·min and β = 0.1883, and for us-AC/Fe3O4, α = 52.238 mg/g·min and β = 0.3355, respectively. These results indicate a high initial rate of adsorption and a moderate decrease in the rate with increasing sorption on the surface [82].
In the next step, the applicability of the Bangham model was tested. The ensuing results are delineated in Table 3 and Figure S4. For all three sorbents, the α values range from 0.07 to 0.14. The fact that α < 1 indicates that the process is controlled by both intrapore diffusion and external mass transfer, i.e., a mixed transport mechanism [84]. The highest α value is observed for h-AC/Fe3O4 (0.1407), possibly due to the high resistance of the pore structure after hydrothermal treatment, due to complete or partial pore blockage due to blockage by Fe3O4 nanoparticles. All values of R2 > 0.91, which confirm the model’s suitability for describing diffusive behavior. The optimal data correspondence (R2 = 0.937–0.9389) is observed between AC and us-AC/Fe3O4, indicating the most suitable description of the kinetics of Pb2+ penetration into the porous structure.
The experimental data was also evaluated using the Boyd model [87] (Figure S5). Linear approximation results yielded coefficients of determination (R2) for the sorption process using each sorbent. R2 values range from 0.7283 to 0.8893, indicating a relatively weak agreement with the experimental data. Furthermore, all graphs constructed using the Boyd model demonstrate that the line does not pass through the origin, indicating that intraparticle diffusion is not the sole limiting factor.
The adsorption of Pb2+ ions by sorbents was modeled using Freundlich and Langmuir isotherms. The adsorption isotherm is defined as the relationship between the amount of substance adsorbed on the surface of the sorbent and its equilibrium concentration in solution. The study of adsorption isotherms is of key importance for understanding the mechanisms of adsorbate–adsorbent interaction and improving the efficiency of sorption materials. To interpret the experimental data in this study, two of the most common models were used: Langmuir and Freundlich. The quality of compliance was assessed using the correlation coefficient. The isotherms obtained in this study facilitate the visualization of the relationship between the amount of substance retained by the adsorbent and its equilibrium concentration in solution.
The Langmuir model was originally proposed to describe the adsorption of gas molecules on homogeneous solid surfaces (crystalline materials) with one type of adsorption centers [88,89]. The model under consideration is of an empirical nature and is predicated on kinetic principles. In a state of equilibrium, the rates of adsorption and desorption on the surface are balanced, and there is no total accumulation. The following assumptions underpin this study: (a) the process proceeds in the form of monolayer adsorption, (b) all active centers of the surface are energetically homogeneous, (c) the adsorption energy remains constant, and (d) there are no lateral interactions between the adsorbed molecules [90]. The Langmuir isotherm can be described by Equation (9):
q e =   1   K L q m a x C e +   1 q m a x
where qe is the amount of adsorbate, adsorbent substance (mg/g), Ce is the equilibrium concentration of adsorbate (mg/L), KL is the coefficient of adsorption energy (L/mg), qmax is the maximum adsorption capacity (mg/g) [91].
In contrast to the Langmuir isotherm, the empirical Freundlich model is capable of being utilized for multilayer adsorption on heterogeneous sites. The Freundlich model has been demonstrated to be capable of describing experimental data from adsorption isotherms, irrespective of whether adsorption occurs on homogeneous or heterogeneous centers. Additionally, it has been shown to be independent of the formation of a monolayer. Furthermore, the Freundlich isotherm is frequently employed to characterize multilayer adsorption and adsorption on heterogeneous surfaces, where there is an interaction of neighboring adsorbed molecules [92]. The mathematical model can be represented by Equation (10):
log e q e =   1 n log e C e +   log e K F
where qe is the amount of adsorbate, adsorbent substance (mg/g), Ce is the equilibrium concentration of adsorbate (mg/L), KF is the Freundlich constant (mg/g), heterogeneity factor (dimensionless) [93]. The Pb2+ adsorption isotherms (Figure S6) revealed distinct uptake behaviors for the three sorbents, which can be attributed to differences in their chemical composition as well as in their textural and morphological characteristics.
The parameters of the isotherms for the Freundlich and Langmuir models, calculated for the adsorption of Pb2+ ions by AC and AC/Fe3O4 composites, are presented in Table 4.
According to the obtained results, the following assumptions can be made from the experiments. The maximum sorption capacity (qmax) calculated for AC is 63.69 mg/g. For the us-AC/Fe3O4 and h-AC/Fe3O4 composites, the calculated capacities are 41.32 and 24.16 mg/g, respectively. These values are in good agreement with the experimental data. The result indicates a strong affinity of AC for Pb2+ ions, a typical occurrence for carbons with oxygen-containing groups. For us-AC/Fe3O4, KL is 0.14 L/mg, indicating a moderate affinity for Pb2+ and 0.05 L/mg for h-AC/Fe3O4, suggesting a weaker affinity for Pb2+. The coefficient of determination for all three sorbents ranges from 0.8373 to 0.9997, and it has been demonstrated to be a comparatively superior indicator of compliance with the Langmuir model for describing the processes of adsorption of Pb2+ ions of AC and us-AC/Fe3O4. The constants of the Freundlich isotherms, KF and n, are determined from the slope of the graph of log qe versus log Ce (Figure 15).
In this study, the values of n > 1 indicate chemosorption and reflect the high affinity between adsorbate and adsorbent [94]. The Freundlich constant, KF, which is associated with the adsorption capacity, exhibits an increase in the range KF(h-AC/Fe3O4) < KF(us-AC/Fe3O4) < KF(AC) (Table 4). This observation indicates the endothermic nature of the adsorption process. The Freundlich model is a more suitable model for describing the adsorption process on the h-AC/Fe3O4 composite, as indicated by its higher coefficient of determination.
Comparative analysis of the adsorption capacity (Table 5) shows that the obtained Fe3O4/AC composite demonstrates values comparable to a number of well-known nanomaterials.
Despite exhibiting slightly lower rates in comparison to individual highly effective sorbents, the developed samples demonstrate a favorable performance when utilizing affordable raw materials and environmentally friendly synthesis methods. This observation serves to substantiate their practical significance and to underscore their capacity for further optimization.

4. Conclusions

In this study, two approaches to synthesizing magnetic sorbents based on activated carbon from rice husks with the inclusion of magnetite nanoparticles by hydrothermal and ultrasonic treatment were investigated. The experimental results demonstrated that the us-AC/Fe3O4 composite exhibited an adsorption capacity of 39.15 mg/g and a removal efficiency of 92.84% when adsorbing lead (Pb2+) ions from an aqueous solution with an initial concentration of 50 mg/L over a period of 4 h. These values are comparable with the adsorption capacity of the initial activated carbon (42.03 mg/g and 99.0%). A notable benefit of us-AC/Fe3O4 is its capacity for expeditious extraction following sorption via magnetic separation, a process that significantly streamlines the water purification procedure.
The analysis of kinetic models demonstrated that the adsorption of Pb2+ occurs in two stages: rapid physical adsorption in the initial stages and subsequent chemosorption. A comparison of experimental data with the Weber–Morris, Elovich, Bangham, and Boyd models indicates a high initial sorption rate and a gradual deceleration due to surface saturation. The investigation revealed that the process is governed by two primary mechanisms: external mass transfer and intraparticle diffusion. The most suitable agreement with the experiment for AC and us-AC/Fe3O4 was demonstrated by the Weber–Morris and Bangham models. Conversely, for h-AC/Fe3O4, the Weber–Morris and Elovich models provided a more adequate description. Modeling of adsorption isotherms revealed that the maximum calculated sorption capacity (qmax) according to the Langmuir model is 63.69 mg/g for AC, 41.32 mg/g for us-AC/Fe3O4 and 24.16 mg/g for h-AC/Fe3O4. The determination coefficients confirmed the optimal fit of the Langmuir model for describing the processes of Pb2+ adsorption on AC and us-AC/Fe3O4.
Therefore, it has been demonstrated that the ultrasonic modification of activated carbon with magnetite nanoparticles facilitates the production of magnetic composites that exhibit high sorption activity, comparable to that of the original carbon. An additional technological benefit of this process is the potential for magnetic separation. This renders us-AC/Fe3O4 a promising material for use in water purification systems from heavy metal ions.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/c11040083/s1, Figure S1: Determination of the PZC of AC (a), us-AC/Fe3O4 (b), h-AC/Fe3O4 (c) by the pH-drift method; Table S1: Adsorption capacity of AC and the composites h-AC/Fe3O4 and us-AC/Fe3O4 at different contact times; Figure S2: Magnetic separation of AC/Fe3O4 composite after adsorption; Figure S3: Elovich model for the sorption process of Pb ions on (a) AC sorbent, (b) h-AC/Fe3O4, and (c) us-AC/Fe3O4; Figure S4: Bangham model for the sorption process of Pb ions on (a) AC sorbent, (b) h-AC/Fe3O4, and (c) us-AC/Fe3O4; Figure S5: Boyd model for the sorption process of Pb ions on (a) AC sorbent, (b) h-AC/Fe3O4, and (c) us-AC/Fe3O4; Figure S6: Adsorption isotherms (qₑ vs. Cₑ) for Pb ions on (a) AC sorbent, (b) h-AC/Fe3O4, and (c) us-AC/Fe3O4.

Author Contributions

Conceptualization, G.S. and A.L.; methodology, G.S.; software, A.I.; validation, G.S., A.K. and A.B.; formal analysis, A.B.; investigation, A.A.; resources, R.K.; data curation, A.A.; writing—original draft preparation, A.I.; writing—review and editing, Z.M.; visualization, A.I.; supervision, Z.M.; project administration, A.L.; funding acquisition, A.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Committee of Science of the Ministry of Science and Higher Education of the Republic of Kazakhstan under Grant No. AP19576899.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

ACActivated carbon
BETBrunauer–Emmett–Teller
BCBiochar
BJHBarrett–Joyner–Halenda method
CCarbon
EDSEnergy-dispersive X-ray spectroscopy
ESEggshell/starch
SEMField-emission scanning electron microscopy
GOGraphene oxide
HAPHydroxyapatite
JCPDSJoint Committee on Powder Diffraction Standards
NCNanoclay
PANiPolyaniline
PZCThe point of zero charge
SCCMStandard Cubic Centimeter per Minute
SEMScanning electron microscopy
XRDX-ray diffraction

References

  1. World Health Organization. Arsenic. Available online: https://www.who.int/news-room/fact-sheets/detail/arsenic (accessed on 25 July 2025).
  2. Ng, J.C.; Wang, J.; Shraim, A. A Global Health Problem Caused by Arsenic from Natural Sources. Chemosphere 2003, 52, 1353–1359. [Google Scholar] [CrossRef]
  3. World Health Organization. Lead Poisoning and Health. Available online: https://www.who.int/news-room/fact-sheets/detail/lead-poisoning-and-health (accessed on 25 July 2025).
  4. Bjørklund, G.; Chirumbolo, S.; Dadar, M.; Pivina, L.; Lindh, U.; Butnariu, M.; Aaseth, J. Mercury Exposure and Its Effects on Fertility and Pregnancy Outcome. Basic Clin. Pharmacol. Toxicol. 2019, 125, 317–327. [Google Scholar] [CrossRef]
  5. Nawaz, W.; Bashir, H. Managing the Unintended Consequences of Radical Sustainability Innovations: The Case of Cat-astrophic Failure of Leaded Gasoline Industry. J. Clean. Prod. 2022, 375, 134175. [Google Scholar] [CrossRef]
  6. Galindo, J.M.; Andreu, C.M.; Merino, S.; Herrero, M.A.; Vázquez, E.; Sánchez-Migallón, A.M.; Castañeda, G. Few-Layer Graphene-Hybrid Sulfonate Hydrogels for High-Efficient Adsorption of Heavy Metal (Pb2+, Ni2+, Cd2+) in Water Treatment. Ecotoxicol. Environ. Saf. 2025, 292, 117934. [Google Scholar] [CrossRef] [PubMed]
  7. Taharia, M.; Dey, D.; Das, K.; Sukul, U.; Chen, J.-S.; Banerjee, P.; Dey, G.; Sharma, R.K.; Lin, P.-Y.; Chen, C.-Y. Microbial Induced Carbonate Precipitation for Remediation of Heavy Metals, Ions and Radioactive Elements: A Comprehensive Exploration of Prospective Applications in Water and Soil Treatment. Ecotoxicol. Environ. Saf. 2024, 271, 115990. [Google Scholar] [CrossRef]
  8. Sahu, P.; Patel, A.R.; Pandey, A.; Hait, M.; Patra, G.K. Assessment of Heavy Metal Ion Toxicity in Wastewater: A Com-prehensive Review. Inorganica Chim. Acta 2025, 585, 122751. [Google Scholar] [CrossRef]
  9. Soni, H.; Kanjariya, P.; Ballal, S.; Panigrahi, R.; Ariffin, I.A.; Tantawi, D.; Abosaoda, M.K.; Nathiya, D.; Jayabalan, K.; Chauhan, A.S. Recent Advances in the Synthesis of Magnetic Nanocomposites for the Adsorption of Heavy Metal Ions from Wastewater. J. Mol. Struct. 2025, 1345, 143019. [Google Scholar] [CrossRef]
  10. Ren, Z.-H.; Xiao, L.-P.; Wang, Q.; Fu, X.; Xu, Q.; He, H.; Qin, C.; Zheng, M.; Sánchez, J.; Sun, R.-C. Removal of Heavy Metal Ions in Wastewater and Reuse as a Hydrogenolysis Catalyst: Multistage Application of Lignin-Based Adsorbent. Ind. Crops Prod. 2025, 228, 120919. [Google Scholar] [CrossRef]
  11. Jahan, K.; Thankachan, D.; Shakya, K.; Mehrotra, N.; Nimish, C.S.; Verma, V. Removal of Heavy Metal Ions (Pb2+, Cu2+, Cr3+, and Cd2+) from Multimetal Simulated Wastewater Using 3-Aminopropyl Triethoxysilane Grafted Agar Porous Cryogel. Int. J. Biol. Macromol. 2024, 282, 136784. [Google Scholar] [CrossRef]
  12. Liu, M.; Lu, Q.; Yu, W. The Improvement of Heavy Metals Removal by Wood Membrane in Drinking Water Treatment: Comparison with Polymer Membrane and Associated Mechanism. Chemosphere 2023, 324, 138297. [Google Scholar] [CrossRef] [PubMed]
  13. Hai, A.; Bharath, G.; Patah, M.F.A.; AlMohamadi, H.; Banat, F.; Daud, W.M.A.W. Incorporation of Transition Met-al-Oxide on 3D Porous Carbon Network for Electrochemical Water Desalination and Heavy-Metals Removal. Electrochim. Acta 2025, 528, 146337. [Google Scholar] [CrossRef]
  14. Bandehali, S.; Moghadassi, A.; Hosseini, S.M.; Parvizian, F.; Ghanbari, D.; Rahimizadeh, R.; Ebrahimi, A.A. Smart Ion Exchange Membrane with High Impact in Heavy Metals Separation Prepared by Electrospinning Process with Ultra-sensitive Responsiveness for Water Treatment. Chem. Eng. Res. Des. 2024, 205, 763–774. [Google Scholar] [CrossRef]
  15. De Napoli, I.E.; Zanetti, E.M.; Fragomeni, G.; Giuzio, E.; Audenino, A.L.; Catapano, G. Transport Modeling of Convec-tion-Enhanced Hollow Fiber Membrane Bioreactors for Therapeutic Applications. J. Membr. Sci. 2014, 471, 347–361. [Google Scholar] [CrossRef]
  16. Zhu, C.; Peng, Y.; Yang, W. Modification Strategies for Metal-Organic Frameworks Targeting at Membrane-Based Gas Separations. Green. Chem. Eng. 2021, 2, 17–26. [Google Scholar] [CrossRef]
  17. Cheng, P.; Moehring, N.K.; Idrobo, J.C.; Ivanov, I.N.; Kidambi, P.R. Scalable Synthesis of Nanoporous Atomically Thin Graphene Membranes for Dialysis and Molecular Separations via Facile Isopropanol-Assisted Hot Lamination. Nanoscale 2021, 13, 2825–2837. [Google Scholar] [CrossRef] [PubMed]
  18. Zhao, Y.; Wu, M.; Guo, Y.; Mamrol, N.; Yang, X.; Gao, C.; Van Der Bruggen, B. Metal-Organic Framework Based Mem-branes for Selective Separation of Target Ions. J. Membr. Sci. 2021, 634, 119407. [Google Scholar] [CrossRef]
  19. Bhol, P.; Yadav, S.; Altaee, A.; Saxena, M.; Misra, P.K.; Samal, A.K. Graphene-Based Membranes for Water and Wastewater Treatment: A Review. ACS Appl. Nano Mater. 2021, 4, 3274–3293. [Google Scholar] [CrossRef]
  20. Tarighat, M.A.; Barghandan, F.; Hashemifard, S.A.; Abdi, G. Development and Characterization of PAN/GO-Tyrosine Hollow Fiber Membranes for Enhanced Heavy Metal Adsorption and SPME-Spectrophotometric Detection. RSC Adv. 2025, 15, 3721–3737. [Google Scholar] [CrossRef]
  21. Li, P.; Li, Y.-X.; Wu, Y.-Z.; Xu, Z.-L.; Zhang, H.-Z.; Gao, P.; Xu, S.-J. Thin-Film Nanocomposite NF Membrane with GO on Macroporous Hollow Fiber Ceramic Substrate for Efficient Heavy Metals Removal. Environ. Res. 2021, 197, 111040. [Google Scholar] [CrossRef] [PubMed]
  22. Hu, Q.; Tang, D.; Xiang, Y.; Chen, X.; Lin, J.; Zhou, Q. Magnetic Ion-Imprinted Polyacrylonitrile-Chitosan Electro-Spun Nanofibrous Membrane as Recyclable Adsorbent with Selective Heavy Metal Removal and Antibacterial Fouling in Water Treatment. Int. J. Biol. Macromol. 2023, 241, 124620. [Google Scholar] [CrossRef]
  23. Ezzat, H.A.; Sebak, M.A.; Aladim, A.K.; Shahat, M.A. Innovative Experimental and Theoretical Strategies for Sustain-able Heavy Metal Ion Removal Using chitosan@TiO2 Composites Functionalized with Nanostructured Metal Oxides. J. Mol. Liq. 2025, 431, 127814. [Google Scholar] [CrossRef]
  24. Somyanonthanakun, W.; Ahmed, R.; Krongtong, V.; Thongmee, S. Studies on the Adsorption of Pb(II) from Aqueous Solutions Using Sugarcane Bagasse-Based Modified Activated Carbon with Nitric Acid: Kinetic, Isotherm and Desorption. Chem. Phys. Impact 2023, 6, 100181. [Google Scholar] [CrossRef]
  25. Nguyen, T.H.; Nguyen, T.H.; Nguyen, T.N.; Nghiem, X.S.; Le Minh, T.; Vu, A.-T. Preparation of Functional Materials Based on Activated Carbon from Bagasse and Its Application in Pb2+ Adsorption in an Aqueous Environment. Mater. Today Commun. 2025, 47, 113158. [Google Scholar] [CrossRef]
  26. Heidarinejad, Z.; Rahmanian, O.; Heidari, M. Production of KOH-Activated Carbon from Date Press Cake: Effect of the Activating Agent on Its Properties and Pb(II) Adsorption Potential. Desalination Water Treat. 2019, 165, 232–243. [Google Scholar] [CrossRef]
  27. Altıntıg, E.; Altundag, H.; Tuzen, M.; Sarı, A. Effective Removal of Methylene Blue from Aqueous Solutions Using Magnetic Loaded Activated Carbon as Novel Adsorbent. Chem. Eng. Res. Des. 2017, 122, 151–163. [Google Scholar] [CrossRef]
  28. Liu, X.; Tian, J.; Li, Y.; Sun, N.; Mi, S.; Xie, Y.; Chen, Z. Enhanced Dyes Adsorption from Wastewater via Fe3O4 Nano-particles Functionalized Activated Carbon. J. Hazard. Mater. 2019, 373, 397–407. [Google Scholar] [CrossRef] [PubMed]
  29. Mohammad, S.G.; Abulyazied, D.E.; Ahmed, S.M. Application of Polyaniline/Activated Carbon Nanocomposites De-rived from Different Agriculture Wastes for the Removal of Pb(II) from Aqueous Media. Desalination Water Treat. 2019, 170, 199–210. [Google Scholar] [CrossRef]
  30. Long, Y.; Jiang, J.; Hu, J.; Hu, X.; Yang, Q.; Zhou, S. Removal of Pb(Ⅱ) from Aqueous Solution by Hydroxyapatite/Carbon Composite: Preparation and Adsorption Behavior. Colloids Surf. A Physicochem. Eng. Asp. 2019, 577, 471–479. [Google Scholar] [CrossRef]
  31. Hashem, B.Y.; Alswat, A.A.; Ali, S.L.; Al-Odaini, N.A.; Alshorifi, F.T. Facile Synthesis of NiO–CuO/Activated Carbon Nanocomposites for Use in the Removal of Lead and Cadmium Ions from Water. ACS Omega 2022, 7, 47183–47191. [Google Scholar] [CrossRef]
  32. Mojoudi, F.; Hamidian, A.H.; Zhang, Y.; Yang, M. Synthesis and Evaluation of Activated Carbon/Nanoclay/Thiolated Graphene Oxide Nanocomposite for Lead(II) Removal from Aqueous Solution. Water Sci. Technol. 2019, 79, 466–479. [Google Scholar] [CrossRef]
  33. Zhang, Z.; Wang, T.; Zhang, H.; Liu, Y.; Xing, B. Adsorption of Pb(II) and Cd(II) by Magnetic Activated Carbon and Its Mechanism. Sci. Total Environ. 2021, 757, 143910. [Google Scholar] [CrossRef]
  34. Saeidi, N.; Parvini, M.; Niavarani, Z. High Surface Area and Mesoporous Graphene/Activated Carbon Composite for Adsorption of Pb(II) from Wastewater. J. Environ. Chem. Eng. 2015, 3, 2697–2706. [Google Scholar] [CrossRef]
  35. Wibowo, Y.G.; Anwar, D.; Safitri, H.; Surya, I.; Sudibyo, S.; Yuliansyah, A.T.; Murti Petrus, H.T.B. Functionalized Magnetite-Biochar with Live and Dead Bacteria for Adsorption-Biosorption of Highly Toxic Metals: Cd, Hg, and Pb. Next Mater. 2025, 6, 100487. [Google Scholar] [CrossRef]
  36. Manyangadze, M.; Chikuruwo, N.H.M.; Narsaiah, T.B.; Chakra, C.S.; Radhakumari, M.; Danha, G. Enhancing Adsorp-tion Capacity of Nano-Adsorbents via Surface Modification: A Review. S. Afr. J. Chem. Eng. 2020, 31, 25–32. [Google Scholar] [CrossRef]
  37. Sibiya, N.P.; Mahlangu, T.P.; Tetteh, E.K.; Rathilal, S. Review on Advancing Heavy Metals Removal: The Use of Iron Oxide Nanoparticles and Microalgae-Based Adsorbents. Clean. Chem. Eng. 2025, 11, 100137. [Google Scholar] [CrossRef]
  38. Shah, S.I.A.; Ahmad, W.; Anwar, M.; Shah, R.; Khan, J.A.; Shah, N.S.; Al-Anazi, A.; Han, C. Synthesis, Properties, and Applications of Fe3O4 and Fe3O4-Based Nanocomposites: A Review. Appl. Catal. O Open 2025, 203, 207049. [Google Scholar] [CrossRef]
  39. Rudravarapu, K.; Yarramuthi, V.; Munagapati, V.S.; Gutha, Y.; Wen, J.-C.; Prasad, C.; Pilla, V.A.R.; Gumma, L.; Won, J.S.; Choi, H.Y. Recent Progresses in the Development of Magnetic Fe3O4 Nanoparticles Supported Cellulose-Based Composites: A Review. Inorg. Chem. Commun. 2025, 182, 115588. [Google Scholar] [CrossRef]
  40. Kolluru, S.S.; Agarwal, S.; Sireesha, S.; Sreedhar, I.; Kale, S.R. Heavy Metal Removal from Wastewater Using Nano-materials-Process and Engineering Aspects. Process Saf. Environ. Prot. 2021, 150, 323–355. [Google Scholar] [CrossRef]
  41. Tee, G.T.; Gok, X.Y.; Yong, W.F. Adsorption of Pollutants in Wastewater via Biosorbents, Nanoparticles and Magnetic Biosorbents: A Review. Environ. Res. 2022, 212, 113248. [Google Scholar] [CrossRef]
  42. Akbar, A.; Bhavani Lakshmi, M.; Das, T.K.; Ghosh, M. Spinel Ferrites in the Photocatalytic and Adsorptive Remediation of Dyes and Heavy Metals: A Review. J. Water Process Eng. 2025, 71, 107259. [Google Scholar] [CrossRef]
  43. Wang, C.; Wang, H.; Cao, Y. Pb(II) Sorption by Biochar Derived from Cinnamomum Camphora and Its Improvement with Ultrasound-Assisted Alkali Activation. Colloids Surf. A Physicochem. Eng. Asp. 2018, 556, 177–184. [Google Scholar] [CrossRef]
  44. Latif, A.; Sheng, D.; Sun, K.; Si, Y.; Azeem, M.; Abbas, A.; Bilal, M. Remediation of Heavy Metals Polluted Environment Using Fe-Based Nanoparticles: Mechanisms, Influencing Factors, and Environmental Implications. Environ. Pollut. 2020, 264, 114728. [Google Scholar] [CrossRef]
  45. Yoo, J.; Kim, H.-S.; Park, S.-Y.; Kwon, S.; Lee, J.; Koo, J.; Seo, Y.-S. Instantaneous Integration of Magnetite Nanoparticles on Graphene Oxide Assisted by Ultrasound for Efficient Heavy Metal Ion Retrieval. Ultrason. Sonochemistry 2020, 64, 104962. [Google Scholar] [CrossRef]
  46. Liu, C.; Yan, X.; Zhang, H.-X.; Yang, J.; Yoon, K.-B. Biochars and Modified-Biochars for Toxic-Metal/Metalloid Ions Sorption in Various Mixed Solution Systems: A Review on Kinetic and Isotherm Models. Desalination Water Treat. 2024, 319, 100404. [Google Scholar] [CrossRef]
  47. Raturi, S.; Kumari, S.; András, K.; Khargotra, R.; Sebestyén, V.; Singh, T. Advancements of Nanotechnological Strategies as Conventional Approach for Heavy Metal Removal from Industrial Wastewater: Start-of-the-Art Review. Curr. Res. Green Sustain. Chem. 2024, 9, 100428. [Google Scholar] [CrossRef]
  48. Yeleuov, M.; Seidl, C.; Temirgaliyeva, T.; Taurbekov, A.; Prikhodko, N.; Lesbayev, B.; Sultanov, F.; Daulbayev, C.; Ku-mekov, S. Modified Activated Graphene-Based Carbon Electrodes from Rice Husk for Supercapacitor Applications. Energies 2020, 13, 4943. [Google Scholar] [CrossRef]
  49. Mansurov, Z.; Smagulova, G.; Kaidar, B.; Imash, A.; Lesbayev, A. PAN—Composite Electrospun-Fibers Decorated with Magnetite Nanoparticles. Magnetochemistry 2022, 8, 160. [Google Scholar] [CrossRef]
  50. Chowdhury, I.R.; Chowdhury, S.; Mazumder, M.A.J.; Al-Ahmed, A. Removal of Lead Ions (Pb2+) from Water and Wastewater: A Review on the Low-Cost Adsorbents. Appl. Water Sci. 2022, 12, 185. [Google Scholar] [CrossRef]
  51. Wang, Y.; Wu, D.; Wei, Q.; Wei, D.; Yan, T.; Yan, L.; Hu, L.; Du, B. Rapid Removal of Pb(II) from Aqueous Solution Using Branched Polyethylenimine Enhanced Magnetic Carboxymethyl Chitosan Optimized with Response Surface Method-ology. Sci. Rep. 2017, 7, 10264. [Google Scholar] [CrossRef]
  52. Li, D.; Zhou, L. Adsorption of Heavy Metal Tolerance Strains to Pb2+ and Cd2+ in Wastewater. Env. Sci. Pollut. Res. 2018, 25, 32156–32162. [Google Scholar] [CrossRef]
  53. Al-Maliky, E.A.; Gzar, H.A.; Al-Azawy, M.G. Determination of Point of Zero Charge (PZC) of Concrete Particles Ad-sorbents. IOP Conf. Ser. Mater. Sci. Eng. 2021, 1184, 012004. [Google Scholar] [CrossRef]
  54. Putra, N.R.; Zaini, M.A.A.; Kusuma, H.S.; Darmokoesoemo, H.; Faizal, A.N.M. Advances in Chromium Removal Using Biomass-derived Activated Carbon: A Comprehensive Review and Bibliometric Analysis. Env. Prog. Sustain. Energy 2025, 44, e14598. [Google Scholar] [CrossRef]
  55. Kozyatnyk, I.; Yakupova, I. Impact of Chemical and Physical Treatments on the Structural and Surface Properties of Activated Carbon and Hydrochar. ACS Sustain. Chem. Eng. 2025, 13, 2500–2507. [Google Scholar] [CrossRef]
  56. Sinha, P.; Banerjee, S.; Kar, K.K. Characteristics of Activated Carbon. In Springer Series in Materials Science; Springer International Publishing: Cham, Switzerland, 2020; pp. 125–154. [Google Scholar] [CrossRef]
  57. Mansūrov, Z.A. (Ed.) Carbon Nanomaterials in Biomedicine and the Environment; Jenny Stanford Publishing Pte. Ltd.: Singapore, 2020. [Google Scholar]
  58. Yadav, B.K.; Kumar, A. Study of Graphitic Crystalline Structure in Highly Porous Activated Carbons Derived from Rice Husk Biomass. Chem. Pap. 2025, 79, 5451–5464. [Google Scholar] [CrossRef]
  59. Nwanya, A.C.; Musheghyan-Avetisyan, A.; György, E.; Pérez Del Pino, Á. Fabrication of Asymmetric Supercapacitors by Laser Processing of Activated Carbon-Based Electrodes Produced from Rice Husk Waste. Surf. Interfaces 2024, 54, 105200. [Google Scholar] [CrossRef]
  60. Bagheri, A.R.; Ghaedi, M.; Asfaram, A.; Bazrafshan, A.A.; Jannesar, R. Comparative Study on Ultrasonic Assisted Ad-sorption of Dyes from Single System onto Fe3O4 Magnetite Nanoparticles Loaded on Activated Carbon: Experimental Design Methodology. Ultrason. Sonochemistry 2017, 34, 294–304. [Google Scholar] [CrossRef]
  61. Zhuang, L.; Zhang, W.; Zhao, Y.; Shen, H.; Lin, H.; Liang, J. Preparation and Characterization of Fe3O4 Particles with Novel Nanosheets Morphology and Magnetochromatic Property by a Modified Solvothermal Method. Sci. Rep. 2015, 5, 9320. [Google Scholar] [CrossRef] [PubMed]
  62. Kurien, U.; Hu, Z.; Lee, H.; Dastoor, A.P.; Ariya, P.A. Radiation Enhanced Uptake of Hg0(g) on Iron (Oxyhydr)Oxide Nanoparticles. RSC Adv. 2017, 7, 45010–45021. [Google Scholar] [CrossRef]
  63. Moradi, B.; Wang, D.; Botte, G.G. Carbon-Coated Fe3O4 Nanospindles as Solid Electrolyte Interface for Improving Graphite Anodes in Lithium Ion Batteries. J. Appl. Electrochem. 2020, 50, 321–331. [Google Scholar] [CrossRef]
  64. Hariani, P.L.; Faizal, M.; Ridwan; Marsi; Setiabudidaya, D. Removal of Procion Red MX-5B from Songket’s Industrial Wastewater in South Sumatra Indonesia Using Activated Carbon-Fe3O4 Composite. Sustain. Environ. Res. 2018, 28, 158–164. [Google Scholar] [CrossRef]
  65. Arrebola, J.C.; Caballero, A.; Hernán, L.; Morales, J. Graphitized Carbons of Variable Morphology and Crystallinity: A Comparative Study of Their Performance in Lithium Cells. J. Electrochem. Soc. 2009, 156, A986. [Google Scholar] [CrossRef]
  66. Teng, Z.; Zeng, S.; Feng, W.; Zhu, L.; Tan, Y.; Han, X.; Chen, C.; Zhang, H. Facile Synthesis and Enhanced Microwave Absorption Properties of Fe-Fe3C@C Composites. J. Mater. Sci. Mater. Electron. 2019, 30, 14573–14579. [Google Scholar] [CrossRef]
  67. Abdisattar, A.; Yerdauletov, M.; Yeleuov, M.; Napolskiy, F.; Merkulov, A.; Rudnykh, A.; Nazarov, K.; Kenessarin, M.; Zhomartova, A.; Krivchenko, V. The Impact of Biowaste Composition and Activated Carbon Structure on the Electro-chemical Performance of Supercapacitors. Molecules 2024, 29, 5029. [Google Scholar] [CrossRef]
  68. Ji, S.; Miao, C.; Liu, H.; Feng, L.; Yang, X.; Guo, H. A Hydrothermal Synthesis of Fe3O4@C Hybrid Nanoparticle and Magnetic Adsorptive Performance to Remove Heavy Metal Ions in Aqueous Solution. Nanoscale Res. Lett. 2018, 13, 178. [Google Scholar] [CrossRef]
  69. Chen, K.; Zhang, T.; Chen, X.; He, Y.; Liang, X. Model Construction of Micro-Pores in Shale: A Case Study of Silurian Longmaxi Formation Shale in Dianqianbei Area, SW China. Pet. Explor. Dev. 2018, 45, 412–421. [Google Scholar] [CrossRef]
  70. 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]
  71. Heo, Y.-J.; Park, S.-J. Synthesis of Activated Carbon Derived from Rice Husks for Improving Hydrogen Storage Capacity. J. Ind. Eng. Chem. 2015, 31, 330–334. [Google Scholar] [CrossRef]
  72. Dong, Z.; Li, B.; Shang, H.; Zhang, P.; Chen, S.; Yang, J.; Zeng, Z.; Wang, J.; Deng, S. Ultramicroporous Carbon Granules with Narrow Pore Size Distribution for Efficient CH4 Separation from Coal-bed Gases. AIChE J. 2021, 67, e17281. [Google Scholar] [CrossRef]
  73. Long, X.; Zhang, R.; Rong, R.; Wu, P.; Chen, S.; Ao, J.; An, L.; Fu, Y.; Xie, H. Adsorption Characteristics of Heavy Metals Pb2+ and Zn2+ by Magnetic Biochar Obtained from Modified AMD Sludge. Toxics 2023, 11, 590. [Google Scholar] [CrossRef]
  74. Ofomaja, A.E.; Naidoo, E.B.; Modise, S.J. Kinetic and Pseudo-Second-Order Modeling of Lead Biosorption onto Pine Cone Powder. Ind. Eng. Chem. Res. 2010, 49, 2562–2572. [Google Scholar] [CrossRef]
  75. Tsai, M.-L.; Lo, A.-Y.; Liu, J.-H.; Dai, Y.-M. Optimized Adsorptive Desulfurization Using Waste Tire-Derived Carbon. C 2025, 11, 47. [Google Scholar] [CrossRef]
  76. Awad, A.A.S. Equilibrium and Kinetics Investigations for Adsorption of Aqueous Lead Ions Using Olive Stone Waste. J. Ecol. Eng. 2025, 26, 114–128. [Google Scholar] [CrossRef]
  77. Hubbe, M.; Azizian, S.; Douven, S. Implications of Apparent Pseudo-Second-Order Adsorption Kinetics onto Cellulosic Materials: A Review. BioRes 2019, 14, 7582–7626. [Google Scholar] [CrossRef]
  78. Bao, X.; Qiang, Z.; Chang, J.-H.; Ben, W.; Qu, J. Synthesis of Carbon-Coated Magnetic Nanocomposite (Fe3O4@C) and Its Application for Sulfonamide Antibiotics Removal from Water. J. Environ. Sci. 2014, 26, 962–969. [Google Scholar] [CrossRef]
  79. Azizzadeh, S.E.; Bariki, S.G.; Movahedirad, S. Magnetic Orange Leaf Biochar for Favipiravir Removal from Wastewater. Sci. Rep. 2025, 15, 25388. [Google Scholar] [CrossRef] [PubMed]
  80. Bektaş, T.E.; Uğurluoğlu, B.K.; Tan, B. Phosphate Removal by Ion Exchange in Batch Mode. Water Pract. Technol. 2021, 16, 1343–1354. [Google Scholar] [CrossRef]
  81. Campos, N.F.; Barbosa, C.M.; Rodríguez-Díaz, J.M.; Duarte, M.M. Removal of Naphthenic Acids Using Activated Charcoal: Kinetic and Equilibrium Studies. Adsorpt. Sci. Technol. 2018, 36, 1405–1421. [Google Scholar] [CrossRef]
  82. Wu, F.-C.; Tseng, R.-L.; Juang, R.-S. Characteristics of Elovich Equation Used for the Analysis of Adsorption Kinetics in Dye-Chitosan Systems. Chem. Eng. J. 2009, 150, 366–373. [Google Scholar] [CrossRef]
  83. Nworie, F.S.; Nwabue, F.I.; Oti, W.; Mbam, E.; Nwali, B.U. Removal of methylene blue from aqueous solution using activated rice husk biochar: Adsorption isotherms, kinetics and error analysis. J. Chil. Chem. Soc. 2019, 64, 4365–4376. [Google Scholar] [CrossRef]
  84. Mishra, V. Modeling of Batch Sorber System: Kinetic, Mechanistic, and Thermodynamic Modeling. Appl. Water Sci. 2017, 7, 3173–3180. [Google Scholar] [CrossRef]
  85. Sumanjit; Rani, S.; Mahajan, R.K. Equilibrium, Kinetics and Thermodynamic Parameters for Adsorptive Removal of Dye Basic Blue 9 by Ground Nut Shells and Eichhornia. Arab. J. Chem. 2016, 9, S1464–S1477. [Google Scholar] [CrossRef]
  86. Hasani, N.; Selimi, T.; Mele, A.; Thaçi, V.; Halili, J.; Berisha, A.; Sadiku, M. Theoretical, Equilibrium, Kinetics and Thermodynamic Investigations of Methylene Blue Adsorption onto Lignite Coal. Molecules 2022, 27, 1856. [Google Scholar] [CrossRef] [PubMed]
  87. Benjelloun, M.; Miyah, Y.; Akdemir Evrendilek, G.; Zerrouq, F.; Lairini, S. Recent Advances in Adsorption Kinetic Models: Their Application to Dye Types. Arab. J. Chem. 2021, 14, 103031. [Google Scholar] [CrossRef]
  88. Tan, I.A.W.; Hameed, B.H. Adsorption Isotherms, Kinetics, Thermodynamics and Desorption Studies of Basic Dye on Activated Carbon Derived from Oil Palm Empty Fruit Bunch. J. Appl. Sci. 2010, 10, 2565–2571. [Google Scholar] [CrossRef]
  89. Khayyun, T.S.; Mseer, A.H. Comparison of the Experimental Results with the Langmuir and Freundlich Models for Copper Removal on Limestone Adsorbent. Appl. Water Sci. 2019, 9, 170. [Google Scholar] [CrossRef]
  90. Kalam, S.; Abu-Khamsin, S.A.; Kamal, M.S.; Patil, S. Surfactant Adsorption Isotherms: A Review. ACS Omega 2021, 6, 32342–32348. [Google Scholar] [CrossRef]
  91. Osmari, T.A.; Gallon, R.; Schwaab, M.; Barbosa-Coutinho, E.; Severo, J.B.; Pinto, J.C. Statistical Analysis of Linear and Non-Linear Regression for the Estimation of Adsorption Isotherm Parameters. Adsorpt. Sci. Technol. 2013, 31, 433–458. [Google Scholar] [CrossRef]
  92. Imla Syafiqah, M.S.; Yussof, H.W. Kinetics, Isotherms, and Thermodynamic Studies on the Adsorption of Mercury (Ii) Ion from Aqueous Solution Using Modified Palm Oil Fuel Ash. Mater. Today Proc. 2018, 5, 21690–21697. [Google Scholar] [CrossRef]
  93. Salim, N.A.A.; Puteh, M.H.; Khamidun, M.H.; Fulazzaky, M.A.; Abdullah, N.H.; Yusoff, A.R.M.; Zaini, M.A.A.; Ahmad, N.; Lazim, Z.M.; Nuid, M. Interpretation of Isotherm Models for Adsorption of Ammonium onto Granular Activated Carbon. Biointerface Res. Appl. Chem. 2020, 11, 9227–9241. [Google Scholar] [CrossRef]
  94. Boparai, H.K.; Joseph, M.; O’Carroll, D.M. Kinetics and Thermodynamics of Cadmium Ion Removal by Adsorption onto Nano Zerovalent Iron Particles. J. Hazard. Mater. 2011, 186, 458–465. [Google Scholar] [CrossRef]
  95. El-Sheeta, M.A.; Goher, M.E.; Abd El-Moghny, M.G.; El-Deab, M.S. Efficient Elimination of Pb(II) Ions from Aqueous Solutions Using Magnetic Fe3O4-Nanoparticles/Activated Carbon Derived from Agricultural Waste. Desalination Water Treat. 2022, 258, 241–260. [Google Scholar] [CrossRef]
  96. Vaithianathan, R.; Anitha, P.; Ramachandran, A.; Sudha, R. A Green Synthesis of Fe3O4 Nanocomposites Loaded Castor Seed Shell Carbon for Lead(II) Removal from Water: Equilibrium and Kinetic Studies. Desalination Water Treat. 2022, 280, 271–281. [Google Scholar] [CrossRef]
  97. Hosseini, S.S.; Hamadi, A.; Foroutan, R.; Peighambardoust, S.J.; Ramavandi, B. Decontamination of Cd2+ and Pb2+ from Aqueous Solution Using a Magnetic Nanocomposite of Eggshell/Starch/Fe3O4. J. Water Process Eng. 2022, 48, 1021. [Google Scholar] [CrossRef]
  98. Rafiullah, W.; Liu, X.; Xu, J.; Khan, M.A.; Zheng, X.; Ge, X.; Wang, X. Fabrication of Fe3O4/Biochar Composites for Ef-fective Lead (II) Removal. Water Air Soil. Pollut. 2025, 236, 545. [Google Scholar] [CrossRef]
  99. Karunanayake, A.G.; Todd, O.A.; Crowley, M.; Ricchetti, L.; Pittman, C.U.; Anderson, R.; Mohan, D.; Mlsna, T. Lead and Cadmium Remediation Using Magnetized and Nonmagnetized Biochar from Douglas Fir. Chem. Eng. J. 2018, 331, 480–491. [Google Scholar] [CrossRef]
  100. Zadeh, K.K.; Jafari, D. Activated Carbon/Alginate/Fe3O4 Magnetic Nanocomposite as a Superior Functional Material for Removal of Lead from Aqueous Media. Biomass Conv. Bioref. 2024, 14, 19025–19043. [Google Scholar] [CrossRef]
Figure 1. Surface morphology (a,c,d) and elemental distribution (b) of the activated carbon sample via SEM and EDS analysis.
Figure 1. Surface morphology (a,c,d) and elemental distribution (b) of the activated carbon sample via SEM and EDS analysis.
Carbon 11 00083 g001
Figure 2. Raman spectrum of activated carbon derived from rice husk.
Figure 2. Raman spectrum of activated carbon derived from rice husk.
Carbon 11 00083 g002
Figure 3. Morphological features (a,b), elemental mapping (c,e) and particle size distribution (d) of Fe3O4 obtained by SEM and EDS.
Figure 3. Morphological features (a,b), elemental mapping (c,e) and particle size distribution (d) of Fe3O4 obtained by SEM and EDS.
Carbon 11 00083 g003
Figure 4. X-ray diffraction pattern of Fe3O4 nanoparticles.
Figure 4. X-ray diffraction pattern of Fe3O4 nanoparticles.
Carbon 11 00083 g004
Figure 5. SEM images (a,c,d) and EDS elemental mapping (b) of the us-AC/Fe3O4 composite synthesized via ultrasonic treatment.
Figure 5. SEM images (a,c,d) and EDS elemental mapping (b) of the us-AC/Fe3O4 composite synthesized via ultrasonic treatment.
Carbon 11 00083 g005
Figure 6. X-ray diffraction pattern of us-AC/Fe3O4.
Figure 6. X-ray diffraction pattern of us-AC/Fe3O4.
Carbon 11 00083 g006
Figure 7. SEM images (a,c,d) and EDS elemental mapping (b) of the h-AC/Fe3O4 composite synthesized via the hydrothermal treatment.
Figure 7. SEM images (a,c,d) and EDS elemental mapping (b) of the h-AC/Fe3O4 composite synthesized via the hydrothermal treatment.
Carbon 11 00083 g007
Figure 8. X-ray diffraction pattern of h-AC/Fe3O4.
Figure 8. X-ray diffraction pattern of h-AC/Fe3O4.
Carbon 11 00083 g008
Figure 9. N2 adsorption–desorption isotherms of the samples (a), and pore size distribution (b).
Figure 9. N2 adsorption–desorption isotherms of the samples (a), and pore size distribution (b).
Carbon 11 00083 g009
Figure 10. Adsorption capacity of AC, us-AC/Fe3O4 and h-AC/Fe3O4, composites for lead ions.
Figure 10. Adsorption capacity of AC, us-AC/Fe3O4 and h-AC/Fe3O4, composites for lead ions.
Carbon 11 00083 g010
Figure 11. Plots of pseudo-first-order (a) AC, (c) h-AC/Fe3O4, (e) us-AC/Fe3O4 and pseudo-second-order (b) AC, (d) h-AC/Fe3O4, (f) us-AC/Fe3O4 kinetic models for Pb2+ ion adsorption on the composites.
Figure 11. Plots of pseudo-first-order (a) AC, (c) h-AC/Fe3O4, (e) us-AC/Fe3O4 and pseudo-second-order (b) AC, (d) h-AC/Fe3O4, (f) us-AC/Fe3O4 kinetic models for Pb2+ ion adsorption on the composites.
Carbon 11 00083 g011
Figure 12. (a) Weber-Morris intraparticle diffusion model and (b) Bangham model for the sorption of Pb2+ ions on the AC sorbent.
Figure 12. (a) Weber-Morris intraparticle diffusion model and (b) Bangham model for the sorption of Pb2+ ions on the AC sorbent.
Carbon 11 00083 g012
Figure 13. (a) Weber-Morris intraparticle diffusion model and (b) Elovich model for the sorption of Pb2+ ions on the h-AC/Fe3O4 sorbent.
Figure 13. (a) Weber-Morris intraparticle diffusion model and (b) Elovich model for the sorption of Pb2+ ions on the h-AC/Fe3O4 sorbent.
Carbon 11 00083 g013
Figure 14. (a) Weber-Morris intraparticle diffusion model and (b) Bangham model for the sorption of Pb2+ ion on the us-AC/Fe3O4 sorbent.
Figure 14. (a) Weber-Morris intraparticle diffusion model and (b) Bangham model for the sorption of Pb2+ ion on the us-AC/Fe3O4 sorbent.
Carbon 11 00083 g014
Figure 15. Langmuir isotherm model (a) AC, (c) us-AC/Fe3O4 (e) h-AC/Fe3O4; Freundlich isotherm model (b) AC, (d) us-AC/Fe3O4 (f) h-AC/Fe3O4.
Figure 15. Langmuir isotherm model (a) AC, (c) us-AC/Fe3O4 (e) h-AC/Fe3O4; Freundlich isotherm model (b) AC, (d) us-AC/Fe3O4 (f) h-AC/Fe3O4.
Carbon 11 00083 g015
Table 1. Overview of modern carbon-based composite adsorbents for lead ion removal.
Table 1. Overview of modern carbon-based composite adsorbents for lead ion removal.
Composite SorbentAC PrecursorComponent Synthesis MethodComposite Synthesis ApproachPollutantAdsorption Capacity (mg/g)Ref.
AC/PANiWaste biomass from fruit skins and pitsPANi—PolymerizationPolymerizationPb (II)6.81[29]
AC/HAPCommercial productHAP—Co-precipitationCo-precipitationPb (II)401.58[30]
AC/NiO−CuOArabo-khata leavesNiO, CuO—Commercial productUltrasonic treatmentPb (II)182.78[31]
AC/NC/GOSugar beetNanoclay—Carbonization, graphene oxide—Hummers’ methodMixingPb (II)208[32]
ACRapeseed straw-Hydrothermal methodPb (II)253.2[33]
AC/GOGlucoseGO—Chemical exfoliation, Hummers’ methodUltrasonic treatmentPb(II)217[34]
Table 2. Comparison of kinetic parameters for Pb2+ ion adsorption by AC and AC/Fe3O4 composites.
Table 2. Comparison of kinetic parameters for Pb2+ ion adsorption by AC and AC/Fe3O4 composites.
SorbentPseudo-First-OrderPseudo-Second-Order
k1 (min−1)R2k2 (g·min·mg−1)R2
AC0.00010.86490.06320.9999
h-AC/Fe3O40.00650.72820.10820.9994
us-AC/Fe3O40.00210.88920.06710.9997
Table 3. Comparison of kinetic parameters for Pb2+ ion adsorption by AC, us-AC/Fe3O4 and h-AC/Fe3O4 composites in various models.
Table 3. Comparison of kinetic parameters for Pb2+ ion adsorption by AC, us-AC/Fe3O4 and h-AC/Fe3O4 composites in various models.
SorbentWeber-MorrisBoydBanghamElovich
kid1 mg/(g/min)R2
Step 1
kid2
mg/(g/min)
R2
Step 2
kB, min−1R2αk, mg/g·min−αR2α (mg g−1 min−1)β (g mg−1)R2
AC0.06480.83020.02660.99820.00020.86490.06811.4130.937~3.24 × 10302.00160.9113
h-AC/Fe3O40.91250.66640.09960.49050.00660.72830.1407440.190.913858.010.18830.9306
us-AC/Fe3O40.81820.97810.36870.98980.00220.88930.112874.430.938952.2380.33550.9215
Table 4. The isotherm parameters for Freundlich and Langmuir models for Pb2+ ion adsorption by AC and AC/Fe3O4 composites.
Table 4. The isotherm parameters for Freundlich and Langmuir models for Pb2+ ion adsorption by AC and AC/Fe3O4 composites.
SorbentLangmuirFreundlich
qmax, mg/gKL, L/mgR2nKF, mg/g·(L/mg)1/nR2
AC63.69437.13640.99975.333340.84130.8117
h-AC/Fe3O424.15460.04780.83731.87831.85690.8568
us-AC/Fe3O441.32230.13670.98998.410422.04450.9873
Table 5. Comparison of Fe3O4-composite sorbents synthesized from different supports for Pb(II) adsorption.
Table 5. Comparison of Fe3O4-composite sorbents synthesized from different supports for Pb(II) adsorption.
Composite SorbentSupport MaterialComponent Synthesis MethodComposite Synthesis ApproachPollutantAdsorption Capacity (mg/g)Ref.
AC/Fe3O4Rice straw-In situ co-precipitationPb(II)33[95]
AC/Fe3O4(2)Rice straw-In situ co-precipitationPb(II)68
AC/Fe3O4Castor seed shellFe3O4—Co-precipitationConventional co-precipitationPb(II)122[96]
ES/Fe3O4Eggshell + starchFe3O4—Co-precipitationUltrasonic-assistedPb(II)57.14[97]
BC/Fe3O4Tangerine peelFe3O4—Co-precipitationCo-precipitation on biocharPb(II)125.23[98]
BC/Fe3O4Waste wood-In situ precipitation on biocharPb(II)40.7[99]
C/Fe3O4Pinewood + feedstocks-Pyrolysis + magnetic precipitationPb(II)~30
C/Fe3O4Switchgrass-Pyrolysis + magnetic precipitationPb(II)~30
AC/Fe3O4Commercial AC-Co-precipitation with alginatePb(II)36.764[100]
AC/Fe3O4Rice husksFe3O4—Co-precipitationHydrothermal-assistedPb(II)23.89This work
AC/Fe3O4Rice husksFe3O4—Co-precipitationUltrasonic-assistedPb(II)39.15This work
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

Smagulova, G.; Imash, A.; Baltabay, A.; Keneshbekova, A.; Abdisattar, A.; Kazhdanbekov, R.; Lesbayev, A.; Mansurov, Z. Mechanistic Evaluation of Pb(II) Adsorption on Magnetic Activated Carbon/Fe3O4 Composites: Influence of Hydrothermal and Ultrasonic Synthesis Routes. C 2025, 11, 83. https://doi.org/10.3390/c11040083

AMA Style

Smagulova G, Imash A, Baltabay A, Keneshbekova A, Abdisattar A, Kazhdanbekov R, Lesbayev A, Mansurov Z. Mechanistic Evaluation of Pb(II) Adsorption on Magnetic Activated Carbon/Fe3O4 Composites: Influence of Hydrothermal and Ultrasonic Synthesis Routes. C. 2025; 11(4):83. https://doi.org/10.3390/c11040083

Chicago/Turabian Style

Smagulova, Gaukhar, Aigerim Imash, Akniyet Baltabay, Aruzhan Keneshbekova, Alisher Abdisattar, Ramazan Kazhdanbekov, Aidos Lesbayev, and Zulkhair Mansurov. 2025. "Mechanistic Evaluation of Pb(II) Adsorption on Magnetic Activated Carbon/Fe3O4 Composites: Influence of Hydrothermal and Ultrasonic Synthesis Routes" C 11, no. 4: 83. https://doi.org/10.3390/c11040083

APA Style

Smagulova, G., Imash, A., Baltabay, A., Keneshbekova, A., Abdisattar, A., Kazhdanbekov, R., Lesbayev, A., & Mansurov, Z. (2025). Mechanistic Evaluation of Pb(II) Adsorption on Magnetic Activated Carbon/Fe3O4 Composites: Influence of Hydrothermal and Ultrasonic Synthesis Routes. C, 11(4), 83. https://doi.org/10.3390/c11040083

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop