Next Article in Journal
Nafion-212 Membrane Solvated by Ethylene and Propylene Carbonates as Electrolyte for Lithium Metal Batteries
Next Article in Special Issue
Preparation of Cationic Polyacrylamide Suspension and Its Application in Oilfield Wastewater Treatment
Previous Article in Journal
Evaluation of Viscosity Changes and Rheological Properties of Diutan Gum, Xanthan Gum, and Scleroglucan in Extreme Reservoirs
Previous Article in Special Issue
A Comparative Study of Cr(VI) Sorption by Aureobasidium pullulans AKW Biomass and Its Extracellular Melanin: Complementary Modeling with Equilibrium Isotherms, Kinetic Studies, and Decision Tree Modeling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Myrica esculenta Leaf Extract—Assisted Green Synthesis of Porous Magnetic Chitosan Composites for Fast Removal of Cd (II) from Water: Kinetics and Thermodynamics of Adsorption

1
Department of Chemistry, Banasthali Vidyapith, Banasthali 304022, India
2
Department of Chemistry, Nirankari Baba Gurubachan Singh Memorial College, Sohna 122103, India
3
Departamento de Química, Facultad de Ciencias, Universidad de Tarapacá, Avda. General, Velásquez, Arica 1775, Chile
4
Department of Chemistry, University of Allahabad, Prayagraj 211002, India
*
Authors to whom correspondence should be addressed.
Polymers 2023, 15(21), 4339; https://doi.org/10.3390/polym15214339
Submission received: 1 September 2023 / Revised: 29 September 2023 / Accepted: 7 October 2023 / Published: 6 November 2023
(This article belongs to the Special Issue Recent Advances in Functional Polymer Materials for Water Treatment)

Abstract

:
Heavy metal contamination in water resources is a major issue worldwide. Metals released into the environment endanger human health, owing to their persistence and absorption into the food chain. Cadmium is a highly toxic heavy metal, which causes severe health hazards in human beings as well as in animals. To overcome the issue, current research focused on cadmium ion removal from the polluted water by using porous magnetic chitosan composite produced from Kaphal (Myrica esculenta) leaves. The synthesized composite was characterized by BET, XRD, FT-IR, FE-SEM with EDX, and VSM to understand the structural, textural, surface functional, morphological-compositional, and magnetic properties, respectively, that contributed to the adsorption of Cd. The maximum Cd adsorption capacities observed for the Fe3O4 nanoparticles (MNPs) and porous magnetic chitosan (MCS) composite were 290 mg/g and 426 mg/g, respectively. Both the adsorption processes followed second-order kinetics. Batch adsorption studies were carried out to understand the optimum conditions for the fast adsorption process. Both the adsorbents could be regenerated for up to seven cycles without appreciable loss in adsorption capacity. The porous magnetic chitosan composite showed improved adsorption compared to MNPs. The mechanism for cadmium ion adsorption by MNPs and MCS has been postulated. Magnetic-modified chitosan-based composites that exhibit high adsorption efficiency, regeneration, and easy separation from a solution have broad development prospects in various industrial sewage and wastewater treatment fields.

1. Introduction

Water is the main component for the survival of all living organisms. However, it is gravely endangered due to the massive amount of pollution caused by domestic, industrial, and agricultural actions. Water scarcity and quality have emerged as major issues for long-term development [1,2]. Many water contaminants, including organic and inorganic constituents, have been mentioned in the literature [3]. Heavy metal-contaminated water bodies pose serious problems owing to their toxic behavior and bioaccumulation [4,5,6,7].
Cadmium is one of the non-essential heavy metals that is regarded as highly noxious and carcinogenic due to its non-biodegradability and ability to bioaccumulate in the environment. It is classified as a human carcinogen by the Environmental Protection Agency, (EPA), US, with a maximum permissible limit of 0.005 mg/L in drinking water [8].
Various methods for removing Cd (II) from the aqueous environment that is often used include chemical precipitation, membrane treatment, ion exchange, electro-dialysis, membrane flotation, electrochemical methods, ultra-filtration, etc. [9]. However, these processes, have several limitations, including lower effectiveness, high energy requirements, expensive disposal, and incomplete removal. Adsorption is extensively used for wastewater treatment due to its, quick response, low cost, simple operation, absence of sludge production, and reusability [10,11,12].
Various types of adsorbents have been used to remove Cd from wastewater. Among them, natural polysaccharide polymers, particularly chitosan, and its derivatives, have gained the attention of researchers owing to their renewable nature, sustainability, and adsorption efficiency [13,14,15]. Chitosan (CS), which is formed by the deacetylation of chitin, is present in the shells of crustacean crabs and shrimps, insect carapaces, and fungal and plant cell walls. CS, as one of nature’s most abundant biopolymers, has sparked widespread scientific interest due to its low cost, non-toxicity, inadequate hydrophilicity, biocompatibility, and biodegradability. It has many functional groups, including -NH2 and -OH, which facilitate good sorption of heavy metal ions. Raw chitosan, like many other widely used materials, cannot be efficiently separated from aqueous medium using traditional separation methods. As a result, attempts have been made to recycle CS by combining it with magnetic nanoparticles [16,17]. Because magnetic chitosan composites benefit from both chitosan (excellent adsorption performance) and magnetic material (easy magnetic separation), combining CS with a magnetic component is an effective way to address the above-mentioned inadequacies [18,19].
Due to their high chelating capacity and ease of magnetic separation, porous magnetic chitosan (MCS) material has been considered an efficient adsorbent for Cd removal [20]. A schematic representation of the application of MCS for the adsorption of Cd (II) and the reuse of the adsorbent is provided in Figure 1. The combination of MNPs and CS successfully excludes chitosan’s inherent drawbacks for application as an adsorbent, and its separation and reuse. Furthermore, the -OH groups on the surface of MNPs can form a hydrogen bonding network with chitosan by interacting with its -OH and -NH2 groups. As a result, the stability and integrity of the composite are retained even under acidic or alkaline environments [21]. Various chitosan-based materials were synthesized with magnetic NPs because of their lower internal diffusion resistance and higher specific surface area. MNPs, on the other side, get easily oxidized in air and are chemically active resulting in aggregation and loss of magnetism. Because of their small size, these NPs cause secondary contamination. Magnetic cores with larger sizes, on the other hand, can aid in the magnetic separation process [22]. Porous Magnetic chitosan composites have been used extensively and successfully to eliminate pollutants such as dyes, heavy metal ions, and other organic contaminants.
During the modification process, functional groups such as hydroxyl and carboxyl groups increased. Various studies are performed for the removal of Cd (II) by magnetically modified material, and the results showed that after the material was magnetically modified, the pH value, specific surface area, and polar oxygen-containing functional groups all increased, resulting in a saturated adsorption capacity. The mechanism of MCS to remove Cd is surface complexation and electrostatic adsorption. The strong affinity of iron oxide for Cd can enhance the complexation between them.
Plant extracts have been proposed as an easy and convenient alternative to chemical and physical methods for the preparation of metallic nanoparticles and their composites in recent years owing to the concerns of sustainability and environmental deterioration using conventional non-renewable sources. MNPs are one of these nanoparticles that have caught the interest of many researchers. Myrica esculenta leaf extract contains a variety of bioactive phytoconstituents, including phenolic compounds, glycosides, alkaloids, triterpenoids, and volatile oils [23].
To the best of our knowledge, there has not been much research done to assess the effectiveness of magnetic chitosan employing leaf extract of Myrica esculenta (Kaphal) for removing cadmium from aqueous solutions. Therefore, our aim is to develop a green synthetic process to produce chitosan—Fe3O4 composite using Myrica esculenta leaf extract as a reducing agent. This has resulted in the development of a novel porous magnetic chitosan composite in the present work that turned out to be an efficient material for Cd (II) ion removal. The MCS composite as well as the Fe3O4 nanoparticles were characterized using XRD, BET sorptometry, FT-IR, FE-SEM, EDX, and VSM for evaluating their physicochemical properties and also for determining the effect of such properties on the adsorption process parameters, namely, contact time, pH, adsorbent dose, and initial concentration of the analyte (Cd2+). The potential of MNPs and magnetic chitosan composite as potential adsorbents for Cd (II) was further probed using kinetic models and adsorption isotherms. Moreover, the reusability of the composite was investigated up to seven cycles.

2. Materials and Method

2.1. Materials

FeCl2·4H2O, Chitosan (degree of deacetylation ≥ 95%; viscosity: 100–200 mPa.s), FeCl3·6H2O, and acetic acid (CH3COOH), used during the preparation for adsorbents were of analytical-grade and were procured from Sigma Aldrich (St. Louis, MO, USA). Cd (NO3)2·4H2O was purchased from E. Merck, Mumbai, India. All aqueous solutions for the adsorption studies were prepared using Millipore Milli-Q® ultrapure water (Jaipur, Rajasthan). As Cd (II) has been chosen as the model contaminant in this investigation, a solution of Cd (II) of desired concentrations was prepared and used as an adsorbate solution. Dried Kaphal leaves were sourced from the farmlands of Uttarakhand.

2.2. Preparation of Aqueous Leaf Extract of Myrica esculenta (Kaphal)

To remove any pollutants, the Kaphal leaves were thoroughly washed with distilled water after being collected from the highlands of Uttarakhand, India. The leaves were manually cut into little pieces, air-dried for a week at room temperature, and then manually ground in a home kitchen grinder; 4 g of the resulting fine powder was then added into double-distilled water and stirred at 70 °C for 45 min. It was then filtered, and the filtrate was stored at 4 °C (Figure 2). The synthesis process for Fe3O4 nanoparticles and Fe3O4/Chitosan Composite is shown in the Supplementary file.

2.3. Characterization of the Adsorbents

The XRD of the MNPs and the CS composite were recorded on (Bruker D8 Discover X-ray Diffractometer, Karlsruhe, Germany) using Cu Kα radiation. The XRD pattern was recorded in the 2θ range of 10° and 70°.
By using a surface area analyzer N2 adsorption desorption studies were carried out on Quantachrome Autosorb iQ Surface Analyzer, CIQTEK, Hefei, China). For the analysis of the surface area, the samples were first degassed for 3.0 h to remove volatile gases and were then placed in a surface area analyzer for N2 adsorption-desorption.
A Perkin-Elmer FT–IR spectrometer was used to record the FTIR spectra. A spectrum of the adsorbents in the range of 400–4000 cm−1 to characterize the nature of chemical bonding and the type of surface functional group.
FE-SEM was used to investigate the morphology of MNPs and MCS composites as well as their elemental composition. FE-SEM equipped with EDAX (FEI QUANTA FEG250, Oregon, OH, USA) and an INCA Energy X-MAX-50, Oregon, OH, USA) was employed to characterize the morphology and chemical composition of the adsorbents used in the study.
A LakeShore 7404 (Lakeshore cryotronics, Westerville, OH, USA) vibrating sample magnetometer was used to determine the magnetization properties of the adsorbents. The magnetic sample is positioned on the sample holder and placed between the electromagnetic poles, normally horizontally, for VSM measurement.

2.4. Batch Adsorption Studies for the Adsorption of Cd (II) Using MNPs and Chitosan/Fe3O4 Composite as Adsorbents

To study the optimal adsorption performance and investigate the adsorption mechanism, batch adsorption experiments were performed. Different parameters were examined in fixed ranges, including, adsorbent dose (0.01–0.1 g), pH (2–11), temperature (303, 313, and 323 K), contact time (5–50 min), initial metal concentration (10–100 mgL−1), and adsorption/desorption studies, to better understand the potential of MNPs and MCS as adsorbents for Cd (II). The C0 (initial concentration) values were used to conduct the adsorption isotherms. After the adsorption process, atomic absorption spectroscopy was utilized to examine the heavy metal ion concentration remaining in the filtrate. The removal effectiveness (%) and equilibrium adsorption capacity (qe) of the adsorbents, namely, MNPs and MCS composite. For Cd (II) was given by Equations (1) and (2).
%   A d s o r p t i o n = C 0 C e × 100 C 0
A d s o r p t i o n   C a p a c i t y q e = C 0 C e × V m
where, V is the volume of metal ion solution, m is the mass of adsorbent dose, C0 is the initial metal ion concentrations, Ce is the residual Cd (II) ion concentration, and qe is the adsorption capacity at Ce, respectively.

3. Result and Discussion

3.1. Adsorbent Characterization

3.1.1. X-ray Diffraction

To understand the chemical and physical structure of the magnetic particles encapsulated in the CS matrix, X-ray diffraction (XRD) is a highly valuable technique. Figure 3a,b represent the XRD patterns of MNPs and porous MCS composite. The crystalline phase of magnetic NPs and porous magnetic chitosan composite were analyzed by XRD.
The diffraction peaks at 2θ values of 30.2°, 35.6°, 43.2°, 53.6°, 57.2°, and 62.9° correspond to the (220), (311), (400), (422), (511), and (440) planes of MNPs (JCPDS file number 01-075-0033) [24]. These peaks closely resemble the standard XRD pattern of Fe3O4. Thus, the XRD pattern demonstrated the formation of magnetic particles. Diffraction peaks at 2θ values in the range of 21° to 28° were attributed to amorphous chitosan (Figure 3b). Moreover, the peaks typical of Fe3O4 were also present in the composite as expected [25].

3.1.2. BET Sorptometry for Evaluating the Textural Properties of Adsorbents

N2 gas adsorption-desorption isotherms were observed in the relative pressures (p/p0) values ranging from 0 to 1 to estimate the surface area and porosity values of the synthesized adsorbents. BET sorptometry was performed to investigate the average pore radius, surface area, and pore volume of the material. About 25.0 mg of material was degassed at 300 °C for analysis. According to IUPAC standards, the N2 adsorption-desorption isotherms are Type IV, which agrees with the mesoporous nature of the composite [26]. The surface area and total pore volume of MNPs and MCS were 105 m2/g, 173 m2/g, and 0.3410 cc/g, and 0.4305 cc/g, respectively (Table 1). The Langmuir surface area of the MNPs and MCS was 201.084 m2/g and 536.934 m2/g, respectively (Figure 4). Thus, the higher surface area is responsible for the higher adsorption capacity of porous MCS compared to MNPs.
The Langmuir curve and BET multipoint of MNPs and porous MCS are expressed in detail in the Supporting Information (Figure S3).

3.1.3. Fourier Transform Infrared (FT-IR) Spectroscopy

For understanding the nature of chemical bonding and the kind of surface functionalities the adsorbents used in this study were examined using FT-IR spectroscopy. The FT-IR spectra of chitosan and MNPs are shown in Figure 5a. Likewise, the FT-IR spectra of the MCS composite before and after the adsorption process are shown in Figure 5b.
The absorption bands in the range of 3200–3400 cm−1 corresponded to N–H and O–H stretching vibrations of CS (Figure 5a). The bands at 2922 cm−1 and 2861 cm−1 were due to the C-H symmetric and unsymmetric stretching vibrations of the –CH2 groups in CS. The -NH deformation, C-N axial deformation, -CH3 bending vibration, and stretching vibration of C-O-C in the chitosan structure appeared at, around 1413, 1367, and 1021 cm−1. The existence of chitosan in the composite and its structural integrity was confirmed by the appearance of peaks at 1652 cm−1 and 1550 cm−1, which corresponds to the N-H bending vibration of primary amine [27].
The absorption bands characteristic of MNPs were observed in the FT-IR spectrum (Figure 5a). Metal oxygen bonds typical of MNPs were observed in the range of 400 and 850 cm−1. The absorption band at 546 cm−1 was attributed to the Fe-O stretching vibration of MNPs. This band is very sharp and has strong intensity, indicating the crystallinity of the sample. The broad bands around 3404 cm−1 and 1612 cm−1 were due to the O–H stretching vibration of surface-adsorbed water molecules. The band located around 1388 cm−1 and 1076 cm−1 were attributed to the unsymmetric and symmetric stretching vibrations of COO- [28].
Similar peaks were observed in the FT-IR spectrum of the MCS composite. In addition, some specific bands correspond to the original formation of new bonds between chitosan and Fe3O4 and some bands of chitosan have moved slightly from their original positions owing to atomic-level interactions between chitosan and MNPs. In the FT-IR spectrum of the MCS composite (Figure 5b), specific bands were observed at 3370 cm−1 which was attributed to N-H stretching vibration. The band at 1652 and 1550 cm−1 was attributed to C=O stretching vibration in the CS, for the N-H scissoring from the primary amine due to free amino groups in the CS and was compared with the standard chitosan [29]. The peak at 552 cm−1 for the Fe-O group was due to bare magnetic nanoparticles. This spectrum (Figure 5b) unambiguously showed the presence of both CS and MNPs in the composite [30]. The NMR spectrum of Myrica esculenta leaf extract has been already discussed by Nguyan et al. [31].

3.1.4. FE-SEM-EDX

FE-SEM was used to find out the morphology of MNPs and MCS composite, and EDAX was utilized to determine the elemental composition. SEM and EDX images of both the adsorbents were displayed in Figure 6 (EDX result shown in Figure S4). Spherical-shaped particles with agglomeration were observed in synthesized magnetic NPs (Figure 6a) as well as MCS (Figure 6b) composite in the FE-SEM images. Fe3O4 NPs exhibited a smooth surface, and the particles had an irregular shape, as observed in Figure 6a. In Figure 6b, the FE-SEM image of MCS was shown to exhibit a rough, granular surface after chitosan was bonded to Fe3O4. Agglomeration of the particles on the surface of the composite cannot be ruled out [32]. EDX spectra of the Fe3O4 sample confirm the presence of iron (Figure S4). The EDX spectra of MNPs and MCS are shown in Supporting Information (Figure S4). The mass percentage of carbon in MCS composite is higher than in MNPs, which proved MCS composite had lesser hydrophilicity properties.

3.1.5. Vibrating Sample Magnetometer Study for Evaluating the Magnetic Property of the Adsorbents

To evaluate the adsorbents’ magnetic properties, VSM (Vibrating-sample magnetometer) tests were utilized [20]. The magnetic properties of investigated magnetic materials were studied by measuring the magnetization curves (Figure 7). The results showed that all magnetization curves pass through the origin, which indicated that there was no residual magnetization occurring in test samples and these materials have super-paramagnetism [33]. By changing H between Oe +10,000 and −10,000 Oe, magnetization hysteresis was produced. MNPs had a saturation magnetization of 55.070 emu/g. The saturation magnetization of MCS decreased to 24.186 emu/g after the formation of composite between MNPs and chitosan, indicating that the adsorbent was still super-paramagnetic and that it would be easier to separate the adsorbent from solution by gravity and magnetism within a short period of time [34]. The lower Ms value for MCS could be explained by the quenching of surface magnetic moment in the material owing to the presence of non-magnetic species, such as CS as expected [35]. Even though MCS’ saturation magnetization value was obviously lower than that of MNPs, the application of a magnet allowed the adsorbent to immediately aggregate and to be separated in the 20 s [32].

3.2. Adsorption Experiments

3.2.1. Study of Kinetics of Adsorption of Cd (II) onto the Adsorbents, MNPs, and MCS Composite

To find out the appropriate rate expression and adsorption mechanism, the adsorption kinetics was examined [36]. Several models were examined for the adsorption of contaminants in water onto the surface of adsorbents and also to identify the main mechanism of such adsorption process and the kinetics of adsorption [37]. Adsorption kinetics was investigated in this work by fitting Lagergren’s pseudo-first-order (PFO) and Ho’s second-order models. The adsorption process was determined using the Elovich and intra-particle diffusion (IPD) models [38].
The PFO kinetic model represents weak interaction between sorbate and sorbent predominantly proceeding via physisorption. The PFO kinetic model is represented below [39]
ln (qe − qt) = lnqe − k1 × t
PSO kinetic model is based on chemisorption. Pseudo–second order kinetics is represented as [40].
t/qt = 1/(k2 × qe2) + t/qe
The IPD model is given by Equation (5).
qt = kid × √t + C
The Elovich model is specified by Equation(6).
qt = (1/β) ln (α × β × t) + (1/β) lnt
where, k1 (min−1) and k2 (g/mg/min) are the rate constants of PFO and PSO, kid—intraparticle diffusion rate constant (mg/(gmin−0.5)), constant C is the y-intercept, α is the initial rate of adsorption, β is the desorption constant [41].
Four kinetic models were studied, and the kinetic parameters deduced from various models are summarized in Table S1 [42,43].
From the R2 analysis of kinetic models, the best-fitted model for Cd (II) adsorption on MNPs at 50 mg/L was Elovich, and at 100 mg/L was IPD. Also, from the R2 analysis of kinetic models, the PSO model was the best fit for MCS. The pseudo-second-order model’s correlation coefficient (R2) was observed to be significantly greater than other models employed for MCS adsorbent, which means that the mechanism of adsorption was governed by this model (Figure 8).

3.2.2. Adsorption Isotherm Models

Adsorption isotherms describe how adsorbents and adsorbates interact in aqueous media at the attained saturation point. The most popular isotherm models, including Langmuir, Freundlich, Dubinin–Radushkevich, and Temkin isotherm models, were utilized.
The Langmuir isotherm is characterized by single-layer sorption onto a surface with countless open sites without interaction between the adsorbate molecules under the assumption that the material is regular and homogeneous. The model’s primary problem is its assumption that the energies of the adsorbent sites at each location are uniform. Equation (7) describes the Langmuir isotherm’s linear for [44]
1 q e = 1 K L q m a x C e + 1 q m a x
where qmax represents the material’s maximum adsorption capacity, and KL is the Langmuir adsorption constant (L/mg) representing the attraction of binding sites [45].
RL, the dimensionless constant separation factor is employed to express the important characteristic of the Langmuir isotherm:
R L = 1 1 + K L C 0
The value of RL shows whether the adsorption isotherm is favorable, linear, or unfavorable. The value of RL was found in the range of 0.0717–0.898 for the MCS composite. This shows the efficient interaction between the MCS composite and cadmium ions [46]. Figure 9a shows the Langmuir plot for MNPs and MCS composite. The values of KL and qmax are calculated using the slope and intercept of the linear regression plot of 1/Ce vs. 1/qe, which also provides the R2 value, which indicates how well the experimental results correspond with the mathematical isotherm model.
If adsorption happens on heterogeneous surfaces, the Freundlich adsorption isotherm is an empirical equation that is utilized to describe multilayer (physisorption) adsorption as well as monolayer (chemisorption) [42].The linearized Freundlich isotherm equation is explained by Equation (8).
L o g q e = log K F   +   1 n log C e
where KF is the Freundlich constant of the adsorbent, and n is the adsorption intensity of the adsorbent. 1/n is the adsorption intensity which signifies the heterogeneity of the adsorbent sites as well as the relative distribution of energy. KF and n are dimensionless constants. The value of n > 1 and 1/n < 1 indicates the favorable condition for adsorption. In this study, the value of n ranged between 1.075 and 1.222 which is greater than 1 for MCS composite, indicating favorable adsorption. The graph of lnCeVs lnqe gives a straight line with intercept KF and slope 1/n as shown in Figure 9b. The R2 values are 0.97, 0.95, and 0.97 for MNPs at 303, 313, and 323 K and 0.94, 0.92, 0.92 for MCS at 303 K, 313 K and 323 K (Table S2) [45].
The Temkin Isotherm Model was used to characterize the adsorption considering the interaction between adsorbate and adsorbent, which resulted in a linear reduction in isotherm when the heat of adsorption of all molecules in the layer was included Equation (9) provides the isotherm:
qe = βlnKT + βlnCe
where, β = RT/b.
β represents the heat of adsorption (J/mol) and KT is the Temkin isotherm constant (L/g). The plot of qeVs lnCe produces a straight line with slope β and intercept βlnKT. b is the Temkin constant which is associated with the sorption heat (J/ mg) (Figure S6).
The adsorption is characterized by a uniform distribution of binding energies. The binding energies were 199.28, 185.86, and 146.10 J/mol for MNPs and 132.85, 94.68, and 69.43 J/mol for MCS at 308, 303, and 298 K, respectively [46,47,48].
The D-R isotherm can be used to find out the adsorption mechanism.
Dubinin-Radushkevich (D-R) isotherm (Dubinin, 1960) is:
lnqe = lnqm − βε2
ε2 and b are constants. The D-R constants qm and b were calculated from the slope and intercept of lnqeVs ε2 [37]. The qm calculated by the D-R model was 640.699 mg/g and 598.7594 mg/g for MNPs and MCS composite at 303 K, respectively (Figure S4).
Four isotherm models were studied, and the significant parameters are shown in Table S2.

3.3. Thermodynamic Studies

To investigate the spontaneity, feasibility, and endo/exothermic nature, thermodynamic data plays an important role. The effect of solution temperature (20–50 °C) on Cd (II) ion adsorption was investigated under optimal conditions. Standard entropy (S), enthalpy (H), and Gibbs free energy (G) were calculated as thermodynamic parameters. The Van’t Hoff equation was used to calculate these thermodynamic parameters.
K d = q e ( w / v ) C e
lnKd = ΔS0/R − ΔH0/RT
ΔG0 = −RT lnKd
where ΔG0 is the Gibbs free energy, ΔS0 is the entropy, and ∆H0 is the enthalpy, Kd is the distribution coefficient for the adsorption process.
Equation (11) can be used to calculate the value of Kd after the values of qe and Ce have been determined experimentally. Using the value of Kd in Equation (12), the values of enthalpy and entropy can be calculated. Using Equation (12), the slope and intercept of the plot ln (Kd) vs. 1/T in Figure 10 would give the values ∆H0 and ∆S0, respectively.
From Table S3, the positive value of ΔH shows that the adsorption process is endothermic, which confirms the chemisorption nature of adsorption. The negative value of ΔG at all three temperatures for MNPs and at 303 K for MCS composite indicated that the spontaneity of the process was favored at these temperatures. The positive value of ΔS showed a rise in randomness during the adsorption of Cd (II).

3.4. Reusability of Adsorbent

Metal ion desorption from the sorbent and adsorbent regeneration are critical challenges in terms of adsorbent reusability. The major goals of the regeneration process are to recover useful components from the adsorbed phase and to restore the sorption capability of the exhausted material [49]. Figure 11 displays the results of seven adsorption-regeneration cycles (Figure 11a,b). Using various agents including HCl (0.05 M, 0.1 M), and NaOH (0.05 M, 0.1 M), the batch technique was used to evaluate the desorption of the sorbed cadmium from Fe3O4 and MCS composite. It was discovered that 0.1 M HCl produced the greatest amount of Cd (II) desorption (98%) while using 0.1 M NaOH, the desorption efficiency was observed to be 84%. Adsorption, as well as desorption cycles, were repeated seven times utilizing the same adsorbents to assess the adsorbent’s reusability. After each cycle of adsorption, the solid portion of the adsorbent was centrifuged and further mixed in DI water. This mixture was then agitated for about 1/2 h. The remaining suspension was once again used for a different batch experiment. Seven rounds of a similar process were performed. The removal capacity of the regenerated sorbent gradually reduced in comparison to the original adsorbent. As a result, MCS was an effective reusable adsorbent that could be used to recover Cd (II) ions from aqueous medium. The percentage removal of Cd (II) decreased from the first to seventh cycle (i) 99.99% to 80.44% for MCS, (ii) 95.1% to 42% for MNPs. This demonstrates that the MCS may be reused at least seven times while retaining good sorption efficiency. As a result, it was possible to conclude that magnetic chitosan composite can be employed as an efficient material in the removal process.

3.5. Comparative Studies

MnFe2O4/CS microspheres were prepared by coating chitosan on MnFe2O4. Maximum sorption capacity was observed to be 60.6 mg/g for Cd (II) removal. The experiments showed that the composite could maintain sorption capacities after three cycles of adsorption–regeneration [50]. A new nanobiosorbent based on methionine-glutaraldehyde Schiff’s base-modified cross-linked chitosan magnetic beads was prepared by Salehi et al. (2020) for the elimination of Cd (II). The utmost sorption capacity was observed to be 163.9 mg/g [51]. Chitosan and sodium tripolyphosphate cross-linked chitosan beads were synthesized for cadmium removal from an aqueous medium by Babakhani and Sartaj (2020). The maximum sorption capacity was observed to be 99.87 mg/g [52]. Attapulgite/CoFe2O4@SiO2-chitosan/EDTA was synthesized by solvothermal and sol-gel techniques. The maximum sorption capacity was observed to be 127.79 mg/g [53]. Fan et al. (2017) studied the sorption of cadmium by using magnetic chitosan nanoparticles. The maximum sorption capacity was observed to be 36.42 mg/g for Cd (II). Relatively, the maximum sorption capacity for pure Fe3O4 nanoparticles was observed to be 13.04 mg/g [54]. Li et al. (2017) prepared chitosan/polyethylenimine grafted magnetic gelatin to eliminate cadmium from wastewater. According to the Cd (II) sorption results, the process can be explained by a monolayer forming on the surface of the material with 321 mg/g from the Langmuir isotherm [55]. The relative adsorption capacity of various adsorbents for Cd (II) and the relative adsorption performance of MCS-based adsorbents for heavy metal ions removal are shown in Table 2 and Table 3, respectively.

3.6. Adsorption Studies

3.6.1. Variation of Adsorbent Dose

The amount of the adsorbent used is a crucial factor in calculating sorption capacity. Studies were carried out by changing the sorbent dosage from 0.01 to 0.1 g/L, while all other factors such as contact duration, temperature, and rpm were held constant to optimize the sorbent dose for the elimination of Cd (II) ions from the aqueous medium. Adsorption capacity experienced a substantial reduction with increasing adsorbent doses. This may be explained by the fact that a greater sorbent dose makes it easier for Cd (II) ions to access active sites on the pores of magnetic chitosan composite, which enhances removal rates. More easily accessible functional groups and adsorption sites for metal ions were responsible for the increment. However, higher adsorbent dosage most certainly enhances particle interactions, like aggregation, which is due to high sorbent concentration. As a result, the adsorbent’s active surface area is significantly reduced, which lowers its capacity for adsorption. The adsorption process achieves the equilibrium point due to the overcrowding of adsorbent particles brought on by the overlapping of adsorption sites after a gradual decline in the sorption capacity of Cd (II) ions from the wastewater. The decline in the ratio of Cd (II) per mass unit of sorbent might potentially account for this [20,22]. The adsorption capacity for MCS fell from 613.75 mg/g to about 249.99 mg/g for Co(100) and from 488.7 mg/g to 124.99 mg/g for Co(50) as the sorbent dosage rose from 0.01 g/L to 0.1 g/L. Also, the adsorption capacity for MNPs fell from 363.75 mg/g to 248.75 mg/g for Co(100) and from 238.75 mg/g to 123.875 mg/g for Co(50) as the sorbent dosage rose from 0.01 g/L to 0.1 g/L, as shown in Figure 12a.

3.6.2. Variation in pH

The solution’s pH is a very important factor that strongly affects the sorption process because it influences the charge of the surface by protonation and deprotonation of the material and degree of ionization. By conducting equilibrium adsorption experiments at various pH levels, the impact of the suspending medium’s pH on Cd (II) removal was investigated. In the current study, the influence of solution pH on the sorption of Cd (II) by MNPs and MCS was examined throughout a pH range of 2–11 (Figure 12b). Temperature, dose, rpm, dosage, and other factors were all held constant. Low adsorption efficiency was found at low pH which was caused by competition between the Cd (II) ion and the H+ ion for placements on material active sites. Additionally, the concentration of adsorbent positive charges is large in acidic conditions, which results in electrostatic repulsion between the magnetic composite and the Cd (II), decreasing the efficiency of Cd (II) elimination. The removal of cadmium reduces when the pH rises from 2 to 6 because the positive charges on the adsorbent surface are reduced, repulsive forces are stronger, and the positive charges are less attractive [74]. The electrostatic attraction was strengthened as the pH increased, which progressively raised the degree of deprotonation of functional groups and improved the sorption capacity. The utmost removal was seen between pH 6.0 and 8.0 [75].

3.6.3. Variation of Contact Time

It has a pronounced effect on the elimination of adsorbate species from aqueous medium. To assess the Cd (II) adsorption behavior by MNPs and magnetic chitosan composite, the influence of time of contact between the adsorbent and adsorbate was optimized in this work by altering the contact duration from 5 to 50 min. By separating the supernatant at different times, the effect of contact time on sorption capacity was measured. The adsorption rate was quick in the beginning, as seen in Figure 12c. The initial strong absorption of Cd (II) ion was owing to high adsorbate and adsorbent interactions with low solute-solute interaction and more adsorption site accessibility. As the time increased, equilibrium was attained in the adsorption as more and more active sites were occupied. However, after equilibrium was reached, mass transfer diminished and the repulsion of the adsorbate molecule on the surface and inside the solution increased.
It took around 50 min to achieve equilibrium for uptake of Cd (II) [76]. The increment in time can offer plenty of chances for Cd(II) to adsorb to the composite surface. Figure 12c shows that on raising the contact time from 5 to 50 min, the sorption capacity increased from 81.458 to 416.65, 60.625 to 208.33, 39.79 to 414.79, and 27.29 to 206.458 mg/g for MCS (100 mg/L), MCS (50 mg/L), MNPs (100 mg/L), and MNPs (50 mg/L).
The results demonstrated that the adsorbents showed fast removal of Cd (II) within 50 min which is faster when we compare with the reported literature of other adsorbents employed for the Cd (II) removal [77].

3.6.4. Variation of Concentration

The elimination of Cd (II) by adsorption on MNPs and magnetic composite (MSC) was examined in relation to the effect of initial concentration. It appears that the metal ion’s initial concentration is significant and influences the sorption capacity. Because the initial concentration of Cd (II) may offer the driving force required to transfer a mass of Cd (II) between the water phase and the adsorbent surface, the concentration of Cd (II) can also impact the performance of the sorption process.
With the rise in the concentration of Cd (II), there was a decrease in % removal. The experiments were performed at different initial concentrations varying from 10 to 100 mg/L. The graph (Figure 12d) illustrates that with an increase in the concentration of Cd (II) ions, there is a decrease in Cd (II) removal % and a rise in sorption capacity. The presence of active surfaces and sites in the MNPs and MCS composite structure for the assignment of Cd (II) can be related to improving efficiency at low initial concentrations of Cd (II). Additionally, when the concentration of cadmium rises, the %-age of Cd (II) removal decreases, which is explained by the saturation of the MSC’s active sites above a specific level of Cd (II), at which point the system enters equilibrium and no further Cd (II) adsorption takes place [22].
The % removal of Cd (II) by adsorption onto the adsorbent decreases from 99.96 to 97.09, 99.99 to 97.25, and 99.999 to 97.45 for MCS (303 K, 313 K, and 323 K) and from 89.1 to 93.81, 89.9 to 94.1, and 91.0 to 94.5 for MNPs (303 K, 313 K, and 323 K), respectively (Figure 12d).

4. Adsorption Mechanism

The Cd (II) ion has an empty d orbit, and both nitrogen and oxygen atoms have lone electron pairs that can attach metal ions to create the complex via electron pair sharing. Fe3O4 enclosed in CS NPs and MNPs not only provides rich functional groups (-OH, -COOH, and -CO-NH-) as a binding site for heavy metal ions but also has a greater specific surface area to maximize functional group utilization. Functional groups on Fe3O4-loaded CS NPs and MNPs are freely available for metal ion coordination bonding.
Since negatively charged adsorbents might be quickly absorbed by positively charged groups, such as –OH2+ and –NH+3 in acidic conditions, the electrostatic attraction was a potential adsorption mechanism. The negatively charged functional groups like COO- and AOA on the surface of Fe3O4 -loaded CS NPs and the Cd (II) ion may establish an electrostatic interaction. Therefore, under the influence of both coordination bonds with the oxygen atom in Fe3O4 and electrostatic attraction, the adsorbents demonstrated a significant adsorption capacity for the Cd (II) ion.
The amines and secondary alcohol functional groups are the major sites for complexing, according to the FTIR analysis of the magnetic chitosan composite before and after Cd (II) adsorption (Figure 13). There is a slight shift in the position of various bands in the MCS after the adsorption of the metal ion, Cd (II). From the XRD data, there was a slight shift in the diffraction peaks after the adsorption of Cd (II). Also, from the magnetization results, the decrease in magnetization indirectly asserted the formation of a composite between CS and MNPs.
The oxygen of the alcohol and the nitrogen of the amino group has a pair of electrons that can add themselves to a proton by coordinated covalent bonds. While nitrogen has a larger capacity to donate its pair of electrons to a cadmium ion to form a complex through a coordinated and covalent bond, oxygen has a stronger attraction of the electron pair by the atom nucleus. This leads us to suggest that the mechanism shown in Figure 13 governs the formation of the complexes between Cd (II) and adsorbent. In this mechanism, the cadmium ion acts as a Lewis acid that may take in electron pairs due to its vacant orbitals. The amine and hydroxyl groups, on the other hand, which have non-shared electron pairs, serve as Lewis bases by giving their electron pair.

5. Conclusions

In the present study, magnetic chitosan composition was synthesized via a green route and thoroughly characterized by a variety of physical-chemical techniques. Afterward, the composite materials were successfully utilized for the adsorption of Cd (II). Elaborate studies on the kinetics and thermodynamic aspects of adsorption were carried out.For comparison, MNPs were also used as adsorbents for Cd (II). MCS showed excellent and high adsorption capacity compared to native MNPs. The adsorption capacity values of 426 mg/g for MCS and 290 mg/g for NPs were observed. The BET surface area values of MCS and NPs were found to be 178 and 105 m2/g, respectively. Both adsorbents showed second-order kinetics. The optimum process parameters include an adsorbent dose of 0.05 g/L, a contact time of 50 min, and a pH value of 6.0 with an initial Cd (II) concentration of 100 and 50 mg/L. The adsorption processes on both the adsorbents were feasible and spontaneous as evident from the thermodynamic parameters. The composite material showed good regeneration capacity with an 80.44% removal tendency up to the seventh cycle while MNPs showed low removal capacity of 42% by the seventh and the last cycle. This research provides some insights into the aspects that influence the design of adsorbents with superior performance and easy recovery for Cd (II) ion absorption. Magnetic composites have promising applications in water treatment. Its outstanding selective adsorption considerably improves the material’s adsorption efficiency, and it may be specially treated for wastewater enriched with various compounds. Furthermore, the magnetic chitosan adsorption material is easy and simple to recycle, with a high recycling rate. These benefits can lower wastewater treatment costs and increase economic efficiency. So, from the present study, it is concluded that the green synthesized MCS composite with MNPs enhanced the stability of hybrid material and also improved the removal capacity of Cd (II) with good regeneration ability and effective field study results.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/polym15214339/s1, Figure S1: Preparation of Fe3O4 nanoparticles using kaphal leaf extract, Figure S2: Preparation of chitosan/Fe3O4 composite, Figure S3: (a) Langmuir curve, (b) BET multipoint of MNPs and MCS, Figure S4: EDX images of (a) MNPs and, (b) MCS composite, Figure S5: title (a) IPD, (b) Elovich kinetic curves for MNPs and MCS composite, respectively, Figure S6: (a)Temkin, (b) D-Radsorption isotherm curves for MNPs and MCS composite respectively, Figure S7: (a) Langmuir, (b) Freundlich non-linear isotherm curves for MNPs and MCS composite respectively, Table S1: Kinetic parameters of PFO, PSO, IPD, Elovich models for Fe3O4 and MCS composite, Table S2: Adsorption parameters of four isotherm models, Table S3: Thermodynamic parameters for MNPs and magnetic chitosan composite under optimum conditions title; Table S4: Effect of coexisting ions on the adsorption of Cd(II) by MCS composite.

Author Contributions

Writing—original draft, A.Y.; review and editing, S.R.; investigation, formal analysis, data curation, N.K.J.; conceptualization, methodology, A.S.; data curation, writing—review and editing, S.J.; Supervision, M.S.; methodology, data curation, J.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research is receiving no funding.

Data Availability Statement

Data are available upon request.

Acknowledgments

We are grateful to the authorities of the Department of Chemistry, Banasthali Vidyapith (India) for the moral support and for providing the necessary facilities for the overall experimental work. Author Sapana Jadoun is grateful to Universidad de Tarapacá for its support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mahvi, A.H.; Balarak, D.; Bazrafshan, E. Remarkable Reusability of Magnetic Fe3O4-Graphene Oxide Composite: A Highly Effective Adsorbent for Cr (VI) Ions. Int. J. Environ. Anal. Chem. 2023, 103, 3501–3521. [Google Scholar] [CrossRef]
  2. Li, L.; Zou, J.; Han, Y.; Liao, Z.; Lu, P.; Nezamzadeh-Ejhieh, A.; Liu, J.; Peng, Y. Recent advances in Al (iii)/In (iii)-based MOFs for the detection of pollutants. New J. Chem. 2022, 46, 19577–19592. [Google Scholar] [CrossRef]
  3. Adelabu, I.O.; Saleh, T.A.; Garrison, T.F.; Al Hamouz, O.C.S. Synthesis of Polyamine-CNT Composites for the Removal of Toxic Cadmium Metal Ions from Wastewater. J. Mol. Liq. 2020, 297, 111827. [Google Scholar] [CrossRef]
  4. Begum, S.; Yuhana, N.Y.; Saleh, N.M.; Kamarudin, N.H.N.; Sulong, A.B. Review of Chitosan Composite as a Heavy Metal Adsorbent: Material Preparation and Properties. Carbohydr. Polym. 2021, 259, 117613. [Google Scholar] [CrossRef]
  5. Gupta, K.; Joshi, P.; Gusain, R.; Khatri, O.P. Recent Advances in Adsorptive Removal of Heavy Metal and Metalloid Ions by Metal Oxide-Based Nanomaterials. Coord. Chem. Rev. 2021, 445, 214100. [Google Scholar] [CrossRef]
  6. Mahmoud, M.E.; Mohamed, A.K. Novel Derived Pectin Hydrogel from Mandarin Peel Based Metal-Organic Frameworks Composite for Enhanced Cr (VI) and Pb (II) Ions Removal. Int. J. Biol. Macromol. 2020, 164, 920–931. [Google Scholar] [CrossRef] [PubMed]
  7. Yanrong, C.A.I.; Weili, J.; Chun, C. Research Progress on Removal of Metal Ions from Water by Marine Waste Biomass Based Adsorbent. J. Light Ind. 2022, 37, 100–110. [Google Scholar]
  8. Zhang, H.; Dang, Q.; Liu, C.; Yu, D.; Wang, Y.; Pu, X.; Liu, Y.; Liang, Y.; Cha, D. Fabrication of Methyl Acrylate and Tetraethylenepentamine Grafted Magnetic Chitosan Microparticles for Capture of Cd (II) from Aqueous Solutions. J. Hazard. Mater. 2019, 366, 346–357. [Google Scholar] [CrossRef]
  9. Imran, M.; Khan, Z.U.H.; Iqbal, J.; Shah, N.S.; Muzammil, S.; Ali, S.; Muhammad, N.; Aziz, A.; Murtaza, B.; Naeem, M.A. Potential of Siltstone and Its Composites with Biochar and Magnetite Nanoparticles for the Removal of Cadmium from Contaminated Aqueous Solutions: Batch and Column Scale Studies. Environ. Pollut. 2020, 259, 113938. [Google Scholar] [CrossRef]
  10. Karimi-Maleh, H.; Ayati, A.; Davoodi, R.; Tanhaei, B.; Karimi, F.; Malekmohammadi, S.; Orooji, Y.; Fu, L.; Sillanpää, M. Recent Advances in Using of Chitosan-Based Adsorbents for Removal of Pharmaceutical Contaminants: A Review. J. Clean. Prod. 2021, 291, 125880. [Google Scholar] [CrossRef]
  11. Ahamad, T.; Naushad, M.; Alshehri, S.M. Fabrication of Magnetic Polymeric Resin for the Removal of Toxic Metals from Aqueous Medium: Kinetics and Adsorption Mechanisms. J. Water Process Eng. 2020, 36, 101284. [Google Scholar] [CrossRef]
  12. Zheng, M.; Chen, J.; Zhang, L.; Cheng, Y.; Lu, C.; Liu, Y.; Singh, A.; Trivedi, M.; Kumar, A.; Liu, J. Metal organic frameworks as efficient adsorbents for drugs from wastewater. Mater. Today Commun. 2022, 31, 103514. [Google Scholar] [CrossRef]
  13. Doshi, B.; Ayati, A.; Tanhaei, B.; Repo, E.; Sillanpää, M. Partially Carboxymethylated and Partially Cross-Linked Surface of Chitosan versus the Adsorptive Removal of Dyes and Divalent Metal Ions. Carbohydr. Polym. 2018, 197, 586–597. [Google Scholar] [CrossRef]
  14. Li, D.; Tian, X.; Wang, Z.; Guan, Z.; Li, X.; Qiao, H.; Ke, H.; Luo, L.; Wei, Q. Multifunctional Adsorbent Based on Metal-Organic Framework Modified Bacterial Cellulose/Chitosan Composite Aerogel for High Efficient Removal of Heavy Metal Ion and Organic Pollutant. Chem. Eng. J. 2020, 383, 123127. [Google Scholar] [CrossRef]
  15. Karimi, F.; Ayati, A.; Tanhaei, B.; Sanati, A.L.; Afshar, S.; Kardan, A.; Dabirifar, Z.; Karaman, C. Removal of Metal Ions Using a New Magnetic Chitosan Nano-Bio-Adsorbent; A Powerful Approach in Water Treatment. Environ. Res. 2022, 203, 111753. [Google Scholar] [CrossRef]
  16. Chen, B.; Long, F.; Chen, S.; Cao, Y.; Pan, X. Magnetic Chitosan Biopolymer as a Versatile Adsorbent for Simultaneous and Synergistic Removal of Different Sorts of Dyestuffs from Simulated Wastewater. Chem. Eng. J. 2020, 385, 123926. [Google Scholar] [CrossRef]
  17. Liu, J.; Chen, Y.; Han, T.; Cheng, M.; Zhang, W.; Long, J.; Fu, X. A Biomimetic SiO2@Chitosan Composite as Highly-Efficient Adsorbent for Removing Heavy Metal Ions in Drinking Water. Chemosphere 2019, 214, 738–742. [Google Scholar] [CrossRef]
  18. Chen, B.; Zhao, H.; Chen, S.; Long, F.; Huang, B.; Yang, B.; Pan, X. A Magnetically Recyclable Chitosan Composite Adsorbent Functionalized with EDTA for Simultaneous Capture of Anionic Dye and Heavy Metals in Complex Wastewater. Chem. Eng. J. 2019, 356, 69–80. [Google Scholar] [CrossRef]
  19. Huang, Y.; Hu, C.; An, Y.; Xiong, Z.; Hu, X.; Zhang, G.; Zheng, H. Magnetic Phosphorylated Chitosan Composite as a Novel Adsorbent for Highly Effective and Selective Capture of Lead from Aqueous Solution. J. Hazard. Mater. 2021, 405, 124195. [Google Scholar] [CrossRef]
  20. Shaumbwa, V.R.; Liu, D.; Archer, B.; Li, J.; Su, F. Preparation and Application of Magnetic Chitosan in Environmental Remediation and Other Fields: A Review. J. Appl. Polym. Sci. 2021, 138, 51241. [Google Scholar] [CrossRef]
  21. Mohseni-Bandpi, A.; Kakavandi, B.; Kalantary, R.R.; Azari, A.; Keramati, A. Development of a Novel Magnetite–Chitosan Composite for the Removal of Fluoride from Drinking Water: Adsorption Modeling and Optimization. RSC Adv. 2015, 5, 73279–73289. [Google Scholar] [CrossRef]
  22. Wu, K.; Su, D.; Liu, J.; Saha, R.; Wang, J.-P. Magnetic Nanoparticles in Nanomedicine: A Review of Recent Advances. Nanotechnology 2019, 30, 502003. [Google Scholar] [CrossRef] [PubMed]
  23. Kabra, A.; Martins, N.; Sharma, R.; Kabra, R.; Baghel, U.S. Myrica Esculenta Buch.-Ham. Ex D. Don: A Natural Source for Health Promotion and Disease Prevention. Plants 2019, 8, 149. [Google Scholar] [CrossRef] [PubMed]
  24. de Jesús Ruíz-Baltazar, Á.; Reyes-López, S.Y.; de Lourdes Mondragón-Sánchez, M.; Robles-Cortés, A.I.; Pérez, R. Eco-Friendly Synthesis of Fe3O4 Nanoparticles: Evaluation of Their Catalytic Activity in Methylene Blue Degradation by Kinetic Adsorption Models. Results Phys. 2019, 12, 989–995. [Google Scholar] [CrossRef]
  25. Appu, M.; Lian, Z.; Zhao, D.; Huang, J. Biosynthesis of Chitosan-Coated Iron Oxide (Fe3O4) Hybrid Nanocomposites from Leaf Extracts of Brassica oleracea L. and Study on Their Antibacterial Potentials. 3 Biotech 2021, 11, 271. [Google Scholar] [CrossRef]
  26. Rai, P.; Usman, U.L.; Banerjee, S. Sequestration of Cadmium Ions from Aqueous Phase Using Magnetic Chitosan Composite. Mater. Today Proc. 2022, 49, 3375–3383. [Google Scholar] [CrossRef]
  27. Jaafari, J.; Barzanouni, H.; Mazloomi, S.; Farahani, N.A.A.; Sharafi, K.; Soleimani, P.; Haghighat, G.A. Effective Adsorptive Removal of Reactive Dyes by Magnetic Chitosan Nanoparticles: Kinetic, Isothermal Studies and Response Surface Methodology. Int. J. Biol. Macromol. 2020, 164, 344–355. [Google Scholar] [CrossRef]
  28. Bouafia, A.; Laouini, S.E.; Khelef, A.; Tedjani, M.L.; Guemari, F. Effect of Ferric Chloride Concentration on the Type of Magnetite (Fe3O4) Nanoparticles Biosynthesized by Aqueous Leaves Extract of Artemisia and Assessment of Their Antioxidant Activities. J. Clust. Sci. 2021, 32, 1033–1041. [Google Scholar] [CrossRef]
  29. Varma, R.; Vasudevan, S. Extraction, Characterization, and Antimicrobial Activity of Chitosan from Horse Mussel Modiolus Modiolus. ACS Omega 2020, 5, 20224–20230. [Google Scholar] [CrossRef]
  30. Subedi, N.; Lähde, A.; Abu-Danso, E.; Iqbal, J.; Bhatnagar, A. A Comparative Study of Magnetic Chitosan (Chi@Fe3O4) and Graphene Oxide Modified Magnetic Chitosan (Chi@Fe3O4GO) Nanocomposites for Efficient Removal of Cr (VI) from Water. Int. J. Biol. Macromol. 2019, 137, 948–959. [Google Scholar] [CrossRef]
  31. Nhiem, N.X.; Van Kiem, P.; Van Minh, C.; Tai, B.H.; Cuong, N.X.; Thu, V.K.; Kim, Y.H. A new monoterpenoid glycoside from Myrica esculenta and the inhibition of angiotensin I-converting enzyme. Chem. Pharm. Bull. 2010, 58, 1408–1410. [Google Scholar] [CrossRef] [PubMed]
  32. Hu, D.; Lian, Z.; Xian, H.; Jiang, R.; Wang, N.; Weng, Y.; Peng, X.; Wang, S.; Ouyang, X. Adsorption of Pb (II) from Aqueous Solution by Polyacrylic Acid Grafted Magnetic Chitosan Nanocomposite. Int. J. Biol. Macromol. 2020, 154, 1537–1547. [Google Scholar] [CrossRef]
  33. Zhu, A.; Yuan, L.; Liao, T. Suspension of Fe3O4 Nanoparticles Stabilized by Chitosan and O-Carboxymethylchitosan. Int. J. Pharm. 2008, 350, 361–368. [Google Scholar] [CrossRef]
  34. Jiang, Z.; Li, N.; Li, P.-Y.; Liu, B.; Lai, H.-J.; Jin, T. One-Step Preparation of Chitosan-Based Magnetic Adsorbent and Its Application to the Adsorption of Inorganic Arsenic in Water. Molecules 2021, 26, 1785. [Google Scholar] [CrossRef]
  35. Peralta, M.E.; Nisticò, R.; Franzoso, F.; Magnacca, G.; Fernandez, L.; Parolo, M.E.; León, E.G.; Carlos, L. Highly Efficient Removal of Heavy Metals from Waters by Magnetic Chitosan-Based Composite. Adsorption 2019, 25, 1337–1347. [Google Scholar] [CrossRef]
  36. Liu, Y.; Shen, L. From Langmuir Kinetics to First-and Second-Order Rate Equations for Adsorption. Langmuir 2008, 24, 11625–11630. [Google Scholar] [CrossRef] [PubMed]
  37. Sarojini, G.; Venkateshbabu, S.; Rajasimman, M. Facile Synthesis and Characterization of Polypyrrole—Iron Oxide—Seaweed (PPy-Fe3O4-SW) Nanocomposite and Its Exploration for Adsorptive Removal of Pb(II) from Heavy Metal Bearing Water. Chemosphere 2021, 278, 130400. [Google Scholar] [CrossRef] [PubMed]
  38. Peers, A.M. Elovich Adsorption Kinetics and the Heterogeneous Surface. J. Catal. 1965, 4, 499–503. [Google Scholar] [CrossRef]
  39. Revellame, E.D.; Fortela, D.L.; Sharp, W.; Hernandez, R.; Zappi, M.E. Adsorption Kinetic Modeling Using Pseudo-First Order and Pseudo-Second Order Rate Laws: A Review. Clean. Eng. Technol. 2020, 1, 100032. [Google Scholar] [CrossRef]
  40. Sahoo, T.R.; Prelot, B. Chapter 7—Adsorption Processes for the Removal of Contaminants from Wastewater: The Perspective Role of Nanomaterials and Nanotechnology. In Micro and Nano Technologies; Bonelli, B., Freyria, F.S., Rossetti, I., Sethi, R.B.T.-N., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 161–222. ISBN 978-0-12-818489-9. [Google Scholar]
  41. Ayub, A.; Irfan, A.; Raza, Z.A.; Abbas, M.; Muhammad, A.; Ahmad, K.; Munwar, A. Development of Poly (1-Vinylimidazole)-Chitosan Composite Sorbent under Microwave Irradiation for Enhanced Uptake of Cd (II) Ions from Aqueous Media. Polym. Bull. 2021, 79, 807–827. [Google Scholar] [CrossRef]
  42. Pavithra, S.; Thandapani, G.; Sugashini, S.; Sudha, P.N.; Alkhamis, H.H.; Alrefaei, A.F.; Almutairi, M.H. Batch Adsorption Studies on Surface Tailored Chitosan/Orange Peel Hydrogel Composite for the Removal of Cr (VI) and Cu (II) Ions from Synthetic Wastewater. Chemosphere 2021, 271, 129415. [Google Scholar] [CrossRef] [PubMed]
  43. Wang, X.; Li, X.; Peng, L.; Han, S.; Hao, C.; Jiang, C.; Wang, H.; Fan, X. Effective Removal of Heavy Metals from Water Using Porous Lignin-Based Adsorbents. Chemosphere 2021, 279, 130504. [Google Scholar] [CrossRef] [PubMed]
  44. Jadoun, S.; Fuentes, J.P.; Urbano, B.F.; Yáñez, J. Highly Efficient Remediation of Cu (II) from Water by Novel Poly(o-Phenylenediamine)/Zinc Oxide Nanohybrids: Kinetic, Equilibrium, and Thermodynamic Studies. J. Water Process Eng. 2023, 53, 103663. [Google Scholar] [CrossRef]
  45. Massoudinejad, M.; Rasoulzadeh, H.; Ghaderpoori, M. Magnetic Chitosan Nanocomposite: Fabrication, Properties, and Optimization for Adsorptive Removal of Crystal Violet from Aqueous Solutions. Carbohydr. Polym. 2019, 206, 844–853. [Google Scholar] [CrossRef] [PubMed]
  46. Sharma, P.; Yadav, V.; Kumari, S.; Ghosh, D.; Rawat, P.; Vij, A.; Srivastava, C.; Saini, S.; Sharma, V.; Hassan, M.I. Deciphering the Potent Application of Nanobentonite and α-Fe2O3/Bentonite Nanocomposite in Dye Removal: Revisiting the Insights of Adsorption Mechanism. Appl. Nanosci. 2023, 13, 883–897. [Google Scholar] [CrossRef]
  47. Inyinbor, A.A.; Adekola, F.A.; Olatunji, G.A. Kinetics and Isothermal Modeling of Liquid Phase Adsorption of Rhodamine B onto Urea Modified Raphia Hookerie Epicarp. Appl. Water Sci. 2017, 7, 3257–3266. [Google Scholar] [CrossRef]
  48. Togue Kamga, F. Modeling Adsorption Mechanism of Paraquat onto Ayous (Triplochiton scleroxylon) Wood Sawdust. Appl. Water Sci. 2019, 9, 1. [Google Scholar] [CrossRef]
  49. Manzoor, K.; Ahmad, M.; Ahmad, S.; Ikram, S. Removal of Pb (Ii) and Cd (Ii) from Wastewater Using Arginine Cross-Linked Chitosan–Carboxymethyl Cellulose Beads as Green Adsorbent. RSC Adv. 2019, 9, 7890–7902. [Google Scholar] [CrossRef]
  50. Zhang, Y.; Wang, Y.; Zhang, Z.; Cui, W.; Zhang, X.; Wang, S. Removing Copper and Cadmium from Water and Sediment by Magnetic Microspheres-MnFe2O4/Chitosan Prepared by Waste Shrimp Shells. J. Environ. Chem. Eng. 2021, 9, 104647. [Google Scholar] [CrossRef]
  51. Salehi, N.; Moghimi, A.; Shahbazi, H. Preparation of Cross-Linked Magnetic Chitosan with Methionine-Glutaraldehyde for Removal of Heavy Metals from Aqueous Solutions. Int. J. Environ. Anal. Chem. 2022, 102, 2305–2321. [Google Scholar] [CrossRef]
  52. Babakhani, A.; Sartaj, M. Removal of Cadmium (II) from Aqueous Solution Using Tripolyphosphate Cross-Linked Chitosan. J. Environ. Chem. Eng. 2020, 8, 103842. [Google Scholar] [CrossRef]
  53. Wang, Y.; Zhou, R.; Wang, C.; Zhou, G.; Hua, C.; Cao, Y.; Song, Z. Novel Environmental-Friendly Nano-Composite Magnetic Attapulgite Functionalized by Chitosan and EDTA for Cadmium (II) Removal. J. Alloys Compd. 2020, 817, 153286. [Google Scholar] [CrossRef]
  54. Fan, H.-L.; Zhou, S.-F.; Jiao, W.-Z.; Qi, G.-S.; Liu, Y.-Z. Removal of Heavy Metal Ions by Magnetic Chitosan Nanoparticles Prepared Continuously via High-Gravity Reactive Precipitation Method. Carbohydr. Polym. 2017, 174, 1192–1200. [Google Scholar] [CrossRef] [PubMed]
  55. Li, J.; Tao, Y.; Chen, S.; Li, H.; Chen, P.; Wei, M.; Wang, H.; Li, K.; Mazzeo, M.; Duan, Y. A Flexible Plasma-Treated Silver-Nanowire Electrode for Organic Light-Emitting Devices. Sci. Rep. 2017, 7, 16468. [Google Scholar] [CrossRef] [PubMed]
  56. Ahamad, T.; Naushad, M.; Eldesoky, G.E.; Alqadami, A.A.; Khan, A. Synthesis and Characterization of Egg-Albumen-Formaldehyde Based Magnetic Polymeric Resin (MPR): Highly Efficient Adsorbent for Cd (II) Ion Removal from Aqueous Medium. J. Mol. Liq. 2019, 286, 110951. [Google Scholar] [CrossRef]
  57. Alosaimi, E.H.; Alsohaimi, I.H.; Dahan, T.E.; Chen, Q.; Melhi, S. Adsorptive Performance of Tetracarboxylic Acid-Modified Magnetic Silica Nanocomposite for Recoverable Efficient Removal of Toxic Cd (II) from Aqueous Environment: Equilibrium, Isotherm, and Reusability Studies. J. Mol. Liq. 2021, 334, 116069. [Google Scholar] [CrossRef]
  58. Mahmoud, M.E.; Amira, M.F.; Seleim, S.M.; Mohamed, A.K. Amino-Decorated Magnetic Metal-Organic Framework as a Potential Novel Platform for Selective Removal of Chromium (Vl), Cadmium (II) and Lead (II). J. Hazard. Mater. 2020, 381, 120979. [Google Scholar] [CrossRef]
  59. Wu, Z.; Deng, W.; Zhou, W.; Luo, J. Novel Magnetic Polysaccharide/Graphene Oxide@Fe3O4 Gel Beads for Adsorbing Heavy Metal Ions. Carbohydr. Polym. 2019, 216, 119–128. [Google Scholar] [CrossRef]
  60. Wang, L.; Hu, D.; Kong, X.; Liu, J.; Li, X.; Zhou, K.; Zhao, H.; Zhou, C. Anionic Polypeptide Poly (γ-Glutamic Acid)-Functionalized Magnetic Fe3O4-GO-(o-MWCNTs) Hybrid Nanocomposite for High-Efficiency Removal of Cd (II), Cu (II) and Ni (II) Heavy Metal Ions. Chem. Eng. J. 2018, 346, 38–49. [Google Scholar] [CrossRef]
  61. Liu, G.; Liao, L.; Dai, Z.; Qi, Q.; Wu, J.; Ma, L.Q.; Tang, C.; Xu, J. Organic Adsorbents Modified with Citric Acid and Fe3O4 Enhance the Removal of Cd and Pb in Contaminated Solutions. Chem. Eng. J. 2020, 395, 125108. [Google Scholar] [CrossRef]
  62. Naushad, M.; Ahamad, T.; Al-Ghanim, K.A.; Ala’a, H.; Eldesoky, G.E.; Khan, A.A. A Highly Porous Nanocomposite (Fe3O4@ BFR) for the Removal of Toxic Cd (II) Ions from Aqueous Environment: Adsorption Modelling and Regeneration Study. Compos. Part B Eng. 2019, 172, 179–185. [Google Scholar] [CrossRef]
  63. Behbahani, E.S.; Dashtian, K.; Ghaedi, M. Fe3O4-FeMoS4: Promise Magnetite LDH-Based Adsorbent for Simultaneous Removal of Pb (II), Cd (II), and Cu (II) Heavy Metal Ions. J. Hazard. Mater. 2021, 410, 124560. [Google Scholar] [CrossRef] [PubMed]
  64. Sun, C.; Xie, Y.; Ren, X.; Song, G.; Alsaedi, A.; Hayat, T.; Chen, C. Efficient Removal of Cd (II) by Core-Shell Fe3O4@ Polydopamine Microspheres from Aqueous Solution. J. Mol. Liq. 2019, 295, 111724. [Google Scholar] [CrossRef]
  65. Liu, C.; Jiang, X.; Wang, X.; Wang, Q.; Li, L.; Zhang, F.; Liang, W. Magnetic Polyphenol Nanocomposite of Fe3O4/SiO2/PP for Cd (II) Adsorption from Aqueous Solution. Environ. Technol. 2022, 43, 935–948. [Google Scholar] [CrossRef] [PubMed]
  66. Li, Y.; Dong, X.; Zhao, L. Application of Magnetic Chitosan Nanocomposites Modified by Graphene Oxide and Polyethyleneimine for Removal of Toxic Heavy Metals and Dyes from Water. Int. J. Biol. Macromol. 2021, 192, 118–125. [Google Scholar] [CrossRef]
  67. Abou El-Reash, Y.G. Magnetic Chitosan Modified with Cysteine-Glutaraldehyde as Adsorbent for Removal of Heavy Metals from Water. J. Environ. Chem. Eng. 2016, 4, 3835–3847. [Google Scholar] [CrossRef]
  68. Li, B.; Zhou, F.; Huang, K.; Wang, Y.; Mei, S.; Zhou, Y.; Jing, T. Environmentally Friendly Chitosan/PEI-Grafted Magnetic Gelatin for the Highly Effective Removal of Heavy Metals from Drinking Water. Sci. Rep. 2017, 7, 43082. [Google Scholar] [CrossRef]
  69. Zeng, X.; Zhang, G.; Wen, J.; Li, X.; Zhu, J.; Wu, Z. Simultaneous Removal of Aqueous Same Ionic Type Heavy Metals and Dyes by a Magnetic Chitosan/Polyethyleneimine Embedded Hydrophobic Sodium Alginate Composite: Performance, Interaction and Mechanism. Chemosphere 2023, 318, 137869. [Google Scholar] [CrossRef]
  70. Le, T.T.N.; Le, V.T.; Dao, M.U.; Nguyen, Q.V.; Vu, T.T.; Nguyen, M.H.; Tran, D.L.; Le, H.S. Preparation of Magnetic Graphene Oxide/Chitosan Composite Beads for Effective Removal of Heavy Metals and Dyes from Aqueous Solutions. Chem. Eng. Commun. 2019, 206, 1337–1352. [Google Scholar] [CrossRef]
  71. Fan, L.; Luo, C.; Sun, M.; Li, X.; Qiu, H. Highly Selective Adsorption of Lead Ions by Water-Dispersible Magnetic Chitosan/Graphene Oxide Composites. Colloids Surf. B Biointerfaces 2013, 103, 523–529. [Google Scholar] [CrossRef]
  72. Lv, L.; Chen, N.; Feng, C.; Zhang, J.; Li, M. Heavy Metal Ions Removal from Aqueous Solution by Xanthate-Modified Cross-Linked Magnetic Chitosan/Poly (Vinyl Alcohol) Particles. RSC Adv. 2017, 7, 27992–28000. [Google Scholar] [CrossRef]
  73. Liu, T.; Han, X.; Wang, Y.; Yan, L.; Du, B.; Wei, Q.; Wei, D. Magnetic Chitosan/Anaerobic Granular Sludge Composite: Synthesis, Characterization and Application in Heavy Metal Ions Removal. J. Colloid Interface Sci. 2017, 508, 405–414. [Google Scholar] [CrossRef] [PubMed]
  74. Peighambardoust, S.J.; Foroutan, R.; Peighambardoust, S.H.; Khatooni, H.; Ramavandi, B. Decoration of Citrus Limon Wood Carbon with Fe3O4 to Enhanced Cd2+ Removal: A Reclaimable and Magnetic Nanocomposite. Chemosphere 2021, 282, 131088. [Google Scholar] [CrossRef] [PubMed]
  75. Homayonfard, A.; Miralinaghi, M.; Shirazi, R.H.S.M.; Moniri, E. Efficient Removal of Cadmium (II) Ions from Aqueous Solution by CoFe2O4/Chitosan and NiFe2O4/Chitosan Composites as Adsorbents. Water Sci. Technol. 2018, 78, 2297–2307. [Google Scholar] [CrossRef] [PubMed]
  76. Wu, D.; Wang, Y.; Li, Y.; Wei, Q.; Hu, L.; Yan, T.; Feng, R.; Yan, L.; Du, B. Phosphorylated Chitosan/CoFe2O4 Composite for the Efficient Removal of Pb (II) and Cd (II) from Aqueous Solution: Adsorption Performance and Mechanism Studies. J. Mol. Liq. 2019, 277, 181–188. [Google Scholar] [CrossRef]
  77. Song, X.; Li, L.; Zhou, L.; Chen, P. Magnetic Thiolated/Quaternized-Chitosan Composites Design and Application for Various Heavy Metal Ions Removal, Including Cation and Anion. Chem. Eng. Res. Des. 2018, 136, 581–592. [Google Scholar] [CrossRef]
  78. Reddy, D.H.K.; Lee, S.M. Application of magnetic chitosan composites for the removal of toxic metal and dyes from aqueous solutions. Adv. Colloid Interface Sci. 2013, 201, 68–93. [Google Scholar] [CrossRef]
Figure 1. Systematic illustration of magnetic chitosan composite for Cd(II) removal.
Figure 1. Systematic illustration of magnetic chitosan composite for Cd(II) removal.
Polymers 15 04339 g001
Figure 2. Preparation of plant extract of Myrica esculenta (Kaphal).
Figure 2. Preparation of plant extract of Myrica esculenta (Kaphal).
Polymers 15 04339 g002
Figure 3. XRD spectra for (a) MNPs [before and after Cd (II) adsorption], (b) Porous MCS adsorbent [before and after Cd (II) adsorption].
Figure 3. XRD spectra for (a) MNPs [before and after Cd (II) adsorption], (b) Porous MCS adsorbent [before and after Cd (II) adsorption].
Polymers 15 04339 g003
Figure 4. N2 sorption isotherm of (a) MNPs, (b) MCS composite.
Figure 4. N2 sorption isotherm of (a) MNPs, (b) MCS composite.
Polymers 15 04339 g004
Figure 5. TIR spectra for (a) Chitosan and Fe3O4 nanoparticles, (b) MCS adsorbent [before and after Cd (II) adsorption].
Figure 5. TIR spectra for (a) Chitosan and Fe3O4 nanoparticles, (b) MCS adsorbent [before and after Cd (II) adsorption].
Polymers 15 04339 g005
Figure 6. FE-SEM image of (a) MNPs, and (b) MCS composite.
Figure 6. FE-SEM image of (a) MNPs, and (b) MCS composite.
Polymers 15 04339 g006
Figure 7. Magnetization curves of the Fe3O4, MCS composite, and MCS composite after Cd (II) adsorption.
Figure 7. Magnetization curves of the Fe3O4, MCS composite, and MCS composite after Cd (II) adsorption.
Polymers 15 04339 g007
Figure 8. (a) PFO and (b) PSO for MNPs and MCS composite.
Figure 8. (a) PFO and (b) PSO for MNPs and MCS composite.
Polymers 15 04339 g008
Figure 9. (a) Langmuir, (b) Freundlich curves for Fe3O4 and MCS composite as adsorbents for Cd (II) at different temperatures.
Figure 9. (a) Langmuir, (b) Freundlich curves for Fe3O4 and MCS composite as adsorbents for Cd (II) at different temperatures.
Polymers 15 04339 g009
Figure 10. Linear dependence of ln (Kd) on 1/T based on adsorption thermodynamics for (a) MNPs and (b) MCS at various temperatures.
Figure 10. Linear dependence of ln (Kd) on 1/T based on adsorption thermodynamics for (a) MNPs and (b) MCS at various temperatures.
Polymers 15 04339 g010
Figure 11. Regeneration studies using (a) 0.1 M HCl and 0.1 M NaOH, (b) Regeneration studies of MNPs and MCS (0.1 M HCl).
Figure 11. Regeneration studies using (a) 0.1 M HCl and 0.1 M NaOH, (b) Regeneration studies of MNPs and MCS (0.1 M HCl).
Polymers 15 04339 g011
Figure 12. Effect of (a) adsorbent dose, (b) pH, (c) contact time, (d) initial concentration for adsorption of Cd (II) on Fe3O4 and MCS composite, respectively.
Figure 12. Effect of (a) adsorbent dose, (b) pH, (c) contact time, (d) initial concentration for adsorption of Cd (II) on Fe3O4 and MCS composite, respectively.
Polymers 15 04339 g012
Figure 13. Proposed mechanism for Cd (II) capture by MCS composite [78].
Figure 13. Proposed mechanism for Cd (II) capture by MCS composite [78].
Polymers 15 04339 g013
Table 1. Surface parameters of MNPs and MCS adsorbents before cadmium adsorption.
Table 1. Surface parameters of MNPs and MCS adsorbents before cadmium adsorption.
ParametersMNPsMCS
BET specific surface area (m2/g)105173
Langmuir surface area (m2/g)201537
Average pore size (Å) 64.3649.77
Total pore volume (cc/g)0.34100.4305
Table 2. The relative adsorption capacity of various adsorbents for Cd (II).
Table 2. The relative adsorption capacity of various adsorbents for Cd (II).
S. No.MaterialsQe(mg/g)Reference
1Egg-albumen-formaldehyde-based magnetic polymeric resin149.3[56]
2Mesoporous magnetic nanocomposite158.68[57]
3Amino-decorated magnetic metal-organic framework693.0[58]
4Carboxymethyl chitosan/sodium alginate/graphene oxide@ Fe3O4 beads86.28[59]
6Poly(γ-glutamic acid) modified magnetic Fe3O4-GO-(o-MWCNTs) hybrid nanocomposite625.00[60]
7Citric acid- and Fe3O4-modified sugarcane bagasse33.2[61]
8Fe3O4@Biuret-formaldehyde pre polymeric resin92.6[62]
9Fe3O4/FeMoS4/MgAl-LDH nanocomposite140.50[63]
10Fe3O4@PDA microspheres296.4[64]
11Fe3O4/SiO2/PP30.1[65]
12Fe3O4 nanoparticles290This work
13Fe3O4/Chitosan composite 426This work
Table 3. The relative adsorption performance of MCS-based adsorbents for heavy metal ions removal.
Table 3. The relative adsorption performance of MCS-based adsorbents for heavy metal ions removal.
S. No.MaterialsHeavy MetalsQe(mg/g)Reference
1.Magnetic chitosan compositeNi(II)
Cu(II)
Pb(II)
108.9
216.8
220.9
[35]
2. Magnetic chitosan nanocomposites modified with graphene oxide and polyethyleneimine As(V)
Hg(II)
220.26
124.84
[66]
3.Chitosan magnetic beads modified with cysteine glutaraldehyde Schiff’s baseCu(II)
Cr(VI)
156.49
138.53
[67]
4.PEI-grafted magnetic gelatinPb(II)
Cd (II)
341
321
[68]
5. Magnetic chitosan/polyethyleneimine embedded hydrophobic sodium alginate compositeCr(VI)
Cu(II)
87.53
351.03
[69]
6. Magnetic graphene oxide/chitosan composite beadsNi(II)80.48[70]
7.Magnetic Fe3O4/Chitosan nanoparticlesPb(II)
Cd (II)
79.24
36.42
[54]
8.Magnetic chitosan/graphene oxide (MCGO) materialsPb(II)76.94[71]
9.Xanthate-modified cross-linked magnetic chitosan/poly(vinyl alcohol) particlesPb(II)
Cu(II)
59.855
139.797
[72]
10.Magnetic anaerobic granule sludge/chitosan compositePb(II)
Cu(II)
97.97
83.65
[73]
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

Yadav, A.; Raghav, S.; Jangid, N.K.; Srivastava, A.; Jadoun, S.; Srivastava, M.; Dwivedi, J. Myrica esculenta Leaf Extract—Assisted Green Synthesis of Porous Magnetic Chitosan Composites for Fast Removal of Cd (II) from Water: Kinetics and Thermodynamics of Adsorption. Polymers 2023, 15, 4339. https://doi.org/10.3390/polym15214339

AMA Style

Yadav A, Raghav S, Jangid NK, Srivastava A, Jadoun S, Srivastava M, Dwivedi J. Myrica esculenta Leaf Extract—Assisted Green Synthesis of Porous Magnetic Chitosan Composites for Fast Removal of Cd (II) from Water: Kinetics and Thermodynamics of Adsorption. Polymers. 2023; 15(21):4339. https://doi.org/10.3390/polym15214339

Chicago/Turabian Style

Yadav, Anjali, Sapna Raghav, Nirmala Kumari Jangid, Anamika Srivastava, Sapana Jadoun, Manish Srivastava, and Jaya Dwivedi. 2023. "Myrica esculenta Leaf Extract—Assisted Green Synthesis of Porous Magnetic Chitosan Composites for Fast Removal of Cd (II) from Water: Kinetics and Thermodynamics of Adsorption" Polymers 15, no. 21: 4339. https://doi.org/10.3390/polym15214339

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