Next Article in Journal
Advances in the Clinical Application of Histamine and Diamine Oxidase (DAO) Activity: A Review
Next Article in Special Issue
Nano-ZrO2-Catalyzed Biginelli Reaction and the Synthesis of Bioactive Dihydropyrimidinones That Targets PPAR-γ in Human Breast Cancer Cells
Previous Article in Journal
Solvothermal Synthesis of g-C3N4/TiO2 Hybrid Photocatalyst with a Broaden Activation Spectrum
Previous Article in Special Issue
Au@CdS Nanocomposites as a Visible-Light Photocatalyst for Hydrogen Generation from Tap Water
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Characterization and Investigation of Cross-Linked Chitosan/(MnFe2O4) Nanocomposite Adsorption Potential to Extract U(VI) and Th(IV)

by
Marwa Alaqarbeh
1,*,
Fawwaz Khalili
2,
Mohammed Bouachrine
3,4 and
Abdulrahman Alwarthan
5
1
National Agricultural Research Center, Al-Baqa 19381, Jordan
2
School of Science, Department of Chemistry, The University of Jordan, Amman 11942, Jordan
3
Molecular Chemistry and Natural Substances Laboratory, Faculty of Science, Moulay Ismail University of Meknes, Meknes 50000, Morocco
4
EST Khenifra, Sultan Moulay Sliman University, Benimellal 54010, Morocco
5
Department of Chemistry, College of Science, King Saud University, Riyadh 11451, Saudi Arabia
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(1), 47; https://doi.org/10.3390/catal13010047
Submission received: 5 November 2022 / Revised: 12 December 2022 / Accepted: 17 December 2022 / Published: 26 December 2022
(This article belongs to the Special Issue From Design to Application of Nanomaterials in Catalysis)

Abstract

:
A cross-linked chitosan/(MnFe2O4) CCsMFO nanocomposite was prepared using co-precipitation methods and used as a nanomaterial to extract U(VI) and Th(IV) from an aqueous solution based on adsorption phenomena. The contact time of experiments shows a rapid extraction process within 30 min by the CCsMFO nanocomposite. The solution pH acts a critical role in determining qm value, where pH 3.0 was the best pH value to extract both ions. The pseudo-second-order equilibrium model illustrated the kinetics equilibrium modal extraction process. Adsorption isotherm of U(VI) at pH 3.0 by CCsMFO nanocomposite is an endothermic process. In contrast, the adsorption isotherm of Th(IV) at pH 3.0 by CCsMFO nanocomposite is an exothermic process. The reusability of CCsMFO nanocomposite was tested using basic eluents as suitable conditions for the chemical stability of CCsMFO nanocomposite; the reusability results show promising results for the removal of U(VI) adsorbed onto CCsMFO nanocomposite with 77.27%, after 12 h by Na2CO3 as eluent. At the same time, the reusability results show good reusability for removal of U(VI) adsorbed onto CCsMFO nanocomposite with 21.82%, after 8 h by EDTA as eluent.

Graphical Abstract

1. Introduction

A natural heteropolymer produced from chitin by chemical, enzymatic methods through full or partially chitin’s alkaline deacetylation to produce a poly(β-1-4)-2-amino-2-deoxy-D-glucopyranose (Figure 1) is known as Chitosan (CS), where the deacetylation degree (%DD) plays a significant role in the physical, chemical, and biological properties of CS [1,2]. The measured pKa of chitosan gel-aqueous solution ranges 6.5 to 6.7 because it has three different chemical functional groups: primary and secondary hydroxyl groups at the C-6 and C-3 positions, and an amino group at the C-2 place (Figure 1) [3,4,5]. At a relatively low pH < 6.5, CS has a high positive charge because of amino group protonation [6,7]. Chitosan shows a high sensitivity to high temperatures which causes thermal analysis at high temperatures. In addition, the presence of oxidizable and hydrolyzable bonds in the backbone of chitosan makes it chemically sensitive to some materials. In addition, enhanced CS accessibility for microorganisms leads to biodegradation with time [2]. Despite these obstacles, and due to Cs’s chemical and physical features, it has several biological and medicinal applications as hemodialysis membrane materials, artificial skin, drug delivery, and substituting or regenerating blood/tissue interfaces [8,9,10]. In addition, it is used as an effective material in water treatment applications in coagulant [7] and membrane manufacturing [2,11]. Moreover, the advantageous characteristics of low price, abundance, nontoxicity, biocompatibility, biodegradability, hydrophilicity, ionic active sites (NH2 and OH group), and high adsorption capability that DD determined CS is used as an adsorbent for organic and dye pollutants removal [5,12]. In addition, it is used as an adsorbent for Fe(II), Cu(II), Zn(II), Cr(VI), Ni(II), Hg(II), Pb(II), Cr(III) and Cd(II) remediation due to its chelation capability [12,13,14,15,16,17]. Moreover, it is used to remove stable nuclide ions such as U(VI), Th(IV), and radionuclide Sr90, Co60, and Cs137 [18,19,20,21,22,23]. Th(IV) and U(VI) ions cause diverse health problems such as kidney toxicity, neurotoxin, immune toxin, mutagen, and carcinogen, which increase the probability of lung, pancreatic, and colorectal cancers, in addition to chronic respiratory diseases, liver damage, anemia, RBC hemolysis, fatigue syndrome, and low blood pressure [24].
Chitosan is widely used as a surface-modified material for several nanomaterials [25,26,27]. Most surface-modified materials are magnetic materials such as Fe2O4 and spinal ferrite MFe2O4 (M = Fe, Mn, Zn, Ni, Mg, and Co) [28]. Mainly, magnetic materials are fabricated using cross-linker molecules such as glyoxal, glutaraldehyde, isatin, epichlorohydrin, and formaldehyde called modified cross-linked nanocomposites [5,19,29,30,31,32,33,34,35,36,37]. The cross-linked chitosan was used for coating a magnetic nanoparticle to enhance the adsorption capacity of magnetic nanoparticles due to NH2 and OH groups which make ionic active sites. At the same time, the magnetic properties of cross-linked chitosan metal-oxide nanocomposites facilitate separating it from aqueous media by a magnet, an important feature that shortens many steps in the water purification process [29,30,31,32,33,34,35,36,37].
The present research aims to fabricate and characterize cross-linked chitosan Manganese spinel ferrite (CCsMFO) nanocomposites by FTIR, TGA, DSC, SEM, and TEM. The extraction of Th(IV) and U(VI) ions in this research is based on adsorption phenomena. The adsorption conditions were examined using batch experiments to optimize the adsorption conditions at different contact times, pH, temperatures, and initial metal-ion concentrations at constant ionic strength. To find the adsorption capacity of Th(IV) and U(VI) by (CCsMFO), the kinetic model, adsorption isotherm, and thermodynamic adsorption results were obtained based on optimization results.

2. Results and Discussion

2.1. Characterization of (CCsMFO) Nanocomposite

The FT-IR spectrum of MnFe2O4, chitosan, and (CCsMFO) are shown in (Figure 2A–C), respectively. A stretching band at 568.4 cm−1 for (Fe–O) and 649.6 cm−1 for (Mn–O) of the spinel structure of iron is shown in (Figure 2A). While a broad band from 3272.70 cm-1 (N–H stretching), 3524.75 cm−1 (O–H stretching), 2933 and 2879.92 cm−1 (C–H stretching), 1664.05 cm−1 (amide II band, C–O stretching of the acetyl group), a 1592.09 cm−1 strong band corresponding to (amide II band, N–H stretching), 1380–1320 cm−1 (asymmetrical C–H bending of the CH2 group) and 1079.35 cm−1 (O bridge stretching) of the glucosamine residue corresponding to chitosan (Figure 2B). In Figure 2C, the FT-IR spectrum shows the stretching band 568.95 cm−1 (Fe–O) and 655.95 cm−1 (Mn–O) for MnFe2O4 and the characteristic band at 3298.73 cm-1 (O–H stretching) and 2932.29 cm−1 (C–H stretching), and a 1635.72 cm−1 medium band corresponding to (C=N (imine) group), 1559.00 cm−1 (amide II band, N–H stretching) 1421.59 cm−1 (asymmetrical C–H bending of the CH2 group) and 1068.00 cm−1 (O–bridge stretching) of the glucosamine residue for CS. The IR results for (CCsMFO) showed the presence of characteristic bands corresponding to (Fe–O), (Mn–O), and (C=N), which proves that the method of preparation is proper and produced the target nanoparticles.
The diffractogram of MnFe2O4 NPs, chitosan, and (CCsMFO) is given in (Figure 3A–C, respectively). Figure 3A shows the characteristic diffraction peaks of MnFe2O4 NPs with an inverse cubic spinel structure (cubic, space group: Fd3m; JCPDS No. 73-1964) at 18.66° (110), 30.32° (220), 35.38° (311), 43.12° (400), 52.12° (422), and 57.12° (511), which is confirmed by the previous literature [38,39]. In Figure 3B, two characteristic peaks of chitosan were suggested, the first at 10.7° (020) and 20.0° (220), which were semi-crystalline due to the hydroxyl and amine groups of chitosan [6]. At the same time, Figure 3C shows the diffraction peaks of the (CCsMFO) nanocomposite which contains peaks of MnFe2O4 and chitosan at 14.84° (110), 31.36° (220), 28.24° (311), 37.12° (400), 51.82° (422), and 56.52° (511). That shows that the width of CS peaks (220) became narrower in size, indicating that the size became smaller and CS peaks (020) vanished due to imine bond forming (C=N). In addition, the peaks of chitosan and MnFe2O4 positions were changed. At the same time, this indicates MnFe2O4 nanoparticle binding to glutaraldehyde-cross linked chitosan.
The particle size of the (CCsMFO) nanocomposite was quantitatively calculated using XRD peaks from Debye–Scherrer (Equation (1)), which finds a relationship between particle size and peak broadening. The peak sizes of the (220), (311), (400), and (511) for MnFe2O4-CS obtained from this equation were found to be about 20, 42, 30, 24, and 51 nm, respectively.
D =   k λ   β   c o s ( θ )  
where k, λ, β, and θ are Sherrer constant (0.89), X-ray wavelength (nm), the peak width at half maximum, and Bragg diffraction angle, respectively.
The thermal stability of MnFe2O4 NPs, chitosan, and CCsMFO nanocomposite was studied by thermal gravity analysis as shown in Figure 4A–C for MnFe2O4 NPs, chitosan, and CCsMFO nanocomposite, respectively. The TGA curve for MnFe2O4 NPs (Figure 4A) shows no mass changes during the overheating process from rt to 1000 °C; the residual mass is 98.32% indicating that MnFe2O4 NPs are thermally stable. For the TGA curve of chitosan (Figure 4B), a 14% weight loss occurred between (100–150 °C) due to water evaporation. However, the weight stayed constant up to 248 °C. After that, a significant weight loss happened up to 750 °C (68.53% total weight loss) because of the thermal degradation of CS organic matter. The TGA curve of the CCsMFO nanocomposite (Figure 4C) shows three main weight-loss stages. The first stage illustrates a solvent loss (water evaporation) at 120 °C in CCsMFO nanocomposite. In comparison, the thermal degradation of chitosan is responsible for the second and third weight-loss phases at approximately 230 and 650 °C, respectively. The residual mass is 39.41% at 800 °C of the heating process, where, due to MnFe2O4, it remained unchanged, indicating CCsMFO nanocomposite consists of 39.41% metal oxide and 60.59% chitosan.
The morphology of CCsMFO nanocomposite was explored using SEM, TEM, and EDS. SEM micrographs (Figure 5) show the irregular particle size and shape of CCsMFO nanocomposite particles. At the same time, TEM images (Figure 6) show miscellaneous flake forms with different size particles, and the particle size was measured using the “Image-J” program and expressed in Table 1. In SEM micrographs, energy dispersive X-ray spectroscopy (EDS) determines each element’s kind and weight percent. The proportion of each element after normalization is displayed in Table 2 and Figure 7. The source of Si Kα peaks detected in EDS charts breaks in using conventional Si(Li) detectors that produced a silicon internal fluorescence peak from a Si dead layer of the Si-Li detector [40].
The electrochemical potential between a charged surface and a liquid flowing near the charged surface’s plane is known as the zeta-potential [40]. According to Figure 8, CCsMFO has a negative value (−20.20 mV), making it a good material for positive species adsorption, such as U(VI) and Th(IV).

2.2. Optimization Adsorption Protocols

2.2.1. Effect of Adsorbent Amount

A total of 5.0 mL of 50.0 ppm of U(VI) and Th(IV) solution at pH 3.0 were contacted with various masses of adsorbent (1.0–5.0 mg) for 12 h at 25 °C. Figure 9A shows that the uptake of U(VI) and Th(IV) increases as CCsMFO nanocomposite mass increases to 87% and 91%, respectively, at mass = 5.0 mg for both metal ions. The increased amount of the adsorbent substance increases with its active adsorption sites, and, thus, the rate of uptake increases [41]. Where the uptake increases as U(VI) and Th(IV) concentration increases, which may be explained by the improved ratio of total active sites to the metal ions in solution; these ions interact fully with the active site [42,43,44,45].

2.2.2. Solution pH Effect

A total of 5.0 mg of CCsMFO nanocomposite was conducted on 5.0 mL of 50.0 ppm of different pH (1.0–7.0) metal ions solution at 25 °C for 12 h. Figure 9B,C show the maximum% uptake of U(VI) at 48% and Th(IV) at 87% at pH 3.0.
As noted in Figure 9B, the maximum uptake of U(VI) is at pH = 7 ˃ 6 ˃ 3 ≥ 5 = 4 ˃ 2 ˃ 1. In the aqueous solution, uranium forms oligomeric hydrolysis species depending on the pH of the solution. Dissociation of UO2(NO3)2 uranyl nitrate in water involves several ions at pH > 3.0, but (UO2)2(OH)22+ uranyl hydroxide is the main ion as pH increases from pH = 3 to pH = 7, in addition to oligomeric and monomeric hydrolysis species (UO2)3(OH)5+, (UO2)3(OH)42+, (UO2)2OH3+, (UO2)3(OH)7-, (UO2)4(OH)7+, UO2OH+, UO2(OH)2, UO2(OH)3- and UO2(OH)42- [42,43,44,45]. In addition, as pH increases from pH = 3 to pH = 7, uranyl cation UO22+ reacts with carbonate anion CO32− to form monomeric carbonate compounds: UO2CO3, UO2(CO3)22− and UO2(CO3)34−, and oligomeric carbonate compounds: (UO2)3(CO3)66− [43]. In the chemistry of uranium ions in an aqueous solution, the most soluble species of total U(VI) is UO22+. In contrast, the other species UO2(OH)+, (UO2)2(OH)22+, (UO2)3(OH)53+, (UO2)2(OH)2, UO2(CO3)22−, UO2(CO3)34−, and (UO2)3(CO3)66− enhance the precipitation as a yellow precipitate on the surface of the adsorbent and compete for sorption efficiency of UO22+ on the adsorption active sites, so the suitable pH for the adsorption of uranyl ion on CCsMFO nanocomposite is 3.0. At the same time, Figure 9B notes that the maximum% uptake of Th(IV) is at pH 3.0, which corresponds to the hydrolysis of Thorium ion in an aqueous solution which depends on the pH of the solution. As pH increases from pH 3 to pH 7, Th(IV) salts hydrolyze to produce monomeric hydrolysis species [ThOH]3+ and [Th(OH)2]2+. Monomeric species have a strong tendency to polymerize [ThOH(H2O)n]3+, where n = 7 or 8, and dimerize to [Th2(OH)2(H2O)12]6+, which makes the slightly soluble white precipitate compete with a soluble Thorium species to adsorb on the adsorbent surface [42,43,44,45]. As noted in (Figure 9C) at pH 3.0, Th4+ ions are a predominant soluble species over other thorium ions species that can easily adsorb on the surface of CCsMFO nanocomposite.

2.2.3. Contact Time Effect

Using a solution of 5.0 mL of 50.0 ppm U(VI) and Th(IV) at pH 3.0 and 25.0 °C, the contact time of adsorption by CCsMFO nanocomposite was performed at regular periods (0.5, 1, 3, 6, 8, and 12 h). Figure 9D expresses that the maximum adsorption capacity (qm) of U(VI) and Th(IV) onto CCsMFO nanocomposite increases as contact time increases. The value of qm for U(VI) required 3 h at pH values of 3.0, whereas the value of qm for Th(IV) needed 1 h.

2.2.4. Metal Ion Concentration Effect

A total of 5.0 mL of solutions of different U(VI) and Th(IV) concentrations in the range (10.0–100.0) ppm at pH 3.0 contacted 5.0 mg of CCsMFO nanocomposite for 12 h. Figure 9E shows that the uptake reaches the maximum at 50.0 ppm. This may be illustrated by referring to the linear driving-force law accumulation rate of adsorbent (U(VI) and Th(IV)) on adsorbate (CCsMFO nanocomposite) as a concentration gradient function [45].

2.2.5. Adsorption Kinetics Modeling Studies

Most adsorption reactions are controlled by several successive steps, including (a) resistance to film diffusion, (b) resistance to intraparticle diffusion, and (c) the proper sorption reaction rate [46,47]. Since the first step is excluded by sufficiently shaking the solution, the rate-determining step is one of the other two steps. Two kinetic models are used to postulate time-dependent adsorption models: pseudo-first-order and pseudo-second-order. The proposed model should explain the sorption kinetics modal of U(VI) and Th(IV) on CCsMFO nanocomposite, which is comprised of two phases: a rapid adsorption-driven early phase which greatly influences equilibrium sorbent uptake (first step) and a second, slower phase which adds to the minimal amount of metal-ion sorption [46,47]. The slower step, known as the rate-determining step of metal-ion sorption on sorbent surfaces, involves interactions between bulk solution and sorbent surface [46,47].
Equations (2) and (3) describe liner pseudo-first-order and pseudo-second-order adsorption, respectively.
ln (qₑqt) = ln (qₑ) − kt
t q e = 1 k 2 q e 2 + 1 q e   t
Figure 10A,B expresses pseudo-first-order and pseudo-second-order kinetic models results. At the same time, the values of (qe) calculated, (qe) experimental, and R2 correlation coefficients are tabulated in Table 3, where the values of R2, qe calculated, and qe experimental fit the pseudo-second-order kinetic model more than the pseudo-first-order model. This indicates that in the rate-determining step of the adsorption of U(VI) and Th(IV) on CCsMFO, the nanocomposite mechanism occurs by chemisorption [48,49,50]. The adsorption behavior involves the valency forces through sharing electrons between the U(VI) and Th(IV) ions and the adsorbent active sites of chitosan on CCsMFO as amine, carboxymethyl, and hydroxyl groups [48,49,50]. The value of k2 (Table 2) shows the uptake of U(VI) by CCsMFO nanocomposite at pH 3.0 (0.08 g·mg−1·h−1) and k2 for uptake of Th(IV) by CCsMFO nanocomposite at pH 3.0, (1.07 g·mg−1·h−1) achieves equilibrium faster than the uptake of U(VI) because the hydrolysis of Th(IV) and U(VI), as mentioned in Section 2.2, drives the hydration radius of thorium ions less than uranyl ions species; thus, the adsorption of thorium ions is easier than uranyl ions [51].

2.2.6. Temperature’s Effect and Adsorption Isotherms Modeling

The temperature effect in this work was investigated using pH, contact time, and metal-ions-concentration optimized conditions by variation in temperature at 25, 35, and 45 °C. The results of temperature variations (Figure 11A) show the adsorption isotherm of U(VI) by CCsMFO nanocomposite belonging to the S-type adsorption isotherm, which explains the adsorption process experienced by cooperative adsorption mechanisms, where adsorbate interaction at the surface of the adsorbent is more vital than adsorbate interaction in bulk by creating a cluster of the multilayers of the adsorbate at the surface of the sorbent [49,50]. Meanwhile, Figure 11A shows the adsorption isotherm of Th(IV) by CCsMFO nanocomposite belonging to the H-type adsorption isotherm, which suggests extremely high affinity of (adsorbate) Th(IV) for (adsorbent) CCsMFO nanocomposite [51,52,53].
The adsorption isotherms model is often described using linear equations of the Langmuir, Freundlich, and Dubinin–Radushkevich (D-R) isotherm models (Equations (4), (6) and (7), respectively). Equation (4) is the linear equation form of the Langmuir (II) model, where the intercept is used to find qm (mg/g), and the obtained value is used with slop to find KL (L/mg). The calculated value of qm and KL are used to find the value of RL (Equation (5)), which expresses the adsorption process to be either unfavorable if RL > 1), favorable if 0 < RL < 1, linear if RL = 1, and irreversible if RL = 0 [51,52,53].
1 q e = ( 1 q m K L ) 1 C e + 1 q m  
R L = 1 1 + K L C o
Equation (6) is the linear equation form of the Freundlich modal, where n (heterogeneity factor) was obtained from slop, and the intercept was used to find the adsorption capacity value (Kf). A value of (n) above one reflects normal sorption, but if it is less than one, this reflects that the sorption process undergoes a cooperative adsorptions process [54,55].
l o g   q e = l o g   K f   + 1 n l o g   C e  
Equation (7) is the linear (D-R) model that explains the adsorption process’s physical and chemical characteristics. It helps to calculate potential binding energy E (kJ/mol), which describes the transferring of one mol of adsorbate from the solution to the sorbent surface [56]. Polanyi potential (ε) is from Equation (11), where R is the gas constant (kJ K−1 mol−1), and T is the temperature (K).
l n ( q e ) = l n ( q m ) β ε 2
ε = R   T l n ( 1 + 1 C e )
The constant (mol2/kJ2) (β) value is obtained from Equation (7) and Equation (9) was used to obtain E.
E = 1 2 β      
The results of adsorption isotherms for U(VI) and Th(IV) uptake by CCsMFO nanocomposite are expressed in Figure 12A–C and Figure 13A–C and the parameters qm, n, linear regression (R2), and E listed in Table 4 were used to explain the adsorption process.
As shown in Table 4 and Figure 12A–C, Langmuir (II) and Freundlich isotherms, the model’s R2 values are > 0.89, signifying that the U(VI) adsorption by CCsMFO nanocomposite can be described using the Langmuir (II) and Freundlich models. RL values were calculated and found to be 0 < RL < 1, indicating a favorable adsorption of U(VI) onto CCsMFO nanocomposite by the formation of a monolayer at the surface of the sorbent [56,57]. The values of qm are 500.00, 250.00, and 111.11 (mg/g) at T = 25, 35, 45 °C, respectively. The Langmuir (qm) values of U(VI) uptake onto CCsMFO nanocomposite decrease as the temperature increases, which reflects exothermic adsorption. At the same time, n obtained from the liner Freundlich model (Table 4) suggested that the adsorption process underwent some difficulties on the CCsMFO nanocomposite surface as a heterogeneous surface [58,59].
For the adsorption isotherm modals parameters of Th(IV) onto CCsMFO nanocomposite listed in Table 3 and Figure 13A–C, where Langmuir (II), Freundlich, and D-R models R2 values are >0.89, but the R2 value obtained from Freundlich modal was the best value, implying that the adsorption of Th(IV) onto CCsMFO nanocomposite can be described using the Freundlich model, where (n) was obtained from the liner Freundlich model (Table 3), suggesting that the adsorption process experiences some difficulty on the CCsMFO nanocomposite surface as a heterogeneous surface. At the same time, R2 values of the D-R model were better than the R2 of Langmuir value, where binding energy (E) values obtained from the D-R model are ≥ 8.00 kJ/mol at different temperatures, indicating that the adsorption mechanism occurred by the chemical adsorption process [52,59,60].

2.2.7. Adsorption Thermodynamics

Gibbs free energy (ΔG°), enthalpy (ΔH°), and entropy (ΔS°) as thermodynamic parameters were obtained at 25 °C and listed in Table 5 using Van’t Hoff (Equation (11)). Kd was calculated using (Equation (12)), then ΔH°/R and ΔS°/R were calculated from the slope and intercept values of lnKd Vs. 1/T plot (Figure 14). Where ΔH° and entropy ΔS° were easily calculated and facilitated the calculation of ΔG° using Equation (12) [61,62,63,64].
K d = q e C e
L n K d = Δ S ° R Δ H ° R T
Δ G ° = Δ H ° T Δ S °
The Van’t Hoff plot shows that U(VI) adsorption processes onto CCsMFO nanocomposite at pH 3.0 are endothermic. In contrast, Th(IV) adsorption processes onto CCsMFO nanocomposite at pH 3.0 are classified as exothermic [61,62,63,64].

2.3. Reusability of CCsMFO Nanocomposite

The reusability of CCsMFO nanocomposite was examined by batch desorption test, started by loading U(VI) and Th(IV) onto CCsMFO nanocomposite surfaces, then leached using five cycles with different intervals of time. Figure 15A,B shows the % desorption of U(VI) and Th(IV) from CCsMFO nanocomposite using 0.10 M NaCl, EDTA, and Na2CO3 eluents. The used eluents have a basic pH to keep the CCsMFO nanocomposite stable. The results in Figure 15A show that the highest desorption percentage of U(VI) was 77.27% using Na2CO3 eluent for 12 h, and the highest desorption percentage of Th(IV) was 21.82% using EDTA eluent for 8 h (Figure 15B) [38].

2.4. Comparative Study

Table 6 shows qm for adsorption U(VI) and Th(IV) by different nanosorbents at different pH and 25 °C. CCsMFO nanocomposite has a unique maximum adsorption capacity qm with 500.00 mg/g, describing the adsorption of U(VI) compared to other nanomaterials that contain NH2 and OH groups and MnFe2O4 [38], as shown in Table 6. In contrast, the qm of adsorption Th(IV) by CCsMFO with 77.19 mg/g compared with MnFe2O4 [38] and Fe3O4 magnetic ion-imprinted chitosan in [65].

3. Material, Instruments, and Methods

3.1. Material

All reagents used were analytical-grade reagents with no further purification. Chitosan (CS) shrimp shells, ≥75%, and glutaraldehyde (25% v/v) in aqueous form were from SIGMA-ALDRICH. Sodium, hydrochloric acid 37% (HCl), sodium carbonate (Na2CO3), EDTA, and potassium hydrogen phthalate (KHP) were from BDH PROLABO. Thorium(IV) nitrate tetrahydrate (Th(NO3)4∙4H2O) and Uranyl(VI) nitrate hexahydrate (UO2(NO3)2.6H2O) from BDH Chemicals Ltd. Poole England. Sodium chloride (NaCl) and hydroxide (NaOH) were from GAINLAND CHEMICAL COMPANY (GCC). Acetone and Ethanol were from SELVO CHEM. 99.5% glacial acetic acid was from TEDIA. Arsenazo(III) indicator was from JANSSEN CHIMICA.

3.2. Instruments

The metal ions’ concentration was determined using a Vis Spectrophotometer from the METASH model V-5100 and a 1.0 cm quartz cell. Fourier-transform Infrared Spectroscopy (FT-IR) spectra were measured using Thermo Nicolet NEXUS 670 FT–IR Spectrophotometer. Thermal gravimetric analysis (TGA) was examined using NETZCH STA 409 PG/PC, a thermal analyzer in the temperature range of 25 °C–1000 °C at a heating rate of 20 °C/min. X-ray diffraction (XRD) was measured using Philips X pert PW 3060, operated at 45 kV and 40 mA. The shape with 3-dimensional (3D) surface morphology was examined with NCFL’s FEI QUANTA 600 FEG scanning electron microscope (SEM). The surface’s nature and average size were carried out using a magneton MORGAGNI FEI 500 tunneling electron microscopy. Solution pH was measured using a METTLER TOLEDO pH–meter. Zeta potential was measured using Zetasizer Nano ZS90 (Malvern Instruments). The weighing was performed by RADWAG ® AS 220. R2 Electronic Balance. Samples were shaken using a GFL-1083 thermostatic shaker.

3.3. Methods

3.3.1. Cross-Linked Chitosan/MnFe2O4 (CCsMFO) Nanocomposite Synthesis

A 2% w/v chitosan solution was prepared by dissolving 1.00 g chitosan into 50 mL of 2% (v/v) acetic acid with vigorous stirring (1300 rpm) for 1 h to obtain a clear, homogenous gel. Then, 0.25 g of previously fabricated MnFe2O4 NPs [38] was added under vigorous stirring (1300 rpm) for 1 h to obtain a homogenous black gel. After that, 0.5 mL of 25 wt.% glutaraldehyde was added under vigorous stirring (1300 rpm) for 3 h. Then, the pH was slowly raised to 12.0 by adding 5.0 M NaOH solution dropwise under vigorous stirring (1300 rpm), where the mixture separated into a transparent aqueous layer and a black magnetic precipitate as shown in Scheme 1. A neodymium magnet was used to separate the magnetic product, and then washed with distilled water, ethanol, and acetone, then dried at 100 °C for 24 h.

3.3.2. Sorption and Recovery Experiments

Th(IV) and U(VI) standard solutions were prepared by dissolving UO2(NO3)2.6H2O) and (Th(NO3)4.4H2O) salt with different concentrations (10–100 ppm), with adjusted pH using buffer solution (HCl)/(KHP) and NaOH/KHP to adjust the solution pH of 1.0, 2.0, and 5.0–7.0, but pH 3.0 and 4.0 were adjusted using NaCl/HCl buffer solution.
Batch methods were performed to optimize adsorption-experiment conditions by executing them at 25 °C using CCsMFO nanocomposite. A total of 5.0 mL of 50 ppm of Th(IV) and U(VI) was used to optimize the mass of sorbent, solution pH, initial metal-ions concentration (Ci), and contact time. Certain amounts of CCsMFO nanocomposite ranging from (1.0–5.0) mg were used to find the optimum mass of sorbent condition. The optimized mass of sorbent was used with 5.0 mL of 50 ppm U(VI) and Th(IV) solutions to optimize solution pH. Then, at optimized pH conditions, 5.0 mL of 50 ppm of U(VI) and Th(IV) samples were shaken for 24 h. A regular sampling was taken every 30 min to investigate the adsorption-process equilibrium conditions. The data produced from optimization experiments were manipulated to investigate Langmuir, Freundlich, and D-R isotherm models by finding adsorption capacity qe (mg/g) (equilibrium amount of adsorbate Th(IV) and U(VI) adsorbed per unit mass of sorbent (CCsMFO nanocomposite)); and removal yield obtained from Equations (13) and (14), respectively.
q e = ( C o C e ) V m  
%   R = ( C o C e ) C o   × 100
where Ce is the equilibrium concentration of U(VI) and Th(IV) (mg/L), Co is the initial concentration of U(VI) and Th(IV) (mg/L), V is the solution volume (L), and m is (CCsMFO nanocomposite) mass (mg).
The recovery of CCsMFO nanocomposite was explored using batch techniques starting by loading U(VI) and Th(IV) onto CCsMFO nanocomposite. Then, U(VI) and Th(IV) were leached through 5 cycles with different time intervals, using 0.10 M of NaCl, Na2CO3, and EDTA eluents, where % recovery was obtained by Equation (15):
%   R = ( C a d s     C d e s ) C a d s × 100 %    

3.3.3. Determination of U(VI) and Th(IV) Concentration

The concentration of U(VI) and Th(IV) was obtained for each metal ion separately by Vis absorption spectroscopy, where Arsenazo(III) indicator was used as a specific colorimetric agent by mixing 0.50 mL of 0.1% Arsenazo(III) indicator with 1.0 mL of the U(VI) sample with 10.0 mL of 0.01 M hydrochloric acid solution and 1.0 mL of Th(IV) solution with 10.0 mL of 9.0 M hydrochloric acid solution [38,70]. Absorbance measurements were carried out after sample preparation at 660 nm to detect Th(IV) ions and 650 nm to detect U(VI) ions using a 1.0 cm quartz cell within one hour [38,70].

4. Conclusions

In this work, a green synthesis method was used to prepare a cross-linked chitosan/(MnFe2O4) CCsMFO nanocomposite, and it was used to extract U(VI) and Th(IV) from an aqueous solution based on adsorption phenomena. The Langmuir maximum adsorption capacities (qm) of U(VI) are 500.00, 250.00, and 111.11 (mg/g) at T = 25, 35, and 45 °C, respectively. The (qm) values of U(VI) uptake onto CCsMFO nanocomposite decrease as the temperature increases, reflecting exothermic adsorption. The Langmuir and D-R maximum adsorption capacities (qm) of Th(IV) are unaffected by temperature. Furthermore, the adsorption equilibrium model fitted the pseudo-second-order equilibrium model. Based on adsorption, kinetic and isotherm results prove that the extraction of U(VI) and Th(IV) using CCsMFO nanocomposite is explained by a physical-chemical adsorption process.

Author Contributions

Writing, review, and editing, M.A.; review and editing, A.A. and M.B.; supervision, F.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the School of Graduate Studies at The University of Jordan.

Data Availability Statement

Not Applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. El-Atawy, M.A.; Khalil, K.D.; Bashal, A.H. Chitosan Capped Copper Oxide Nanocomposite: Efficient, Recyclable, Heterogeneous Base Catalyst for Synthesis of Nitroolefins. Catalysts 2022, 12, 964. [Google Scholar] [CrossRef]
  2. Sanjari, A.J.; Asghari, M. A Review on Chitosan Utilization in Membrane Synthesis. ChemBioEng Rev. 2016, 3, 134–158. [Google Scholar] [CrossRef]
  3. Kaur, K.; Dattajirao, V.; Shrivastava, V.; Bhardwaj, U. Isolation and Characterization of Chitosan-Producing Bacteria from Beaches of Chennai, India. Enzym. Res. 2012, 2012, 421683. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Qu, X.; Wirsen, A.; Albertsson, A.C. Effect of lactic/glycolic acid side chains on the thermal degradationkinetics of chitosan derivatives. Polymer 2000, 41, 4841–4847. [Google Scholar] [CrossRef]
  5. Ngah, S.W.; Hanafiah, K.A.M. Adsorption of dyes and heavy metal ions by chitosan composites: A review. Carbohydr. Polym. 2011, 83, 1446–1456. [Google Scholar] [CrossRef]
  6. Ma, J.; Xin, C.; Tan, C. Preparation, physicochemical and pharmaceutical characterization of chitosan from Catharsius molossus residue. Int. J. Biol. Macromol. 2015, 80, 547–556. [Google Scholar] [CrossRef] [PubMed]
  7. Pontius, F.W. Chitosan as a Drinking Water Treatment Coagulant. Am. J. Civil Eng. 2016, 4, 205–215. [Google Scholar] [CrossRef] [Green Version]
  8. Ray, M.; Pal, K.; Anis, A.; Banthia, A.K. Development and Characterization of Chitosan based Polymeric Hydrogel Membranes. Des. Monomers Polym. 2010, 13, 193–206. [Google Scholar] [CrossRef] [Green Version]
  9. Pella, M.C.G.; Lima-Tenorio, M.K.; Tenório-Neto, E.T.; Guilherme, M.R.; Muniz, E.C.; Rubira, A.F. Chitosan-based hydrogels: From preparation to biomedical applications. Carbohydr. Polym. 2018, 196, 233–245. [Google Scholar] [CrossRef]
  10. Jayakumar, R.; Prabaharan, M.; Nair, S.; Tamura, H. Novel chitin and chitosan nanofibers in biomedical applications. Biotechnol. Adv. 2009, 28, 142–150. [Google Scholar] [CrossRef]
  11. Kumar, S.; Ye, F.; Dobretsov, S.; Dutta, J. Chitosan Nanocomposite Coatings for Food, Paints, and Water Treatment Applications. Appl. Sci. 2019, 9, 2409. [Google Scholar] [CrossRef] [Green Version]
  12. Sheth, Y.; Dharaskar, S.; Khalid, M.; Sonawane, S. An environment friendly approach for heavy metal removal from industrial wastewater using chitosan based biosorbent: A review. Sustain. Energy Technol. Assess. 2021, 43, 100951. [Google Scholar] [CrossRef]
  13. Zhang, Y.; Zhao, M.; Cheng, Q.; Wang, C.; Li, H.; Han, X.; Fan, Z.; Su, G.; Pan, D.; Li, Z. Research progress of adsorption and removal of heavy metals by chitosan and its derivatives: A review. Chemosphere 2021, 279, 130927. [Google Scholar] [CrossRef] [PubMed]
  14. Gokila, S.; Gomathi, T.; Sudha, P.N.; Anil, S. Removal of the heavy metal ion chromiuim(VI) using chitosan and alginate nanocomposites. Int. J. Biol. Macromol. 2017, 104, 1459–1468. [Google Scholar] [CrossRef]
  15. Mohanasrinivasan, V.; Mishra, M.; Paliwal, J.S.; Singh, S.K.; Selvarajan, E.; Suganthi, V.; Subathra Devi, C. Studies on heavy metal removal efficiency and antibacterial activity of chitosan prepared from shrimp shell waste. 3 Biotech 2014, 4, 167–175. [Google Scholar] [CrossRef] [Green Version]
  16. Wang, J.; Chen, C. Chitosan-based biosorbents: Modification and application for biosorption of heavy metals and radionuclides. Bioresour. Technol. 2014, 160, 129–141. [Google Scholar] [CrossRef]
  17. Cervera, M.L.; Arnal, M.C.; de la Guardia, M. Removal of heavy metals by using adsorption on alumina or chitosan. Anal Bioanal. Chem. 2003, 375, 820–825. [Google Scholar]
  18. Annaduzzaman, M. Chitosan Biopolymer as an Adsorbent for Drinking Watertreatment-Investigation on arsenic and uranium. Trita-Lwr Lic 2015, 2, 26. [Google Scholar]
  19. Zemskova, L.; Egorin, A.; Tokar, E.; Ivanov, V. Chitosan-based biosorbents: Immobilization of metal hexacyanoferrates and application for removal of cesium radionuclide from aqueous solutions. J. Sol-Gel Sci. Technol. 2019, 92, 459–466. [Google Scholar] [CrossRef]
  20. Yin, Y.; Wang, J.; Yang, X.; Li, W. Removal of Strontium Ions by Immobilized Saccharomyces Cerevisiae in Magnetic Chitosan Microspheres. Nucl. Eng. Technol. 2017, 49, 172–177. [Google Scholar] [CrossRef] [Green Version]
  21. Chen, Y.; Wang, J. Removal of radionuclide Sr2+ ions from aqueous solution using synthesized magnetic chitosan beads. Nucl. Eng. Des. 2012, 242, 445–451. [Google Scholar] [CrossRef]
  22. Zhu, Y.; Hu, J.; Wang, J. Removal of Co2+ from radioactive wastewater by polyvinyl alcohol (PVA)/chitosan magnetic composite. Prog. Nucl. Energy 2014, 71, 172–178. [Google Scholar] [CrossRef]
  23. Kyzas, G.Z.; Bikiaris, D.N.; Mitropoulos, A.C. Chitosan adsorbents for dye removal: A review. Polym. Int. 2017, 66, 1800–1811. [Google Scholar] [CrossRef]
  24. Pinkerton, L.E.; Bloom, T.F.; Hein, M.J. Mortality among a cohort of uranium mill workers: An update. Occup. Environ. Med. 2004, 61, 57–64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Ara, A.; Khattak, R.; Khan, M.S.; Begum, B.; Khan, S.; Han, C. Synthesis, Characterization, and Solar Photo-Activation of Chitosan-Modified Nickel Magnetite Bio-Composite for Degradation of Recalcitrant Organic Pollutants in Water. Catalysts 2022, 12, 983. [Google Scholar] [CrossRef]
  26. Frank, L.A.; Onzi, G.R.; Morawski, A.S.; Pohlmann, A.R.; Guterres, S.S.; Contr, R.V. Chitosan as a coating material for nanoparticles intended for biomedicall applications. React. Funct. Polym. 2020, 147, 104459. [Google Scholar] [CrossRef]
  27. Shukla, S.K.; Mishra, A.K.; Arotiba, O.A.; Mamba, B.B. Chitosan-based nanomaterials: A state-of-the-art review. Int. J. Biol. Macromol. 2013, 59, 46–58. [Google Scholar] [CrossRef]
  28. Liu, T.; Fu, J.; Gou, D.; Hu, Y.; Tang, Q.; Zhao, J.; Li, X. Chitosan-Derived Magnetic Nanomaterials: Synthesis, Characterization, and Nitrite Adsorption in Water. J. Nanomater. 2021, 2021, 6420341. [Google Scholar] [CrossRef]
  29. Ali, A.; Ahmed, S. A Review on Chitosan and its Nanocomposites in Drug Delivery. Int. J. Biol. Macromol. 2017, 109, 273–286. [Google Scholar] [CrossRef]
  30. Sutirman, Z.A.; Sanagi, M.M.; Abd Karim, K.J.; Ibrahim, W.A.W.; Jume, B.H. Equilibrium, kinetic and mechanism studies of Cu(II) and Cd(II) ions adsorption by modified chitosan beads. Int. J. Biol. Macromol. 2018, 116, 255–263. [Google Scholar] [CrossRef]
  31. Monier, M.; Ayad, D.M.; Wei, Y.; Sarhanb, A.A. Adsorption of Cu(II), Co(II), and Ni(II) ions by modified magnetic chitosan chelating resin. J. Hazard. Mat. 2010, 177, 962–970. [Google Scholar] [CrossRef] [PubMed]
  32. Chavan, N.; Mane, S.; Ponrathnam, S. Effect of Chemical Crosslinking on Properties of Polymer Microbeads. A Review. Can. Chem. Trans. 2016, 3, 473–485. [Google Scholar]
  33. Lee, S.M.; Reddy, D.H.K. Application of Magnetic Chitosan Composites for the Removal of Toxic Metal and Dyes from Aqueous Solutions. Adv. Colloid Interface Sci. 2013, 201–202, 68–93. [Google Scholar]
  34. Wang, Z.; Yanzhen, X.; Liang, H. MnFe2O4/Chitosan Nanocomposites as a Recyclable Adsorbent for the Removal of Hexavalent Chromium. Mater. Res. Bull. 2013, 48, 3910–3915. [Google Scholar]
  35. Khodadust, R.; Unsoy, G.; Yalcin, S.; Gunduz, G.; Gunduz, U. Synthesis Optimization and Characterization of Chitosan Coated Iron Oxide Nanoparticles Produced for Biomedical Applications. J Nanopart. Res. 2012, 14, 964–977. [Google Scholar]
  36. Elwakeel, K.Z.; Atia, A.A.; Guibal, E. Fast removal of uranium from aqueous solutions using tetraethylenepentamine modified magnetic chitosan resin. Bioresour. Technol. 2014, 160, 107–114. [Google Scholar] [CrossRef]
  37. Elwakeel, K.Z. Removal of Cr(VI) from Alkaline Aqueous Solutions Using Chemically Modified Magnetic Chitosan Resins. Desalination 2010, 250, 105–112. [Google Scholar] [CrossRef] [Green Version]
  38. Alaqarbeh, M.; Khalili, F.I.; Kanoun, O. Manganese ferrite (MnFe2O4) as potential nanosorbent for adsorption of uranium(VI) and thorium(IV). J. Radioanal. Nucl. Chem. 2020, 323, 515–537. [Google Scholar] [CrossRef]
  39. Dinesha, B.L.; Sharanagouda, H.; Udaykumar, N.; Ramachandr, C.T.; Dandekar, A.B. Removal of Pollutants from Water/Waste Water Using Nano-Adsorbents: A Potential Pollution Mitigation. Int. J. Curr. Microbiol. Appl. Sci. 2017, 6, 4868–4872. [Google Scholar] [CrossRef]
  40. Newbury, D.E. Mistakes encountered during automatic peak identification of minor and trace constituents in electron-excited energy dispersive X-ray microanalysis. Scanning 2009, 31, 91–101. [Google Scholar] [CrossRef]
  41. Szymczyk, A.; Fievet, P.; Mullet, M.; Reggiani, J.C.; Pagetti, J. Comparison of two electrokinetic methods-electroosmosis and streaming potential-to determine the zeta-potential of plane ceramic membranes. J. Membrane Sci. 1998, 143, 189–195. [Google Scholar] [CrossRef]
  42. Katsoyiannis, I.A.; Zouboulis, A.I. Removal of uranium from contaminated drinking water: A mini review of available treatment methods. Desalination Water Treat. 2013, 51, 2915–2925. [Google Scholar] [CrossRef]
  43. Persson, I.; Torapava, N.; Eriksson, I.; Lundberg, D. Hydration and Hydrolysis of Thorium(IV) in Aqueous Solution and the Structures of Two Crystalline Thorium(IV) Hydrates. Inorg. Chem. 2009, 48, 11712–11723. [Google Scholar]
  44. Choppin, G. Actinide speciation in aquatic systems. Mar. Chem. 2006, 99, 83–92. [Google Scholar] [CrossRef]
  45. Grenthe, I.; Szabo, Z.; Toraishi, T.; Vallet, V. Solution coordination chemistry of actinides: Thermodynamics, structure and reaction mechanisms. Coord. Chem. Rev. 2006, 250, 784–815. [Google Scholar]
  46. Bhaumik, M.; Setshedi, K.; Maity, A.; Onyango, M.S. Chromium(VI) removal from water using fixed bed column of polypyrrole/Fe3O4 nanocomposite. Sep. Purif. Technol. 2013, 110, 11–19. [Google Scholar] [CrossRef]
  47. Robati, D. Pseudo-Second-Order Kinetic Equations for Modeling Adsorption Systems for Removal of Lead Ions Using Multi-Walled Carbon Nanotube. J. Nanostruct. Chem. 2013, 3, 55–61. [Google Scholar] [CrossRef] [Green Version]
  48. Salameh, S.I.Y.; Khalili, F.I.; Al-Dujaili, A.H. Removal of U(VI) and Th(IV) from Aqueous Solutions by Organically Modified Diatomaceous Earth: Evaluation of Equilibrium, Kinetic and Thermodynamic Data. Int. J. Miner. Process. 2017, 168, 9–18. [Google Scholar] [CrossRef]
  49. Khalili, F.I.; Al-Kakah, M.S.; Ayoub, M.M.; Ismail, L.S. Sorption of Pb(II), Cd(II) and Zn(II) ions from aqueous solution using Jordanian kaolinite modified by the amino acids methionine or cysteine. Desalination Water Treat. 2019, 151, 280–294. [Google Scholar] [CrossRef]
  50. Zhou, L.; Zou, H.; Wang, Y.; Huang, Z.; Wang, Y.; Luo, T.; Liu, Z.; Adesina, A.A. Adsorption of uranium(VI) from aqueous solution using magnetic carboxymethyl chitosan nano-particles functionalized with ethylenediamine. J. Radioanal. Nucl. Chem. 2016, 308, 935–946. [Google Scholar] [CrossRef]
  51. Giles, C.H.; Smith, D.; Huitson, A. A General Treatment and Classification of the Solute Adsorption Isotherm. I. Theoretical. J. Colloid Interface Sci. 1974, 47, 755–765. [Google Scholar] [CrossRef]
  52. Humelnicu, D.; Dinu, M.V.; Dragan, E.S. Adsorption characteristics of UO22+ and Th4+ ions from simulated radioactive solutions onto chitosan/clinoptilolite sorbents. J. Hazard. Mater. 2011, 185, 447–455. [Google Scholar] [CrossRef] [PubMed]
  53. Alaqarbeh, M. Adsorption Phenomena: Definition, Mechanisms, and Adsorption Types: Short Review. RHAZES Green Appl. Chem. 2021, 13, 43–51. [Google Scholar]
  54. Liu, S. Cooperative adsorption on solid surfaces. J. Colloid Interface Sci. 2015, 450, 224–238. [Google Scholar] [CrossRef] [PubMed]
  55. Nasiri, A.; Rajabi, S.; Amiri, A.; Fattahizade, M.; Hasani, O.; Lalehzari, A.; Hashemi, M. Adsorption of tetracycline using CuCoFe2O4@Chitosan as a new and green magnetic nanohybrid adsorbent from aqueous solutions: Isotherm, kinetic and thermodynamic study. Arab. J. Chem. 2022, 15, 104014. [Google Scholar] [CrossRef]
  56. Al-Wasidi, A.S.; Naglah, A.M.; Saad, F.A.; Abdelrahman, E.A. Modification of silica nanoparticles with 1-hydroxy-2-acetonaphthone as a novel composite for the efficient removal of Ni(II), Cu(II), Zn(II), and Hg(II) ions from aqueous media. Arab. J. Chem. 2022, 15, 104010. [Google Scholar] [CrossRef]
  57. Afolabi, H.K.; Nasef, M.M.; Hadi, N.A.; Nordin, M.; Ting, T.M.; Harun, N.Y.; Saeed, A.A.H. Isotherms, kinetics, and thermodynamics of boron adsorption on fibrous polymeric chelator containing glycidol moiety optimized with response surface method. Arab. J. Chem. 2021, 14, 103453. [Google Scholar] [CrossRef]
  58. Zhang, Y.; Miao, B.; Chen, Q.; Bai, Z.; Cao, Y.; Davaa, B. Synthesis, Structure, and Photocatalytic Activity of TiO2-Montmorillonite Composites. Catalysts 2022, 12, 486. [Google Scholar] [CrossRef]
  59. Chatla, A.; Almanassra, I.W.; Kochkodan, V.; Laoui, T.; Alawadhi, H.; Atieh, M.A. Efficient Removal of Eriochrome Black T (EBT) Dye and Chromium (Cr) by Hydrotalcite-Derived Mg-Ca-Al Mixed Metal Oxide Composite. Catalysts 2022, 12, 1247. [Google Scholar] [CrossRef]
  60. Aouaini, F.; Knani, S.; Yahia, B.M.; Lamine, A.B. Statistical physics studies of multilayer adsorption isotherm in food materials and pore size distribution. Phys. A Stat. Mech. Appl. 2015, 432, 373–390. [Google Scholar] [CrossRef]
  61. Dastbaza, A.; Reza, K.A. Adsorption of Th4+, U6+, Cd2+, and Ni2+ from aqueous solution by a novel modified polyacrylonitrile composite nanofiber adsorbentprepared by electrospinning. Appl. Surf. Sci. 2014, 293, 336–344. [Google Scholar] [CrossRef]
  62. Xue, H.; Guo, X.; Mao, D.; Meng, T.; Yu, J.; Ma, Z. Phosphotungstic Acid-Modified MnOx for Selective Catalytic Reduction of NOx with NH3. Catalysts 2022, 12, 1248. [Google Scholar] [CrossRef]
  63. Ayawei, N.; Ebelegi, A.N.; Wankasi, D. Modelling and Interpretation of Adsorption Isotherm. J. Chem. 2017, 2017, 3039817. [Google Scholar] [CrossRef] [Green Version]
  64. Li, G.; Xu, H.; Li, J.; Chen, C.; Ren, X. Interaction of Th(IV) with graphene oxides: Batch experiments, XPS investigation, and modeling. J. Mol. Liq. 2016, 213, 58–68. [Google Scholar]
  65. Husnain, S.M.; Kim, H.J.; Um, W.; Chang, Y.Y.; Chang, Y.S. Superparamagnetic Adsorbent Based on Phosphonate Grafted Mesoporous Carbon for Uranium Removal. Ind. Eng. Chem. Res. 2017, 56, 9821–9830. [Google Scholar] [CrossRef]
  66. Huang, G.; Chen, Z.; Wang, L.; Lv, T.; Shi, J. Removal of thorium(IV) from aqueous solution using magnetic ion-imprinted chitosan resin. J. Radioanal. Nucl. Chem. 2016, 310, 1265–1272. [Google Scholar] [CrossRef]
  67. Abd El-Magied, M.O.; Tolba, A.A.; El-Gendy, H.S.; Zaki, S.A.; Atia, A.A. Studies on the recovery of Th(IV) ions from nitric acid solutions using amino-magnetic glycidyl methacrylate resins and application to granite leach liquors. Hydrometallurgy 2017, 169, 89–98. [Google Scholar] [CrossRef]
  68. Mirzabe, G.H.; Keshtkar, A.R. Application of response surface methodology for thorium adsorption on PVA/Fe3O4/SiO2APTES nanohybrid adsorbent. J. Ind. Eng. Chem. 2015, 26, 277–285. [Google Scholar] [CrossRef]
  69. Wang, Z.; Wang, Y.; Yao, C. Highly Efficient Removal of Uranium(VI) From Aqueous Solution Using the Polyethyleneimine Modified Magnetic Chitosan. J. Polym. Environ. 2022, 30, 855–866. [Google Scholar] [CrossRef]
  70. Savvin, S.B. Analytical Use of Arsenazo(III), Determination of Thorium, Zirconium, Uranium and Rare Earth Elements. Talanta 1961, 8, 673–685. [Google Scholar] [CrossRef]
Figure 1. General chitosan chemical structure.
Figure 1. General chitosan chemical structure.
Catalysts 13 00047 g001
Figure 2. Spectrum of FT-IR for (A) MnFe2O4 NPs, (B) chitosan, and (C) CCsMFO nanocomposite.
Figure 2. Spectrum of FT-IR for (A) MnFe2O4 NPs, (B) chitosan, and (C) CCsMFO nanocomposite.
Catalysts 13 00047 g002
Figure 3. XRD spectrum for (A) MnFe2O4 NPs, (B) chitosan, and (C) CCsMFO nanocomposite.
Figure 3. XRD spectrum for (A) MnFe2O4 NPs, (B) chitosan, and (C) CCsMFO nanocomposite.
Catalysts 13 00047 g003
Figure 4. TGA curve for (A) MnFe2O4 NPs, (B) chitosan, and (C) CCsMFO nanocomposite.
Figure 4. TGA curve for (A) MnFe2O4 NPs, (B) chitosan, and (C) CCsMFO nanocomposite.
Catalysts 13 00047 g004aCatalysts 13 00047 g004b
Figure 5. SEM image of CCsMFO nanocomposite.
Figure 5. SEM image of CCsMFO nanocomposite.
Catalysts 13 00047 g005
Figure 6. (A,B) TEM image of CCsMFO nanocomposite.
Figure 6. (A,B) TEM image of CCsMFO nanocomposite.
Catalysts 13 00047 g006
Figure 7. EDS for (A) CCsMFO/U(VI) and (B) CCsMFO/Th(IV).
Figure 7. EDS for (A) CCsMFO/U(VI) and (B) CCsMFO/Th(IV).
Catalysts 13 00047 g007
Figure 8. Zeta potential distribution CCsMFO nanocomposite curve.
Figure 8. Zeta potential distribution CCsMFO nanocomposite curve.
Catalysts 13 00047 g008
Figure 9. (A) Effect of (CCsMFO) nanocomposite mass for removal of U(VI) at pH 3.0 and Th(IV) at pH 3.0, T = 25 °C. (B) pH effect on U(VI)% uptake. (C) pH effect on Th(IV)% uptake. (D) Contact sorption time of metal ions on CCsMFO nanocomposite. (E) Metal ion concentration effect.
Figure 9. (A) Effect of (CCsMFO) nanocomposite mass for removal of U(VI) at pH 3.0 and Th(IV) at pH 3.0, T = 25 °C. (B) pH effect on U(VI)% uptake. (C) pH effect on Th(IV)% uptake. (D) Contact sorption time of metal ions on CCsMFO nanocomposite. (E) Metal ion concentration effect.
Catalysts 13 00047 g009aCatalysts 13 00047 g009bCatalysts 13 00047 g009c
Figure 10. (A) Pseudo-first-order, (B) pseudo-second-order kinetics adsorption models of U(VI) and Th(IV) at pH =3.0 and T = 25.0 °C.
Figure 10. (A) Pseudo-first-order, (B) pseudo-second-order kinetics adsorption models of U(VI) and Th(IV) at pH =3.0 and T = 25.0 °C.
Catalysts 13 00047 g010
Figure 11. Adsorption isotherm of (A) U(VI) and (B) Th(IV) by CCsMFO nanocomposite at pH 3.0.
Figure 11. Adsorption isotherm of (A) U(VI) and (B) Th(IV) by CCsMFO nanocomposite at pH 3.0.
Catalysts 13 00047 g011aCatalysts 13 00047 g011b
Figure 12. Linear adsorption (A) Langmuir(II) isotherm, (B) Freundlich isotherm, and (C) D-R isotherm of U(VI), at pH 3.0 at 25.0, 35.0, and 45.0 °C.
Figure 12. Linear adsorption (A) Langmuir(II) isotherm, (B) Freundlich isotherm, and (C) D-R isotherm of U(VI), at pH 3.0 at 25.0, 35.0, and 45.0 °C.
Catalysts 13 00047 g012aCatalysts 13 00047 g012b
Figure 13. Linear adsorption (A) Langmuir(II) isotherm, (B) Freundlich isotherm, and (C) D-R isotherm of Th(IV), at pH 3.0 at 25.0, 35.0, and 45.0 °C.
Figure 13. Linear adsorption (A) Langmuir(II) isotherm, (B) Freundlich isotherm, and (C) D-R isotherm of Th(IV), at pH 3.0 at 25.0, 35.0, and 45.0 °C.
Catalysts 13 00047 g013aCatalysts 13 00047 g013b
Figure 14. Plot of LnKd Vs. 1/T for U(VI) and Th(IV) at pH 3.0.
Figure 14. Plot of LnKd Vs. 1/T for U(VI) and Th(IV) at pH 3.0.
Catalysts 13 00047 g014
Figure 15. Reusability of CCsMFO nanocomposite from adsorbed (A) U(VI) and (B) Th(IV) ions.
Figure 15. Reusability of CCsMFO nanocomposite from adsorbed (A) U(VI) and (B) Th(IV) ions.
Catalysts 13 00047 g015
Scheme 1. Preparation of cross-linked chitosan MnFe2O4 (CCsMFO) nanocomposite.
Scheme 1. Preparation of cross-linked chitosan MnFe2O4 (CCsMFO) nanocomposite.
Catalysts 13 00047 sch001
Table 1. CCsMFO particle size.
Table 1. CCsMFO particle size.
SamplesArea (nm)2Length (nm)
110.2415.10
213.8517.65
312.1117.88
420.9649.65
528.1663.29
Average17.0632.71
Table 2. EDS of CCsMFO nanocomposite and adsorbed metal ions percent.
Table 2. EDS of CCsMFO nanocomposite and adsorbed metal ions percent.
Weight%C KαO KαMn KαFe KαU KαTh Kα
Materials
MnFe2O4-CS35.0218.2412.7834.44--
MnFe2O4-CS /U(VI)30.7414.4114.2332.328.3-
MnFe2O4-CS /Th(IV)33.2316.0711.3230.56-8.82
Table 3. Parameters of kinetic models.
Table 3. Parameters of kinetic models.
Metal Ion, pHPseudo 1st OrderPseudo 2nd Orderqe (mg·g−1) Experimental
qe (mg·g−1)k1 (h−1)R2qe (mg·g−1)k2
(g·mg−1·h−1)
R2
U(VI), pH 324.720.060.9819.490.080.9919.06
Th(IV), pH 319.450.100.9143.291.071.0043.33
Table 4. Parameters Langmuir(II), Freundlich and D-R adsorption isotherm models at 25.0, 35.0, and 45.0 °C for U(VI) and Th(IV) ions onto CCsMFO nanocomposite, at pH 3.0.
Table 4. Parameters Langmuir(II), Freundlich and D-R adsorption isotherm models at 25.0, 35.0, and 45.0 °C for U(VI) and Th(IV) ions onto CCsMFO nanocomposite, at pH 3.0.
Mn+, pHU(VI), pH 3.0Th(IV), pH 3.0
T (°C)253545253545
Langmuir
qm (mg/g)500.00250.00111.1119.3128.9925.64
KL (L/mg)1.90 × 10−34.10 × 10−30.010.060.040.02
RL (mg/g)0.910.830.620.260.330.50
R21.001.001.000.970.960.94
Freundlich
Kf (L/mg)0.970.960.970.470.480.38
n1.001.021.010.560.590.56
R21.001.001.000.990.990.98
D-R
qm (mg/g) 77.1987.9682.35
β (mol2/kJ2) 1.00 × 10−52.00 × 10−52.00 × 10−5
E (kJ/mol) 223.71158.13158.13
R20.750.730.750.970.960.97
Table 5. ΔG°, ΔH°, and ΔS° of U(VI) and Th(IV) adsorption onto CCsMFO nanocomposite at 25 °C.
Table 5. ΔG°, ΔH°, and ΔS° of U(VI) and Th(IV) adsorption onto CCsMFO nanocomposite at 25 °C.
Thermodynamic ParametersU(VI)Th(IV)
pH = 3.0pH = 3.0
ΔG° kJ/mol0.15−10.53
ΔH° kJ/mol12.15−44.65
ΔS° J/K·mol40.03−114.49
Table 6. Maximum adsorption capacity qm (mg/g) of U(VI) and Th(IV) on different nanosorbents with present work MnFe2O4 NPs at 25 °C.
Table 6. Maximum adsorption capacity qm (mg/g) of U(VI) and Th(IV) on different nanosorbents with present work MnFe2O4 NPs at 25 °C.
Metal IonsNanosorbentqmpHReference
U(VI)Fe3O4 magnetic
carboxymethyl chitosan nano-
particles functionalizedwith ethylenediamine
175.404.5[50]
U(VI)phosphonate grafted
mesoporous carbon
150.004.0[65]
Th(IV)Fe3O4 magnetic
ion-imprinted chitosan
147.104.0[66]
Th(IV)amino- Fe3O4 magnetic glycidyl methacrylate nanoparticles50.893.7[67]
Th(IV)amino- Fe3O4 magnetic glycidyl divinylbenzene nanoparticles68.983.7[67]
Th(IV)Magnetic Fe3O4 /SiO2/ PVA/aminopropyltriethoxysilane (APTES) nanoparticles62.505.0[68]
U(VI)Polyethyleneimine Modified Magnetic Chitosan181.806.0[69]
U(VI)MnFe2O480.963[38]
104.044
76.505
Th(IV)MnFe2O4179.813.0[38]
U(VI)CCsMFO nanocomposite500.003.0Current study
Th(IV)CCsMFO nanocomposite77.193.0Current study
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

Alaqarbeh, M.; Khalili, F.; Bouachrine, M.; Alwarthan, A. Synthesis, Characterization and Investigation of Cross-Linked Chitosan/(MnFe2O4) Nanocomposite Adsorption Potential to Extract U(VI) and Th(IV). Catalysts 2023, 13, 47. https://doi.org/10.3390/catal13010047

AMA Style

Alaqarbeh M, Khalili F, Bouachrine M, Alwarthan A. Synthesis, Characterization and Investigation of Cross-Linked Chitosan/(MnFe2O4) Nanocomposite Adsorption Potential to Extract U(VI) and Th(IV). Catalysts. 2023; 13(1):47. https://doi.org/10.3390/catal13010047

Chicago/Turabian Style

Alaqarbeh, Marwa, Fawwaz Khalili, Mohammed Bouachrine, and Abdulrahman Alwarthan. 2023. "Synthesis, Characterization and Investigation of Cross-Linked Chitosan/(MnFe2O4) Nanocomposite Adsorption Potential to Extract U(VI) and Th(IV)" Catalysts 13, no. 1: 47. https://doi.org/10.3390/catal13010047

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