Next Article in Journal
Microbiota “Fingerprint” of Greek Feta Cheese through Ripening
Next Article in Special Issue
Unexpected Ferromagnetism—A Review
Previous Article in Journal
Analysis of Compound Stained Cervical Cell Images Using Multi-Spectral Imaging
Previous Article in Special Issue
Theoretical Study on Specific Loss Power and Heating Temperature in CoFe2O4 Nanoparticles as Possible Candidate for Alternative Cancer Therapy by Superparamagnetic Hyperthemia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Polyethylenimine-Modified Magnetic Chitosan for the Uptake of Arsenic from Water

by
Norah Salem Alsaiari
1,
Fatimah Mohammed Alzahrani
1,*,
Khadijah Mohammedsaleh Katubi
1,*,
Abdelfattah Amari
2,3,*,
Faouzi Ben Rebah
4 and
Mohamed A. Tahoon
5,6
1
Chemistry Department, College of Science, Princess Nourah bint Abdulrahman University, Riyadh 11671, Saudi Arabia
2
Department of Chemical Engineering, College of Engineering, King Khalid University, Abha 61411, Saudi Arabia
3
Research Laboratory of Energy and Environment, Department of Chemical Engineering, National School of Engineers, Gabes University, Gabes 6072, Tunisia
4
Higher Institute of Biotechnology of Sfax (ISBS), Sfax University, P.O. Box 263, Sfax 3000, Tunisia
5
Department of Chemistry, College of Science, King Khalid University, P.O. Box 9004, Abha 61413, Saudi Arabia
6
Chemistry Department, Faculty of Science, Mansoura University, Mansoura 35516, Egypt
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2021, 11(12), 5630; https://doi.org/10.3390/app11125630
Submission received: 3 June 2021 / Revised: 12 June 2021 / Accepted: 14 June 2021 / Published: 18 June 2021
(This article belongs to the Special Issue Nano- and Biomagnetism)

Abstract

:

Featured Application

The present study reports the synthesis of magnetic Fe3O4/CS/PEI nanocomposite for the application as an adsorbent for the removal of arsenic ions from the aqueous solution.

Abstract

The removal of heavy metals from water has become a global environmental problem. Various materials have been applied as adsorbent to remove metals from water. In this field, nanomaterials have been gaining increasing interest due to their exceptional properties. In this work, we discuss the synthesis of a core-shell structure nanocomposite by the modification of magnetic chitosan (CS) (Fe3O4/CS) with polyethylenimine (PEI) to produce Fe3O4/CS/PEI composite for the adsorption of arsenic ions (As(V) and As(III)) from aqueous solution. The synthesized materials were characterized using X-ray diffraction (XRD), Fourier-transform infrared spectroscopy (FT-IR), transmission electron microscope (TEM), and vibrating sample magnetometer (VSM). The results indicated the successful combination of three components of the nanocomposite. The adsorption conditions were optimized by studying the effect of different parameters included pH, contact time, initial concentration, and adsorbent dosage. The optimum adsorption pH was found to be 6.7 while the optimum adsorbent dosage was found to be 2.0 and 1.5 g/L for As(III) and As(V), respectively. The removal efficiency for the uptake of As(III) and As(V) ions over Fe3O4/CS/PEI nanocomposite at optimum conditions was found to be 99.5 and 99.7%, respectively. The experimental results were fitted using Freundlich’s and Langmuir’s isotherms. The data were more fitted to Langmuir isotherm providing a suggestion of monolayer adsorption with maximum adsorption capacity equal to 77.61 and 86.50 mg/g for the removal of As(III) and As(V), respectively. Moreover, linear regression coefficient (R2) indicated that the adsorption of arsenic ions over the synthesized magnetic nanocomposite obeyed pseudo 2nd order suggesting the chemisorption process. The reusability of the nanosorbent for arsenic uptake using sodium hydroxide as eluent was also assessed up to five cycles. Interestingly, Fe3O4/CS/PEI nanocomposite can be considered as a promising adsorbent for As ions’ removal from water and should be tested for the removal of other pollutants.

1. Introduction

The 20th most abundant element on the earth’s crust is arsenic [1], which has become the reason of death in many countries [2]. Arsenic poisoning has claimed the lives of millions of people, and millions more are suffering from its toxicity [3]. Now, the most serious problem is arsenic intake with drinking water [4]. The elevated levels of arsenic in drinking water are associated with several anthropogenic and natural activities [5]. Arsenic is commonly present as arsenite (As(III)) in groundwater while presenting as arsenate (As(V)) in surface water [6]. Arsenite is more poisonous and dangerous to humans than arsenate because of its sensitivity to cell-binding sites [7]. The arsenic limit in drinking water according to WHO is 0.01 ppm [8]. Therefore, the decrease of arsenic concentration in water below this limit is a challenge. Various methods have been applied for the removal of arsenic from water such as the adsorption [9], reverse osmosis [10], ion exchange [11], membrane separation [12], electro-coagulation [13], flocculation [14], and precipitation [15].
The adsorption process has been considered as the commonly used method due to several advantages (eco-friendly, efficient, reversibility, economic, and simplicity) [16,17,18,19,20,21]. In this context, different materials have been used for arsenic removal from water via adsorption method. These materials include metal oxides [22], agricultural wastes [23], biochar [24], clay-based adsorbents [25], aluminum-based adsorbents [26], commercial activated carbon [27], and many other materials. In recent years, nanomaterials have attracted interest as adsorbents due to their exceptional properties [28,29,30,31,32,33]. Although adsorption efficiency is offered by the nanomaterials, it is not easy to recollect the used adsorbents after the pollutant adsorption. Interestingly, several magnetic nanomaterials have been used for the adsorption of contaminants from aqueous media due to their ease of separation by an external magnet, large surface area, and simple functionalization [34,35,36,37,38,39]. By using magnetic nanomaterials, filtration and centrifugation conducted after the adsorption process become unnecessary as they are replaced with an easy magnetic separation. Several reports have described the synthesis and the application of low-cost and eco-friendly magnetic adsorbents for water treatment. Among these adsorbents, Fe3O4 nanoparticles are one of the main selected materials due to their high surface area, high saturation magnetization, high affinity, low cost, ease of surface modifiability, biocompatibility, and low toxicity [40]. However, the small size of the magnetic nanoparticles (NPs) is an obstacle to their application in adsorption systems. To overcome this problem, Fe3O4 nanoparticles were combined with biopolymeric matrix as a supporting platform. Mainly, chitosan (CS) provides great adsorption ability to different types of pollutants such as heavy metals [41], proteins [42], and dyes [43] because of the presence of high hydroxyl and amino groups in its structure. Additionally, chitosan has many advantages that encourage its application in water treatment such as no harmful by-products, high affectivity, and accessibility. However, the practical applications of chitosan face certain difficulties including its dissolution in acidic medium below pH 5.5, and cannot be recovered. Many materials have been impeded by integratation with CS support to overcome this disadvantage by cross-linking (e.g., glutaraldehyde as cross-linker). This integration increases the chitosan adsorption efficiency and decreases its solubility in acidic medium. So, the cross-linked CS can be used as excellent adsorbents for the removal of different types of pollutants including anions [44] and clofibric acid [45] from water, especially if modified with proper amines such as polyethylenimine (PEI). Since the branched amine contains abundant secondary and tertiary amino groups, such modified materials are predicted to have high attraction for arsenic ions in water.
Therefore, we can conclude that the use of modified magnetic chitosan nanocomposite as adsorbent material for water treatment purposes can enhance the adsorption power of the materials toward pollutants such as arsenic ions. In this study, we aim to synthesize and characterize magnetic chitosan modified with polyethylenimine nanocomposite (Fe3O4/CS/PEI) for the application as adsorbent for the removal of arsenic ions from water. The magnetic, structural, and physicochemical properties of the synthesized nanocomposite were investigated. The arsenic adsorption using synthesized nanocomposite was determined with the investigation of different parameters’ effect on the adsorption process such as arsenic initial concentration, pH, adsorbent dose, and contact time. The nanocomposite regeneration for several times of use in arsenic ions’ adsorption was investigated, since the nanocomposite has the advantage of magnetic separation that makes simple its separation and reuse. The current study approved that the fabricated Fe3O4/CS/PEI nanocomposite is a promising adsorbent for water treatment.

2. Materials and Methods

2.1. Chemicals and Reagents

All supplied chemicals were analytical grade and used without any modification. Sodium arsenite (NaAsO2, 98.0%), sodium arsenate dibasic heptahydrate (Na2HAsO4·7H2O, 98.0%), branched polyethylenimine (PEI), CS, sodium tripolyphosphate (TPP), epichlorohydrin as a cross-linking agent, ferric chloride hexahydrate (FeCl3·6H2O), and ferrous chloride tetrahydrate (FeCl2·4H2O) were supplied from Sigma-Aldrich. Isopropanol as a solvent was supplied from Al-Nasr Co., Helwan, Egypt. All solutions were prepared using distilled water.

2.2. Synthesis of Fe3O4 Magnetic Nanoparticles

The common co-precipitation chemical method was used for the synthesis of Fe3O4 magnetic nanoparticles as described in the literature [46]. Under nitrogen atmosphere, deionized H2O (200.0 mL) was used for the dissolving of 4.75 g of ferric chloride hexahydrate and 1.747 g of ferrous chloride tetrahydrate, and this mixture was stirred vigorously and heated up to 60 °C. During the stirring process, approximately 7.5 mL of ammonium solution (33% v/v) was added drop-wise. The reaction was kept for 1.0 h at 60 °C to occur and the black product was obtained. Then, deionized H2O was used for the washing of the product until the neutral filtrate was obtained. A vacuum oven was used for drying the product at 60 °C up to 12 h. Importantly, all steps were performed under nitrogen atmosphere to prevent the transformation of the product from one phase to another in air (i.e., from magnetite to maghemite).

2.3. Synthesis of CS-Modified Fe3O4 Magnetic Nanoparticles

Firstly, 0.01 g of CS powder was used for the preparation of CS solution by dissolution in 11.0 mL of acetic acid (2.0%). After that, the CS solution was mixed with 0.26 g of previously prepared magnetic nanoparticles with the sonication up to half an hour to reach the complete dispersion of the nanoparticles. Then, the mixture was vigorously shacked at 600.0 rpm and 60.0 °C with a drop-wise of 50 mL TPP solution (0.02 g/50.0 mL H2O) for the cross-linking of chitosan on the surface of magnetic nanoparticles [47]. The synthesized core-shell nanocomposite (Fe3O4/CS) was filtered and dried using a freeze dryer for 36.0 h at −20 °C.

2.4. Synthesis of Polyethylenimine-Modified Fe3O4/CS

Firstly, 30.0 mL of isopropanol alcohol was used for washing of 0.250 g of Fe3O4/CS nanocomposite. After that, isopropanol containing epichlorohydrin (25.0 mL) was used to suspend this washed Fe3O4/CS nanocomposite with the continuous shaking of the suspension (3.0 h at 70 °C and 150.0 rpm) for the activation reaction between CS’ hydroxylmethyl groups and epichlorohydrin’s epoxy group. Until being washed three times with deionized water to extract unreacted epichlorohydrin, the Fe3O4/CS nanocomposite had been modified with 25.0 mL of polyethylenimine solution for 5.0 h at 90 °C and 150 rpm. Finally, Fe3O4/CS/PEI nanocomposite was washed by distilled water to remove unattached polyethylenimine and dried at 65.0 °C up to 12.0 h. The Fe3O4/CS/PEI nanocomposite became ready for the investigation and adsorption studies.

2.5. Batch Adsorption Experiment

For the study of the adsorption of arsenic ions (As(III) and As(V)) over the surface of synthesized adsorbents, 500.0 mL stock solutions of both ions were prepared by dissolving Na2HAsO4·7H2O (2.080 g) and NaAsO2 (0.8670 g) in distilled water to reach the concentration of 1000.0 mg/L. Then, these stock solutions were diluted using deionized water to acquire the required concentration for the batch adsorption experiment. After each experiment, the adsorbent was separated using an external magnet and the solution was examined for the presence of arsenic ions using an atomic absorption spectrometer. The effect of different parameters on the adsorption process was studied including pH, adsorbent dosage, initial concentration, and contact time. The pH effect was studied by changing the pH value in the range of 3.0 to 11.0 by using an initial concentration of arsenic ions of 10.0 mg/L, 0.50 g/L of adsorbent, and solution volume of 100.0 mL. The effect of adsorbent dosage was also studied in the range of 0.50 to 2.50 g/L at pH of 6.70, 30.0 °C, 150 rpm of shacking for 1.0 day, and solution volume of 100.0 mL. The initial concentration effect was studied in the range of 10.0 to 500.0 mg/L at pH of 6.70, 30.0 °C, 150 rpm of shacking for 1.0 day, solution volume of 100.0 mL, and adsorbent dosage of 2 and 1.5 g/L for As(III) and As(V), respectively. Additionally, the contact time effect was studied by varying the time in the range of 5.0 to 60.0 min at the same experimental conditions. The removal efficiency can be calculated using Equation (1):
Removal efficiency (%) = ((Ci − Ce)/Ci) × 100,
The symbols Ci and Ce denote the initial and equilibrium concentrations of arsenic ions, respectively.
The adsorption capacity can be calculated using Equation (2):
Adsorption capacity (qe) = ((Ci − Ce)/W) × V,
The symbols W and V denote mass of adsorbent (g) and volume of solution (L), respectively.
The adsorption rate was investigated using the kinetic models of pseudo 1st order and pseudo 2nd order according to Equations (3) and (4), respectively:
ln (qe − qt) = ln qe − k1t,
t/qe = (1/k2qe2) + (1/qe),
where k1 (min−1) and k2 (g mg−1 min−1) denote the equilibrium rate constants and qt denotes the adsorption capacity at t time.
The adsorption mechanism was investigated using the linearized form of Freundlich’s isotherm that can be represented by Equation (5):
Log qe = Log KF + (1/n Log Ce),
The symbols 1/n and KF denote adsorption intensity and Freundlich’s constant, respectively.
Additionally, the adsorption mechanism was investigated using the linearized form of Langmuir’s isotherm that can be represented by Equation (6):
1/qe = (1/KL.qmax.1/Ce + (1/qmax),
The symbols KL and qmax denote Langmuir’s constant and maximum adsorption capacity, respectively.
The regression coefficient R2 value was used to determine the best fitting between experimental data and isotherm that determines the mechanism of adsorption.

2.6. Regeneration and Reusability Study

The reusability of synthesized adsorbent was studied up to five successive cycles. In each cycle, the adsorbent was mixed with the arsenic solution (10.0 mg/L, 100.0 mL) followed by the magnetic separation using external magnet, and the filtrate was examined for the presence of arsenate and arsenite. The separated adsorbent was desorbed and dried at 55 °C for 2.0 h. The desorption process was achieved by immersing the adsorbent in 0.10 M of sodium hydroxide as eluent, and the mixture was shaken for 3.0 h at 30 °C and 150.0 rpm. Then, the adsorbent was washed several times with deionized H2O and dried in an oven to be ready for the next adsorption-desorption cycle. The removal efficiency of arsenic ions over the studied adsorbent was determined in each cycle to evaluate the ability to reuse the adsorbent material for the uptake of arsenic ions from water several times.

3. Results and Discussion

3.1. The Characterization of Synthesized Nanomaterials

The synthesized nanomaterials of magnetic Fe3O4 nanoparticles, Fe3O4/CS nanocomposite, and Fe3O4/CS/PEI nanocomposite were characterized using different techniques included Fourier-transform infrared spectroscopy (FT-IR), X-ray diffraction (XRD), transmission electron microscope (TEM), and vibrating-sample magnetometer (VSM). For the characterization of the presented functional groups, FT-IR spectra were performed for Fe3O4 nanoparticles, Fe3O4/CS nanocomposite, and Fe3O4/CS/PEI nanocomposite as shown in Figure 1a. For Fe3O4 NPs, Fe–O bond vibration, and O–H bending vibration of absorbed H2O molecules have two significant bands at 622 cm−1 and 3402 cm−1, respectively [48,49]. Additionally, the significant bands of CS were recognized at 1082 cm−1, 1662 cm−1, 2882 cm−1, and 3402 cm−1 corresponding to the axial bending vibration of C-N, C=O stretching of amide I of CS, C-H vibrations of aliphatic –CH2 hydrocarbon stretch, and OH besides NH stretching vibration, respectively [50].
For the nanocomposite Fe3O4/CS/PEI, the C–N–C asymmetrical stretch and the C–N vibration in PEI were represented by the two bands at 447 cm−1 and 1072 cm−1, respectively [51]. Additionally, the branched polymer was indicated by two characteristic bands at 1632 cm−1 and 3422 cm−1 corresponding to N-H deformation and N-H stretch, respectively. Moreover, the secondary amine group of the polymer was represented by 1142 cm−1 and 1312 cm−1 small band. The band at 1462 cm−1 was related to the CH2 deformations. Finally, the clear band at 622 cm−1 approved the presence of Fe-O group. So, the FT-IR results indicated the successful implantation of Fe3O4/CS skeleton by the branched polymer PEI.
The crystalline structure of synthesized nanomaterials was studied using XRD as shown in Figure 1b. For Fe3O4 NPs, the six characteristic crystal planes (440), (551), (422), (400), (311), and (220) were represented by the reflexes at 2θ = 62, 57, 53, 43, 36, and 30°, respectively, and indicated the spinel structure of pristine Fe3O4. It can be viewed that Fe3O4 NPs have inverse spinel structure without another phase, only magnetite [46,49]. XRD patterns of Fe3O4/CS and Fe3O4/CS/PEI nanocomposites were found to be similar to that of pure Fe3O4 indicating the retention of Fe3O4 crystal forms, and the reflex positions were unchanged (i.e., the surface modification of Fe3O4 by CS or polymer did not change the phase or crystalline structure of the magnetic nanoparticles). This may be related to the preparation method of the composite that allowed the coating of chitosan on iron oxide nanoparticles [46]. The average size of synthesized nanomaterials was determined using the Debye-Scherrer Equation:
D = κλ/β cosθ,
where β, θ, λ, κ, and D denote the full-width half-maximum (FWHM), diffraction angle, wavelength of X-ray beam, Scherrer constant (0.89), and the particle size, respectively. The average size of synthesized Fe3O4/CS/PEI, Fe3O4/CS, and Fe3O4 magnetic nanomaterials was 12.1 ± 0.2 nm, 10.2 ± 0.4 nm, and 9.2 ± 0.3 nm, respectively. This indicated the higher size of Fe3O4/CS/PEI and Fe3O4/CS nanocomposites than that of pure Fe3O4, similar to other results reported in the literature [49]. The morphology of synthesized nanomaterials was also studied by the TEM analysis as shown in Figure 2.
According to the TEM image of Fe3O4 NPs (Figure 2a), the particles are homogenous with spherical shape, and the nanoparticles are also aggregated. The TEM image of the Fe3O4/CS nanocomposite (Figure 2b) showed a higher particle size of the nanocomposite than that of magnetite, which is compatible with the results of XRD. Additionally, the TEM image of Fe3O4/CS/PEI (Figure 2c) showed an increase in the particles’ size after the polymer was grafted on the magnetic chitosan. Figure 2b,c showed the core-shell structure of the nanocomposites of Fe3O4/CS and Fe3O4/CS/PEI, respectively, that can be seen at the boundary of magnetite spherical NPs. This core-shell structure was reported in many previously synthesized magnetic nanomaterials [52,53,54]. Figure 2d showed the particle size distribution of the nanocomposite Fe3O4/CS/PEI that equals to 12 nm, similar to the results of XRD. The most important characteristic of magnetic adsorbents are their magnetic separation from the experimental solution. Therefore, the magnetization curves of the synthesized Fe3O4, Fe3O4/CS, and Fe3O4/CS/PEI are shown in Figure 3.
According to Figure 3, all synthesized nanomaterials are ferromagnetic with saturation magnetization (Ms) of 71, 62, and 51 emu g−1 for Fe3O4, Fe3O4/CS, and Fe3O4/CS/PEI, respectively. The clear drop of saturation magnetization by the modification of magnetite with chitosan was attributed to the non-magnetic properties of chitosan. Similarly, the drop occurred when the PEI was added. This was attributed to the diamagnetic properties of the polymer. The drop of magnetic properties after the modification of magnetite NPs was reported in previous studies [55,56]. This order of magnetism of the three synthesized nanomaterials was reflected on their time of separation with an external magnet that was found to equal 69 s, 93 s, and 119 s for Fe3O4, Fe3O4/CS, and Fe3O4/CS/PEI, respectively. Figure 3 (inset) showed the magnetic separation of Fe3O4/CS/PEI nanocomposite with an external magnet. The synthesized nanomaterials have excellent magnetic properties sufficient for their application in water treatment with the benefit of the adsorbent magnetic separation. Moreover, Brunauer–Emmett–Teller surface area (SBET) was analyzed for the three synthesized nanomaterials as shown in Figure 4. The SBET values were found to equal 82.5, 83.1, and 74.6 m2 g−1 for Fe3O4, Fe3O4/CS, and Fe3O4/CS/PEI, respectively. The results indicated the small difference in the SBET value between Fe3O4 and Fe3O4/CS while the value was decreased by the addition of branched PEI indicating the effect of PEI on the surface properties of magnetic chitosan. This means that the core-shell structure of Fe3O4/CS/PEI nanocomposite was formed by the adsorption of PEI on the surface of magnetic chitosan that will save additional capturing sites for arsenic ions in water.

3.2. The Effect of Adsorbent Dosage

The effect of Fe3O4/CS/PEI nanocomposite’s amount on the adsorption of As(V) and As(III) ions from water was studied in the range of 0.50 g/L to 2.50 g/L as shown in Figure 5a. During the study of adsorbent dosage effect, the other factors which were kept constant included the initial adsorbate concentration, temperature, and pH value that were 10 mg/L, 30 °C, and 6.7, respectively. According to Figure 5a, the increased amount of adsorbent from 0.50 g/L to 2.0 g/L caused an increase in the removal efficiency of As(III) from 52.0 % to 99.5% while the dosage increase from 0.50 g/L to 1.5 g/L caused an increase in the removal efficiency of As(V) from 95.5% to 99.7%. This relation between adsorbent dosage and the increased removal efficiency was attributed to the vacant adsorption sites that increase by the addition of more adsorbent dosage.
As shown in Figure 5a, there was no increase in the removal efficiency of As(III) and As(V) when the dosage of Fe3O4/CS/PEI nanocomposite was increased from 2.0 g/L to 2.50 g/L and from 1.50 g/L to 2.0 g/L, respectively, meaning that the equilibrium has been reached. So, these values were selected as the optimum adsorbent dosages.

3.3. The Effect of PH Value

The effect of pH value on the adsorption of As(V) and As(III) ions from water over the surface of Fe3O4/CS/PEI nanocomposite was studied in the range of 3.0 to 11.0 as shown in Figure 5b. The pH effect is known to be a very important factor due to its dual effect on ionic form of the adsorbate and the surface charge of the adsorbent. According to Figure 5b, the adsorption of As(III) ions was decreased from pH > 9.0. While in the pH range of 5.0 to 9.0, there was no significant pH effect on their removal. This behavior could be described according to the ionic form of As(III) ions at different pH values. At pH > 9.0, As(III) ions were present in the forms of HAsO32− and H2AsO3 that cause a repulsion with the negatively charged adsorbent besides the competition with hydroxide ions for the adsorption sites, and the overall result is the decrease of As(III) adsorption. While at pH < 9.0, As(III) ions were present in the neutral forms such as H3AsO3 that allow their maximum adsorption via H-bonding and Van der Wall’s forces. Thus, the maximum removal of As(III) ions was reached at pH (6.0 to 8.0). Additionally, the pH increase caused an increase in the removal of As(V) ions over the surface of Fe3O4/CS/PEI nanocomposite until the value of pH (6.7), then the removal decreased with the pH increase. This behavior could be also explained according to the change of ionic form with pH change. At pH of 6.0 to 8.0, the As(V) ions were present as HAsO42− and HAsO4. At pH of 8.0 to 10.0, the As(V) ions were present as HAsO42−. However, at pH of 3.0 to 6.0, As(V) ions were present as HAsO4 making the maximum removal reached at pH of 6.7. These results clarify that the pH value is a key parameter for the removal of arsenic ions over the studied magnetic nanocomposite.

3.4. The Adsorption Kinetics

The effect of contact time was studied for the removal of As(V) and As(III) ions on the surface of Fe3O4/CS/PEI nanocomposite as shown in Figure 6a. The effect of time on the adsorption process was studied in the range of 5.0 to 60.0 min. According to Figure 6a, the maximum and rapid capturing of both arsenic ions was achieved in the first 5.0 min and when the time increased, there was no significant increase in the adsorption capacity of the nanosorbent toward the As(V) and As(III) ions. The adsorption equilibrium was achieved during the first 30.0 min. The large number of available and vacant adsorption sites over the surface of the nanocomposite adsorbent is responsible for this rapid kinetic action. After the equilibrium is reached, all adsorption sites were occupied by the pollutants’ ions and the contact time had no effect on the adsorption process. The kinetic behavior of arsenic removal over Fe3O4/CS/PEI nanocomposite was studied using the pseudo 1st order and the pseudo 2nd order equations as shown in Figure 6b,c, respectively. According to Figure 6b,c, the pseudo 2nd order showed better fit to the adsorption kinetics for the uptake of both arsenic ions over Fe3O4/CS/PEI nanocomposite than the pseudo 1st order and this was clear from the value of the regression coefficient (R2). R2 value is 0.999 for As(V) and As(III) ions in the case of pseudo 2nd order fit, while it is 0.630 and 0.603 for As(V) and As(III) ions, respectively, in case of pseudo 1st order fit. The pseudo 1st order and the pseudo 2nd order kinetic parameters for the uptake of arsenic ions over the studied magnetic nanocomposite were introduced in Table 1.
The agreement between the experimental data and the pseudo 2nd order kinetic model indicated that the capturing of arsenic ions over this polymer-modified magnetic chitosan happened via the chemisorption mechanism. Additionally, the chemical interaction between the arsenic ions and the adsorbent surface was the rate-limiting step [57]. The agreement between experimented data and the pseudo 2nd order for the treatment of aqueous arsenic ions was previously reported in the literature using another magnetic nanocomposite [58].

3.5. The Adsorption Isotherm

The effect of arsenic initial concentration on their adsorption over the surface of Fe3O4/CS/PEI nanocomposite was studied as shown in Figure 7a. According to Figure 7a, the increased initial concentration of arsenic caused an increase of the adsorption capacity indicating the high affinity between As(III) and As(V) ions and the nanosorbent.
This behavior was attributed to the availability of ions to be adsorbed. To understand the adsorption mechanism of arsenic ions over the Fe3O4/CS/PEI nanocomposite, the information related to the binding affinity and adsorption capacity must be obtained and thus achieved via the study of adsorption isotherm. As mentioned in the experimental section, the two classical isotherm models (Freundlich and Langmuir) were used to fit the experimental data of the arsenic ions’ adsorption over the synthesized nanosorbent as shown in Figure 7b,c. The parameters of Freundlich and Langmuir were also presented in Table 1. According to Figure 7b,c, and the R2 values, the experimental data was more fitted to Langmuir’s isotherm than Freundlich’s isotherm. R2 values are 0.999 for both ions of arsenic in case of Langmuir’s fit while equal to 0.94 in case of Freundlich’s fit. Freundlich’s isotherm suggests the heterogeneous adsorption of adsorbates over energetically dissimilar adsorption sites while Langmuir’s isotherm suggests the homogeneous adsorption of adsorbates over energetically equal sites. Thus, the adsorption of As(III) and As(V) ions occurred as a monolayer over Fe3O4/CS/PEI nanocomposite. The calculated maximum adsorption capacity (qmax) was found to equal 86.50 and 77.61 mg/g for the uptake of As(V) and As(III), respectively. These values were compared to previous results in the comparative study section. Moreover, the adsorption process of arsenic ions is favorable due to the value of RL that was less than 1.0.

3.6. Regeneration and Reusability Study

The reusability of any adsorbent is a key factor determining its sustainability, effectiveness, and applicability for real samples treatment. The reusability of Fe3O4/CS/PEI nanocomposite for the uptake of As(III) and As(V) ions was studied up to five sequential cycles as shown in Figure 8. Each cycle included the adsorption of polluted ions followed by desorption of adsorbed ions using alkaline solution (0.1 M of NaOH) as eluent. Herein, the magnetic separation was used to collect the adsorbent to be washed and dried for the next cycle.
According to Figure 8, the removal efficiency of As(III) and As(V) ions using Fe3O4/CS/PEI nanocomposite was dropped from 99.5% to 92.0% and from 99.7% to 94.0%, respectively, after the 5th cycle. The literature reported the use of Fe3O4/CS for the uptake of As(III) and As(V) ions from water with the retention of only 76.0% and 88.0% of the initial efficiency after the 5th cycle [59]. The small drop of the removal efficiencies of As(III) and As(V) ions over Fe3O4/CS/PEI nanocomposite was attributed to the damaged adsorption sites after each cycle. However, the removal efficiencies for the uptake of both ions over Fe3O4/CS/PEI nanocomposite remain over 90.0% after the 5th cycle indicating the ability to reuse this synthesized adsorbent to remove arsenic ions for several times. As a result, the Fe3O4/CS/PEI nanocomposite can be used for real water treatment with a low cost of processing due to the excellent results of reusability.

3.7. Comparative Study

The maximum adsorption capacity and pH value for the removal of arsenic ions (As(V) and As(III)) were compared with previously studied adsorbents as tabulated in Table 2. According to Table 2, the adsorption capacities in the current study are comparable to the values reported in the literature indicating the affinity of the Fe3O4/CS/PEI nanocomposite for arsenic ions. However, the difference in the adsorbent efficiency between materials can be explained by the applied operating conditions and the variability of the interactions between pollutants and the material functional groups. Generally, strong interactions with the specific functional groups of the adsorbent surface permitted high adsorption capacity. Subsequently, the synthesized Fe3O4/CS/PEI nanocomposite can be considered as a promising nanosorbent for water treatment that must be investigated for the capturing of additional metals, dyes, and pharmaceuticals in the future. However, the complete view about the treatment study should be obtained by the determination of the thermodynamic parameters [60,61,62] of the adsorption process.

4. Conclusions

The present study discussed the successful fabrication of magnetic chitosan modified with polyethylenimine to produce core-shell structure nanocomposite defined as Fe3O4/CS/PEI. As described in the experimental section, chitosan and magnetite nanoparticles were cross-linked using Na5P3O10 while the modification of the magnetic chitosan using PEI was achieved using epichlorohydrin. The successful combination between three parts of the nanocomposite (Fe3O4, CS, and PEI) was confirmed by using different characterization techniques including XRD, FT-IR, TEM, and VSM. Additionally, the spinel structure of magnetite was not changed by the modification process as shown from XRD. The synthesized nanocomposite was investigated for the adsorption of arsenic ions (As(V) and As(III)) from aqueous solution. Different parameters such as pH, contact time, adsorbent dosage, and initial concentration were examined to define the optimum conditions of adsorption. The arsenic adsorption over the Fe3O4/CS/PEI magnetic nanocomposite was found to be pH dependent and the optimum pH was detected to be 6.7. The arsenic removal over Fe3O4/CS/PEI nanocomposite according to the adsorption isotherm results occurred as a monolayer over energetically symmetrical adsorption sites. Moreover, the experimental data was better fitted with the pseudo 2nd order kinetics indicating the chemisorption uptake of arsenic ions over the nanocomposite. The magnetic separation of the synthesized nanocomposite enhanced their reusability for arsenic removal up to five successive cycles by using sodium hydroxide as eluent solution. The results indicated the ability to reuse the nanocomposite for arsenic removal with high efficiency that reduces the overall cost of water purification. All results indicated that Fe3O4/CS/PEI nanocomposite can be considered as a promising adsorbent for the removal of As(V) and As(III) ions from aqueous solution, and must be examined in the near future for the removal of additional pollutants.

Author Contributions

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

Funding

This research was funded by Deanship of Scientific Research at King Khalid University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors extend their appreciation to the Deanship of Scientific Research at King Khalid University for funding this work through the research groups’ program under grant number RGP.1/111/42. Additionally, this research was funded by the Deanship of Scientific Research at Princess Nourah Bint Abdulrahman University through the Fast-Track Research Funding Program.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Centeno, J.A.; Tseng, C.-H.; Van Der Voet, G.B.; Finkelman, R.B. Global Impacts Of Geogenic Arsenic: A Medical Geology Research Case. Ambio 2007, 36, 78–81. [Google Scholar] [CrossRef]
  2. Smith, A.H.; Goycolea, M.; Haque, R.; Biggs, M.L. Marked Increase in Bladder and Lung Cancer Mortality in a Region of Northern Chile Due to Arsenic in Drinking Water. Am. J. Epidemiol. 1998, 147, 660–669. [Google Scholar] [CrossRef] [PubMed]
  3. Balasubramaniam, T.; Bakar, A.; Uda, M.; Hashim, U.; Parmin, N.; Anuar, A.; Bakar, M.; Sulaiman, M. Potential of Syntesized Silica Nanoparticles (Si-NPs) using Corn Cob for Arsenic Heavy Metal Removal. In IOP Conference Series: Materials Science and Engineering, Proceedings of the 2nd Joint Conference on Green Engineering Technology & Applied Computing 2020, Bangkok, Thailand, 4–5 February 2020; IOP Publishing: Bristol, UK, 2020; Volume 864, p. 012187. [Google Scholar]
  4. Adeloju, S.; Khan, S.; Patti, A. Arsenic Contamination of Groundwater and Its Implications for Drinking Water Quality and Human Health in Under-Developed Countries and Remote Communities—A Review. Appl. Sci. 2021, 11, 1926. [Google Scholar] [CrossRef]
  5. Huong, P.T.L.; Huy, L.T.; Lan, H.; Thang, L.H.; An, T.T.; Van Quy, N.; Tuan, P.A.; Alonso, J.; Phan, M.-H.; Le, A.-T. Magnetic iron oxide-carbon nanocomposites: Impacts of carbon coating on the As(V) adsorption and inductive heating responses. J. Alloys Compd. 2018, 739, 139–148. [Google Scholar] [CrossRef]
  6. Siddiqui, S.I.; Chaudhry, S.A. A review on graphene oxide and its composites preparation and their use for the removal of As3+ and As5+ from water under the effect of various parameters: Application of isotherm, kinetic and thermodynamics. Process. Saf. Environ. Prot. 2018, 119, 138–163. [Google Scholar] [CrossRef]
  7. Azamat, J.; Khataee, A.; Sadikoglu, F. Computational study on the efficiency of MoS 2 membrane for removing arsenic from contaminated water. J. Mol. Liq. 2018, 249, 110–116. [Google Scholar] [CrossRef]
  8. Lizama-Allende, K.; Ayala, J.; Jaque, I.; Echeverría, P. The removal of arsenic and metals from highly acidic water in horizontal subsurface flow constructed wetlands with alternative supporting media. J. Hazard. Mater. 2021, 408, 124832. [Google Scholar] [CrossRef]
  9. Mahato, B.N.; Krithiga, T.; Thangam, M.M. Rapid adsorption of As(V) from aqueous solution by ZnO embedded in mesoporous aluminosilicate nanocomposite adsorbent: Parameter optimization, kinetic, and isotherms studies. Surf. Interfaces 2021, 23, 100636. [Google Scholar] [CrossRef]
  10. Dhar, R.; Zheng, Y.; Rubenstone, J.; van Geen, A. A rapid colorimetric method for measuring arsenic concentrations in groundwater. Anal. Chim. Acta 2004, 526, 203–209. [Google Scholar] [CrossRef]
  11. Schmidt, S.-A.; Gukelberger, E.; Hermann, M.; Fiedler, F.; Großmann, B.; Hoinkis, J.; Ghosh, A.; Chatterjee, D.; Bundschuh, J. Pilot study on arsenic removal from groundwater using a small-scale reverse osmosis system—Towards sustainable drinking water production. J. Hazard. Mater. 2016, 318, 671–678. [Google Scholar] [CrossRef]
  12. Pu, L.; Xu, Y.; Xia, Q.; Ding, J.; Wang, Y.; Shan, C.; Wu, D.; Zhang, Q.; Gao, G.; Pan, B. Ferroelectric membrane for water purification with arsenic as model pollutant. Chem. Eng. J. 2021, 403, 126426. [Google Scholar] [CrossRef]
  13. Sandoval, M.A.; Fuentes, R.; Thiam, A.; Salazar, R. Arsenic and fluoride removal by electrocoagulation process: A general review. Sci. Total. Environ. 2021, 753, 142108. [Google Scholar] [CrossRef] [PubMed]
  14. Pallier, V.; Feuillade-Cathalifaud, G.; Serpaud, B.; Bollinger, J.-C. Effect of organic matter on arsenic removal during coagulation/flocculation treatment. J. Colloid Interface Sci. 2010, 342, 26–32. [Google Scholar] [CrossRef] [PubMed]
  15. Ostermeyer, P.; Bonin, L.; Folens, K.; Verbruggen, F.; García-Timermans, C.; Verbeken, K.; Rabaey, K.; Hennebel, T. Effect of speciation and composition on the kinetics and precipitation of arsenic sulfide from industrial metallurgical wastewater. J. Hazard. Mater. 2021, 409, 124418. [Google Scholar] [CrossRef]
  16. Siddeeg, S.M.; Tahoon, M.A.; Alsaiari, N.S.; Shabbir, M.; Rebah, F.B. Application of Functionalized Nanomaterials as Effective Adsorbents for the Removal of Heavy Metals from Wastewater: A Review. Curr. Anal. Chem. 2020, 17, 4–22. [Google Scholar] [CrossRef]
  17. Alsaiari, N.; Amari, A.; Katubi, K.; Alzahrani, F.; Rebah, F.; Tahoon, M. Innovative Magnetite Based Polymeric Nanocomposite for Simultaneous Removal of Methyl Orange and Hexavalent Chromium from Water. Processes 2021, 9, 576. [Google Scholar] [CrossRef]
  18. Katubi, K.; Alsaiari, N.; Alzahrani, F.; Siddeeg, S.M.; Tahoon, M.A. Synthesis of Manganese Ferrite/Graphene Oxide Magnetic Nanocomposite for Pollutants Removal from Water. Processes 2021, 9, 589. [Google Scholar] [CrossRef]
  19. Amari, A.; Elboughdiri, N.; Ghernaout, D.; Lajimi, R.H.; Alshahrani, A.M.; Tahoon, M.A.; Ben Rebah, F. Multifunctional crosslinked chitosan/nitrogen-doped graphene quantum dot for wastewater treatment. Ain Shams Eng. J. 2021. [Google Scholar] [CrossRef]
  20. Amari, A.; Alzahrani, F.M.; Katubi, K.M.; Alsaiari, N.S.; Tahoon, M.; Ben Rebah, F. Clay-Polymer Nanocomposites: Preparations and Utilization for Pollutants Removal. Materials 2021, 14, 1365. [Google Scholar] [CrossRef]
  21. Katubi, K.; Amari, A.; Harharah, H.; Eldirderi, M.; Tahoon, M.; Ben Rebah, F. Aloe vera as Promising Material for Water Treatment: A Review. Processes 2021, 9, 782. [Google Scholar] [CrossRef]
  22. Goswami, A.; Raul, P.; Purkait, M. Arsenic adsorption using copper (II) oxide nanoparticles. Chem. Eng. Res. Des. 2012, 90, 1387–1396. [Google Scholar] [CrossRef]
  23. Tian, Y.; Wu, M.; Lin, X.; Huang, P.; Huang, Y. Synthesis of magnetic wheat straw for arsenic adsorption. J. Hazard. Mater. 2011, 193, 10–16. [Google Scholar] [CrossRef]
  24. Zhou, Z.; Liu, Y.-G.; Liu, S.-B.; Liu, H.-Y.; Zeng, G.-M.; Tan, X.; Yang, C.-P.; Ding, Y.; Yan, Z.-L.; Cai, X.-X. Sorption performance and mechanisms of arsenic(V) removal by magnetic gelatin-modified biochar. Chem. Eng. J. 2017, 314, 223–231. [Google Scholar] [CrossRef]
  25. Ren, X.; Zhang, Z.; Luo, H.; Hu, B.; Dang, Z.; Yang, C.; Li, L. Adsorption of arsenic on modified montmorillonite. Appl. Clay Sci. 2014, 97–98, 17–23. [Google Scholar] [CrossRef]
  26. Zhu, N.; Qiao, J.; Ye, Y.; Yan, T. Synthesis of mesoporous bismuth-impregnated aluminum oxide for arsenic removal: Adsorption mechanism study and application to a lab-scale column. J. Environ. Manag. 2018, 211, 73–82. [Google Scholar] [CrossRef]
  27. Nieto-Delgado, C.; Gutiérrez-Martínez, J.; Rangel-Méndez, J.R. Modified activated carbon with interconnected fibrils of iron-oxyhydroxides using Mn2+ as morphology regulator, for a superior arsenic removal from water. J. Environ. Sci. 2019, 76, 403–414. [Google Scholar] [CrossRef]
  28. Alsaiari, N.; Katubi, K.; Alzahrani, F.; Siddeeg, S.; Tahoon, M. The Application of Nanomaterials for the Electrochemical Detection of Antibiotics: A Review. Micromachines 2021, 12, 308. [Google Scholar] [CrossRef] [PubMed]
  29. Amari, A.; Al Mesfer, M.K.; Alsaiari, N.S.; Danish, M.; Alshahrani, A.M.; Tahoon, M.A.; Rebah, F.B. Electrochemical and Optical Properties of Tellurium Dioxide (TeO2) Nanoparticles. Int. J. Electrochem. Sci. 2021, 16, 210235. [Google Scholar] [CrossRef]
  30. Amari, A.; Alalwan, B.; Siddeeg, S.M.; Tahoon, M.A.; Alsaiari, N.S.; Rebah, F.B. Biomolecules Behavior on a Surface of Boron Doped/un-doped Graphene Nanosheets. Int. J. Electrochem. Sci. 2020, 15, 11427–11436. [Google Scholar] [CrossRef]
  31. Siddeeg, S.M.; Alsaiari, N.S.; Tahoon, M.A.; Rebah, F.B. The application of nanomaterials as electrode modifiers for the electrochemical detection of ascorbic acid. Int. J. Electrochem. Sci. 2020, 15, 3327–3346. [Google Scholar] [CrossRef]
  32. Tahoon, M.A.; Siddeeg, S.M.; Alsaiari, N.S.; Mnif, W.; Ben Rebah, F. Effective Heavy Metals Removal from Water Using Nanomaterials: A Review. Processes 2020, 8, 645. [Google Scholar] [CrossRef]
  33. Alzahrani, F.; Alsaiari, N.; Katubi, K.; Amari, A.; Ben Rebah, F.; Tahoon, M. Synthesis of Polymer-Based Magnetic Nanocomposite for Multi-Pollutants Removal from Water. Polymers 2021, 13, 1742. [Google Scholar] [CrossRef] [PubMed]
  34. Simonescu, C.M.; Tătăruş, A.; Culiţă, D.C.; Stănică, N.; Ionescu, I.A.; Butoi, B.; Banici, A.-M. Comparative study of CoFe2O4 nanoparticles and CoFe2O4-chitosan composite for Congo Red and Methyl Orange removal by adsorption. Nanomaterials 2021, 11, 711. [Google Scholar] [CrossRef] [PubMed]
  35. Hamza, M.F.; Fouda, A.; Elwakeel, K.Z.; Wei, Y.; Guibal, E.; Hamad, N.A. Phosphorylation of Guar Gum/Magnetite/Chitosan Nanocomposites for Uranium (VI) Sorption and Antibacterial Applications. Molecules 2021, 26, 1920. [Google Scholar] [CrossRef] [PubMed]
  36. Apriceno, A.; Silvestro, I.; Girelli, A.; Francolini, I.; Pietrelli, L.; Piozzi, A. Preparation and Characterization of Chitosan-Coated Manganese-Ferrite Nanoparticles Conjugated with Laccase for Environmental Bioremediation. Polymers 2021, 13, 1453. [Google Scholar] [CrossRef]
  37. Siddeeg, S.M.; Amari, A.; Tahoon, M.A.; Alsaiari, N.S.; Rebah, F.B. Removal of meloxicam, piroxicam and Cd+2 by Fe3O4/SiO2/glycidyl methacrylate-S-SH nanocomposite loaded with laccase. Alex. Eng. J. 2020, 59, 905–914. [Google Scholar] [CrossRef]
  38. Siddeeg, S.M.; Tahoon, M.A.; Ben Rebah, F. Simultaneous Removal of Calconcarboxylic Acid, NH4+ and PO43− from Pharmaceutical Effluent Using Iron Oxide-Biochar Nanocomposite Loaded with Pseudomonas putida. Processes 2019, 7, 800. [Google Scholar] [CrossRef] [Green Version]
  39. Amari, A.; Alzahrani, F.; Alsaiari, N.; Katubi, K.; Rebah, F.; Tahoon, M. Magnetic Metal Organic Framework Immobilized Laccase for Wastewater Decolorization. Processes 2021, 9, 774. [Google Scholar] [CrossRef]
  40. Gallo-Cordova, A.; Veintemillas-Verdaguer, S.; Tartaj, P.; Mazarío, E.; Morales, M.; Ovejero, J. Engineering Iron Oxide Nanocatalysts by a Microwave-Assisted Polyol Method for the Magnetically Induced Degradation of Organic Pollutants. Nanomaterials 2021, 11, 1052. [Google Scholar] [CrossRef] [PubMed]
  41. 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] [CrossRef]
  42. Wibowo, S.; Velazquez, G.; Savant, V.; Torres, J.A. Surimi wash water treatment for protein recovery: Effect of chitosan–alginate complex concentration and treatment time on protein adsorption. Bioresour. Technol. 2005, 96, 665–671. [Google Scholar] [CrossRef]
  43. Cheung, W.; Szeto, Y.; McKay, G. Enhancing the adsorption capacities of acid dyes by chitosan nano particles. Bioresour. Technol. 2009, 100, 1143–1148. [Google Scholar] [CrossRef] [PubMed]
  44. Kuncoro, E.P.; Roussy, J.; Guibal, E. Mercury Recovery by Polymer-Enhanced Ultrafiltration: Comparison of Chitosan and Poly(Ethylenimine) Used as Macroligand. Sep. Sci. Technol. 2005, 40, 659–684. [Google Scholar] [CrossRef]
  45. Nie, Y.; Deng, S.; Wang, B.; Huang, J.; Yu, G. Removal of clofibric acid from aqueous solution by polyethylenimine-modified chitosan beads. Front. Environ. Sci. Eng. 2014, 8, 675–682. [Google Scholar] [CrossRef]
  46. Unsoy, G.; Yalcin, S.; Khodadust, R.; Gunduz, G.; Gunduz, U. Synthesis optimization and characterization of chitosan-coated iron oxide nanoparticles produced for biomedical applications. J. Nanopart. Res. 2012, 14, 1–13. [Google Scholar] [CrossRef]
  47. Fan, W.; Yan, W.; Xu, Z.; Ni, H. Formation mechanism of monodisperse, low molecular weight chitosan nanoparticles by ionic gelation technique. Colloids Surf. B Biointerfaces 2012, 90, 21–27. [Google Scholar] [CrossRef]
  48. Jiang, H.; Chen, P.; Luo, S.; Tu, X.; Cao, Q.; Shu, M. Synthesis of novel nanocomposite Fe3O4/ZrO2/chitosan and its application for removal of nitrate and phosphate. Appl. Surf. Sci. 2013, 284, 942–949. [Google Scholar] [CrossRef]
  49. Li, G.-Y.; Jiang, Y.-R.; Huang, K.-L.; Ding, P.; Chen, J. Preparation and properties of magnetic Fe3O4–chitosan nanoparticles. J. Alloys Compd. 2008, 466, 451–456. [Google Scholar] [CrossRef]
  50. Labidi, A.; Salaberria, A.M.; Fernandes, S.C.M.; Labidi, J.; Abderrabba, M. Microwave assisted synthesis of poly (N-vinylimidazole) grafted chitosan as an effective adsorbent for mercury (II) removal from aqueous solution: Equilibrium, kinetic, thermodynamics and regeneration studies. J. Dispers. Sci. Technol. 2019, 41, 828–840. [Google Scholar] [CrossRef]
  51. Anirudhan, T.; Rauf, T.A.; Rejeena, S. Removal and recovery of phosphate ions from aqueous solutions by amine functionalized epichlorohydrin-grafted cellulose. Desalination 2012, 285, 277–284. [Google Scholar] [CrossRef]
  52. Campanile, R.; Scardapane, E.; Forente, A.; Granata, C.; Germano, R.; Di Girolamo, R.; Minopoli, A.; Velotta, R.; Della Ventura, B.; Iannotti, V. Core-Shell Magnetic Nanoparticles for Highly Sensitive Magnetoelastic Immunosensor. Nanomaterials 2020, 10, 1526. [Google Scholar] [CrossRef]
  53. Saykova, D.; Saikova, S.; Mikhlin, Y.; Panteleeva, M.; Ivantsov, R.; Belova, E. Synthesis and Characterization of Core–Shell Magnetic Nanoparticles NiFe2O4@ Au. Metals 2020, 10, 1075. [Google Scholar] [CrossRef]
  54. Niemirowicz-Laskowska, K.; Mystkowska, J.; Łysik, D.; Chmielewska, S.; Tokajuk, G.; Misztalewska-Turkowicz, I.; Wilczewska, A.Z.; Bucki, R. Antimicrobial and Physicochemical Properties of Artificial Saliva Formulations Supplemented with Core-Shell Magnetic Nanoparticles. Int. J. Mol. Sci. 2020, 21, 1979. [Google Scholar] [CrossRef] [Green Version]
  55. Rekos, K.; Kampouraki, Z.-C.; Sarafidis, C.; Samanidou, V.; Deliyanni, E. Graphene Oxide Based Magnetic Nanocomposites with Polymers as Effective Bisphenol–A Nanoadsorbents. Materials 2019, 12, 1987. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Shen, H.; Wang, Z.; Zhou, A.; Chen, J.; Hu, M.; Dong, X.; Xia, Q. Adsorption of phosphate onto amine functionalized nano-sized magnetic polymer adsorbents: Mechanism and magnetic effects. RSC Adv. 2015, 5, 22080–22090. [Google Scholar] [CrossRef]
  57. Biswas, B.K.; Inoue, J.-I.; Inoue, K.; Ghimire, K.N.; Harada, H.; Ohto, K.; Kawakita, H. Adsorptive removal of As(V) and As(III) from water by a Zr(IV)-loaded orange waste gel. J. Hazard. Mater. 2008, 154, 1066–1074. [Google Scholar] [CrossRef] [PubMed]
  58. Siddiqui, S.I.; Singh, P.N.; Tara, N.; Pal, S.; Chaudhry, S.A.; Sinha, I. Arsenic removal from water by starch functionalized maghemite nano-adsorbents: Thermodynamics and kinetics investigations. Colloid Interface Sci. Commun. 2020, 36, 100263. [Google Scholar] [CrossRef]
  59. Wang, J.; Xu, W.; Chen, L.; Huang, X.; Liu, J. Preparation and evaluation of magnetic nanoparticles impregnated chitosan beads for arsenic removal from water. Chem. Eng. J. 2014, 251, 25–34. [Google Scholar] [CrossRef]
  60. Tahoon, M.; Gomaa, E.; Suleiman, M. Aqueous Micro-hydration of Na+ (H2O)n=1–7 Clusters: DFT Study. Open Chem. 2019, 17, 260–269. [Google Scholar]
  61. Gomaa, E.A.; Tahoon, M.A.; Negm, A. Aqueous micro-solvation of Li + ions: Thermodynamics and energetic studies of Li+-(H2O)n (n = 1–6) structures. J. Mol. Liq. 2017, 241, 595–602. [Google Scholar] [CrossRef]
  62. Gomaa, E.A.; Tahoon, M.A. Ion association and solvation behavior of copper sulfate in binary aqueous–methanol mixtures at different temperatures. J. Mol. Liq. 2016, 214, 19–23. [Google Scholar] [CrossRef]
  63. Ryu, S.-R.; Jeon, E.-K.; Yang, J.-S.; Baek, K. Adsorption of As(III) and As(V) in groundwater by Fe–Mn binary oxide-impregnated granular activated carbon (IMIGAC). J. Taiwan Inst. Chem. Eng. 2017, 72, 62–69. [Google Scholar] [CrossRef]
  64. Shakoor, M.B.; Niazi, N.K.; Bibi, I.; Shahid, M.; Saqib, Z.A.; Nawaz, M.F.; Shaheen, S.M.; Wang, H.; Tsang, D.; Bundschuh, J.; et al. Exploring the arsenic removal potential of various biosorbents from water. Environ. Int. 2019, 123, 567–579. [Google Scholar] [CrossRef] [PubMed]
  65. Shim, J.; Kumar, M.; Mukherjee, S.; Goswami, R. Sustainable removal of pernicious arsenic and cadmium by a novel composite of MnO2 impregnated alginate beads: A cost-effective approach for wastewater treatment. J. Environ. Manag. 2019, 234, 8–20. [Google Scholar] [CrossRef]
  66. Kim, N.; Park, M.; Yun, Y.-S.; Park, D. Removal of anionic arsenate by a PEI-coated bacterial biosorbent prepared from fermentation biowaste. Chemosphere 2019, 226, 67–74. [Google Scholar] [CrossRef]
  67. Pholosi, A.; Naidoo, E.B.; Ofomaja, A.E. Enhanced Arsenic (III) adsorption from aqueous solution by magnetic pine cone biomass. Mater. Chem. Phys. 2019, 222, 20–30. [Google Scholar] [CrossRef]
  68. Boddu, S.; Dulla, J.B.; Alugunulla, V.N.; Khan, A.A. An assessment on removal performance of arsenic with treated Turbinaria vulgaris as an adsorbent: Characterization, optimization, isotherm, and kinetics study. Environ. Prog. Sustain. Energy 2020, 39, 13313. [Google Scholar] [CrossRef]
  69. Joshi, S.; Sharma, M.; Kumari, A.; Shrestha, S.; Shrestha, B. Arsenic Removal from Water by Adsorption onto Iron Oxide/Nano-Porous Carbon Magnetic Composite. Appl. Sci. 2019, 9, 3732. [Google Scholar] [CrossRef] [Green Version]
  70. Chowdhury, T.; Zhang, L.; Zhang, J.; Aggarwal, S. Removal of Arsenic(III) from Aqueous Solution Using Metal Organic Framework-Graphene Oxide Nanocomposite. Nanomaterials 2018, 8, 1062. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Usman, M.; Katsoyiannis, I.; Mitrakas, M.; Zouboulis, A.; Ernst, M. Performance Evaluation of Small Sized Powdered Ferric Hydroxide as Arsenic Adsorbent. Water 2018, 10, 957. [Google Scholar] [CrossRef] [Green Version]
  72. Usman, M.; Zarebanadkouki, M.; Waseem, M.; Katsoyiannis, I.A.; Ernst, M. Mathematical modeling of arsenic(V) adsorption onto iron oxyhydroxides in an adsorption-submerged membrane hybrid system. J. Hazard. Mater. 2020, 400, 123221. [Google Scholar] [CrossRef] [PubMed]
Figure 1. FT-IR (a) and XRD (b) of the synthesized Fe3O4, Fe3O4/CS, and Fe3O4/CS/PEI magnetic nanomaterials.
Figure 1. FT-IR (a) and XRD (b) of the synthesized Fe3O4, Fe3O4/CS, and Fe3O4/CS/PEI magnetic nanomaterials.
Applsci 11 05630 g001
Figure 2. TEM images of Fe3O4 nanoparticles (a), Fe3O4/CS nanocomposite (b), Fe3O4/CS/PEI nanocomposite (c), and the particle size distribution of the nanocomposite Fe3O4/CS/PEI (d).
Figure 2. TEM images of Fe3O4 nanoparticles (a), Fe3O4/CS nanocomposite (b), Fe3O4/CS/PEI nanocomposite (c), and the particle size distribution of the nanocomposite Fe3O4/CS/PEI (d).
Applsci 11 05630 g002aApplsci 11 05630 g002b
Figure 3. Magnetization curves (inset: magnetic separation of Fe3O4/CS/PEI nanocomposite) of the synthesized Fe3O4 NPs, Fe3O4/CS, and Fe3O4/CS/PEI nanomaterials.
Figure 3. Magnetization curves (inset: magnetic separation of Fe3O4/CS/PEI nanocomposite) of the synthesized Fe3O4 NPs, Fe3O4/CS, and Fe3O4/CS/PEI nanomaterials.
Applsci 11 05630 g003
Figure 4. Nitrogen adsorption desorption isotherm of the synthesized Fe3O4 NPs, Fe3O4/CS, and Fe3O4/CS/PEI nanomaterials.
Figure 4. Nitrogen adsorption desorption isotherm of the synthesized Fe3O4 NPs, Fe3O4/CS, and Fe3O4/CS/PEI nanomaterials.
Applsci 11 05630 g004
Figure 5. The adsorbent dosage effect (a) and pH value effect (b) on the adsorption of As(III) and As(V) ions over the surface of Fe3O4/CS/PEI nanocomposite.
Figure 5. The adsorbent dosage effect (a) and pH value effect (b) on the adsorption of As(III) and As(V) ions over the surface of Fe3O4/CS/PEI nanocomposite.
Applsci 11 05630 g005
Figure 6. The contact time effect (a), pseudo first order plot (b), and pseudo second order plot (c) for the adsorption of As (III) and As(V) ions over the surface of Fe3O4/CS/PEI nanocomposite.
Figure 6. The contact time effect (a), pseudo first order plot (b), and pseudo second order plot (c) for the adsorption of As (III) and As(V) ions over the surface of Fe3O4/CS/PEI nanocomposite.
Applsci 11 05630 g006aApplsci 11 05630 g006b
Figure 7. The arsenic initial concentration effect (a), Freundlich’s isotherm fitting (b), and Langmuir’s isotherm fitting (c) for the adsorption of As(III) and As(V) ions over the surface of Fe3O4/CS/PEI nanocomposite.
Figure 7. The arsenic initial concentration effect (a), Freundlich’s isotherm fitting (b), and Langmuir’s isotherm fitting (c) for the adsorption of As(III) and As(V) ions over the surface of Fe3O4/CS/PEI nanocomposite.
Applsci 11 05630 g007aApplsci 11 05630 g007b
Figure 8. The reusability of Fe3O4/CS/PEI magnetic nanomaterials for the removal of As(V) and As(III) ions up to five successive cycles.
Figure 8. The reusability of Fe3O4/CS/PEI magnetic nanomaterials for the removal of As(V) and As(III) ions up to five successive cycles.
Applsci 11 05630 g008
Table 1. Kinetic parameters for the uptake of arsenic ions over the studied magnetic nanocomposite Fe3O4/CS/PEI.
Table 1. Kinetic parameters for the uptake of arsenic ions over the studied magnetic nanocomposite Fe3O4/CS/PEI.
ItemCoefficientPollutant
As(III)As(V)
Pseudo 1st orderR20.6030.630
k1−7.90 × 10−4−7.70 × 10−4
qe (mg/g)5.16.1
Pseudo 2nd orderR20.9990.999
K20.430.35
qe (mg/g)15.9023.70
Freundlich’s isothermR20.9380.943
1/n0.55630.5200
KF2.464.54
Langmuir’s isothermR20.9990.999
RL0.56490.4249
KL (L/mg)0.01580.0394
qmax (mg/g)77.6186.50
Table 2. Comparison of the adsorption capacity of Fe3O4/CS/PEI nanocomposite toward arsenic ions with other materials reported in the literature.
Table 2. Comparison of the adsorption capacity of Fe3O4/CS/PEI nanocomposite toward arsenic ions with other materials reported in the literature.
AdsorbentpH ValueRemoval Capacity (mg/g)Ref.
As(V)As(III)
Fe3O4/CS/PEI6.7086.5077.61This study
Fe–Mn binary oxide-impregnated granular activated carbon4.016.018.0[63]
Egg shell4.108.012.0[64]
Tea waste7.05.07.30[64]
MnO2-impregnated alginate beads6.506.50-[65]
PEI-coated bacterial biosorbent4.063.0-[66]
Magnetic pinecone biomass8.0-18.0[67]
Turbinaria vulgaris sp.4.026.0-[68]
Iron oxide/nanoporous carbon8.0-6.69[69]
MIL-53(Al)-GO nanocomposites9.0-65.0[70]
Dust ferric hydroxide7.96.903.50[71]
Micro-sized tetravalent manganese feroxyhyte8.015.4-[72]
Micro-sized granular ferric hydroxide8.022.4 -[72]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Alsaiari, N.S.; Alzahrani, F.M.; Katubi, K.M.; Amari, A.; Rebah, F.B.; Tahoon, M.A. Polyethylenimine-Modified Magnetic Chitosan for the Uptake of Arsenic from Water. Appl. Sci. 2021, 11, 5630. https://doi.org/10.3390/app11125630

AMA Style

Alsaiari NS, Alzahrani FM, Katubi KM, Amari A, Rebah FB, Tahoon MA. Polyethylenimine-Modified Magnetic Chitosan for the Uptake of Arsenic from Water. Applied Sciences. 2021; 11(12):5630. https://doi.org/10.3390/app11125630

Chicago/Turabian Style

Alsaiari, Norah Salem, Fatimah Mohammed Alzahrani, Khadijah Mohammedsaleh Katubi, Abdelfattah Amari, Faouzi Ben Rebah, and Mohamed A. Tahoon. 2021. "Polyethylenimine-Modified Magnetic Chitosan for the Uptake of Arsenic from Water" Applied Sciences 11, no. 12: 5630. https://doi.org/10.3390/app11125630

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