Next Article in Journal
Comparison of Mechanical Properties of PMMA Disks for Digitally Designed Dentures
Next Article in Special Issue
Numerical Study Using Microstructure Based Finite Element Modeling of the Onset of Convective Heat Transfer in Closed-Cell Polymeric Foam
Previous Article in Journal
Characteristics and Performance of PTU-Cu Composite Membrane Fabricated through Simultaneous Complexation and Non-Solvent Induced Phase Separation
Previous Article in Special Issue
Synthesis of Optically Tunable and Thermally Stable PMMA–PVA/CuO NPs Hybrid Nanocomposite Thin Films
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Polymer-Based Magnetic Nanocomposite for Multi-Pollutants Removal from Water

by
Fatimah Mohammed Alzahrani
1,
Norah Salem Alsaiari
1,*,
Khadijah Mohammedsaleh Katubi
1,*,
Abdelfattah Amari
2,3,*,
Faouzi Ben Rebah
4,5,* and
Mohamed A. Tahoon
4,6
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, Department of Chemical Engineering, Energy and Environment, National School of Engineers, Gabes University, Gabes 6072, Tunisia
4
Department of Chemistry, College of Science, King Khalid University, Abha 61413, Saudi Arabia
5
Higher Institute of Biotechnology of Sfax (ISBS), Sfax University, Sfax 3000, Tunisia
6
Chemistry Department, Faculty of Science, Mansoura University, Mansoura 35516, Egypt
*
Authors to whom correspondence should be addressed.
Polymers 2021, 13(11), 1742; https://doi.org/10.3390/polym13111742
Submission received: 23 April 2021 / Revised: 22 May 2021 / Accepted: 25 May 2021 / Published: 26 May 2021
(This article belongs to the Special Issue Advances in Multifunctional Polymer-Matrix Composites)

Abstract

:
A magnetic polymer-based nanocomposite was fabricated by the modification of an Fe3O4/SiO2 magnetic composite with polypyrrole (PPy) via co-precipitation polymerization to form PPy/Fe3O4/SiO2 for the removal of Congo red dye (CR) and hexavalent chromium Cr(VI) ions from water. The nanocomposite was characterized using various techniques including X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), scanning electron microscope (SEM), vibration sample magnetometer, and thermogravimetric analysis (TGA). The results confirm the successful fabrication of the nanocomposite in the size of nanometers. The effect of different conditions such as the contact time, adsorbent dosage, solution pH, and initial concentration on the adsorption process was investigated. The adsorption isotherm suggested monolayer adsorption of both contaminants over the PPy/Fe3O4/SiO2 nanocomposite following a Langmuir isotherm, with maximum adsorption of 361 and 298 mg.g−1 for CR dye and Cr(VI), respectively. Furthermore, the effect of water type on the adsorption process was examined, indicating the applicability of the PPy/Fe3O4/SiO2 nanocomposite for real sample treatment. Interestingly, the reusability of the nanocomposite for the removal of the studied contaminants was investigated with good results even after six successive cycles. All results make this nanocomposite a promising material for water treatment.

Graphical Abstract

1. Introduction

Industrial wastewater treatment has become increasingly complex in recent decades as a result of the rapid industrialization and the presence of complex mixtures of toxic metal ions and organic dyes that harm human health and the environment [1,2,3]. The textile, paper, and plastic industries are the most common manufacturing purposes including dyes as important aromatic compounds [4]. In addition to their carcinogenic effects, contaminated drinking water with dyes causes various symptoms including severe headaches, skin irritation, and acute diarrhea [5]. Similarly, metal ions are not less dangerous than organic dyes due to their high toxicity when discharged in water supplies [6]. One of these toxic metals, hexavalent chromium (Cr(VI)), is classified as a very toxic and carcinogenic metal, causing nephritis, gastrointestinal ulceration, and cancer in the digestive tract [7]. The accepted concentrations of hexavalent chromium in drinking water and industrial wastewater are 5.0 and 200.0 μg/L, respectively [8]. Therefore, the removal of such pollutants from water has become a challenge for environmental engineers. Several techniques have been applied for the removal of these harmful metallic ions and dyes from water and wastewater such as membrane filtration [9], photocatalytic reduction [10], biological treatment [11], precipitation [12], electrocoagulation [13], and adsorption [14,15].
The adsorption technique is considered as the most effective treatment method due to several reasons such as its high removal efficiency, economic viability, and simple processing, which make it a cost-effective method for pollutant removal from water [16,17,18]. Recently, nanomaterials with exceptional properties [19,20,21,22] have been found to be useful in different fields, especially water treatment. Magnetic nanoparticles as a class of nanomaterials have been widely used in different fields, due to their ease of separation using an external magnet, high specific surface area, and simple modification. Generally, these magnetic nanomaterials include elements with magnetic properties in their chemical structure such as iron, nickel, and cobalt. The exceptional properties of iron oxide (Fe3O4) nanoparticles such as high adsorption, superparamagnetic behavior, good compatibility, low toxicity, high surface energy, and large surface area have attracted attention in recent years. Thus, these nanoparticles are appropriate for the elimination of targeted molecules. The toxicity and safety of using nanoparticles in water and food applications are important points. Interestingly, the non-toxicity of Fe3O4 allowed its wide applications in water and food fields among all other magnetic nanoparticles [23]. To increase the adsorption capacity and efficiency of Fe3O4 nanoparticles, they are usually modified by other organic or inorganic compounds to save the functional groups for capturing the target molecules. The surface of Fe3O4 magnetic nanoparticles can be easily modified using silica (silicon oxide) that has a high surface/volume ratio and, subsequently, can improve the adsorption capacity of Fe3O4 nanoparticles. Silica is widely used in different industries such as ceramics, ferrosilicon production, porcelain, and glassmaking, and as semiconductors in electronics [24]. This makes silica a promising material for different applications. The structural characteristics of silicon oxide allow carrying macromolecules such as proteins and polymers that associate the potential properties of polymers and the silica mechanical stability. This association of different properties is paving the way to numerous technological uses.
Among various polymers, polypyrrole (PPy) has received considerable interest due to its biocompatibility, low cost, excellent chemical stability, and conductivity. Moreover, PPy can be easily synthesized and used in different applications such as the nanomedicine sector, electrochemical sensors, and water treatment. In this context, the use of PPy to adsorb pollutants such as heavy metals has been reported [25,26,27]. Thus, PPy is an excellent choice for the modification of magnetic composites, especially the PPy polymer previously composited with magnetic nanoparticles for different applications such as lithium-ion batteries [28], due to its good compatibility with different nanoparticles.
Therefore, we can conclude that the use of composite materials for water treatment purposes can enhance the chelation and adsorption power of the materials toward targeted molecules. Herein, we synthesized an Fe3O4/SiO2/PPy magnetic nanocomposite for water treatment purposes. In the first part of the paper, we report the synthesis and characterization of the magnetic nanocomposite. In the second part of the paper, we report the application of the synthesized materials to the removal of Cr(VI) and Congo red dye with a discussion of the compositing process’s role in the improvement of the material removal efficiency.

2. Materials and Methods

2.1. Chemicals

Congo red (97%) and K2Cr2O7 (99.9%) were supplied by Sigma-Aldrich (Cairo, Egypt), while the monomer of pyrrole (98.0%) was purchased from Merck Co. (Cairo, Egypt). Tetraethyl orthosilicate (TEOS, 98%), aqueous ammonia (25%), ammonium ferrous sulfate salts (99%), and ferric chloride (98%) were supplied by Sigma-Aldrich (Cairo, Egypt). Hydrochloric acid (35.0%) and sodium hydroxide (97.0%) for the pH adjustment were supplied by El Nasr Co. (Cairo, Egypt).

2.2. Synthesis of Fe3O4/SiO2/PPy Magnetic Nanocomposite

The synthesis of the Fe3O4/SiO2/PPy magnetic nanocomposite was achieved in three steps. The first step included the synthesis of Fe3O4 nanoparticles, which was followed by the second step consisting of the modification of Fe3O4 nanoparticles with SiO2. The third step included the formation of polypyrrole over the Fe3O4/SiO2 composite. For Fe3O4 nanoparticle synthesis, we followed the literature [29] in which hydrochloric acid (HCl, 0.5 M, 100 mL) was used for dissolving ammonia iron sulfate ((NH4)2Fe(SO4)2·6H2O), 7.85g) and ferric chloride (FeCl3, 10.82 g), followed by placing this mixture at 30 °C in an ultrasound bath for half an hour. After that, the temperature of the mixture was adjusted at 80 °C, aqueous ammonia (25%, 60 mL) was added drop-wise, and the solution was stirred at 1500 rpm to form the Fe3O4 nanoparticles as a black precipitate. The stirring process of the mixture was continued up to one hour followed by collecting the nanoparticles via an external magnet, and the collected Fe3O4 nanoparticles were washed several times with distilled water and dried for half a day at 70 °C in an oven. Finally, Fe3O4 nanoparticles were ready to be used. In the second step, Fe3O4 nanoparticles were modified with SiO2 by using tetraethyl orthosilicate (TEOS). This was achieved by using a mixture of distilled H2O (40 mL) and ethyl alcohol (99.9%, 170 mL) to dissolve 2.0 g of synthesized Fe3O4 nanoparticles. Then, after adding 6 mL of aqueous ammonia, this mixture was placed for 15.0 min at 30 °C in an ultrasound bath. After the complete dispersion of Fe3O4 with vigorous stirring up to half a day under argon gas environment, 2 mL of TEOS was added to the solution. The synthesized Fe3O4/SiO2 was then collected by an external magnet, washed several times with absolute ethyl alcohol, and finally dried at 70 °C in an oven for four hours. The final step was the formation of PPy over the synthesized Fe3O4/SiO2 composite via the co-precipitation chemical method. In a glass beaker, 100 mL of distilled water was used to disperse 1.0 g of Fe3O4/SiO2 nanoparticles. For the complete dispersion of nanoparticles, the beaker was placed for 15.0 min at 30 °C in an ultrasound bath. After that, the solution was stirred vigorously and during that, pyrrole (0.1 M) was added and the stirring was continued up to two hours followed by the addition of ferric chloride (0.1 M, 50.0 mL); the reaction was continued for an additional three hours. Finally, the synthesized nanocomposite of PPy/Fe3O4/SiO2 was collected by an external magnet, washed several times with absolute ethyl alcohol, and then dried in an oven at 60.0 °C for 6.0 h.

2.3. Material Characterization

The synthesized materials were characterized using various techniques including X-ray diffraction (XRD), magnetometer, Fourier transform infrared spectroscopy (FT-IR), scanning electron microscope (SEM), and thermogravimetric analysis (TGA) measurement. A GNR APD-2000 PRO (GNR, Cairo, Egypt) diffractometer was used to measure XRD using Cu Ka radiation (λ = 1.5406 Å), operating at 45 kV. The diffraction intensities were recorded over 2θ ranging from 5° to 60° with a constant scanning rate of 1° min−1. A vibrating sample magnetometer (Lake Shore 7410, Lake Shore (Cryotronics Inc., Westerville, OH, USA) was used to measure the magnetization of the nanocomposite. A Bruker, Tensor 27 FT–IR (BRUKER, Karlsruhe, Germany) spectrophotometer was used to perform FT-IR spectra in the range of 400–4000 cm−1 at room temperature and collected these at a resolution of 4 cm−1. An SEM Hitachi S4800 (Hitachi, Tokyo, Japan) was used to study the morphology of the synthesized material. A Perkin Elmer STA 6000 (PerkinElmer Inc., Shelton, USA) was used to measure the thermogravimetric analysis for the evaluation of thermal stability.

2.4. Adsorption Studies

2.4.1. The Effect of Contact Time

The effect of the contact time on the removal of Cr(VI) and Congo red from water using the PPy/Fe3O4/SiO2 nanocomposite was studied using a volume of 100 mL of contaminated water (100 mg.L−1) and adsorbent mass of 0.02 g at pH of 4.0. The solution shaking was conducted up to 24 h with sampling at various times to investigate the remaining contaminants using ICP and a UV–Vis spectrophotometer. The adsorbent was collected after each experiment using a magnet. All experiments were performed in triplicate.

2.4.2. The Effect of Adsorbent Dosage

Different masses of the adsorbent PPy/Fe3O4/SiO2 nanocomposite were used to study the effect of adsorbent dosage on the removal of Cr(VI) and Congo red. The definite mass of adsorbent was mixed with 100.0 mL of polluted water with pH of 4.0 up to 12 h. When equilibrium was reached, the adsorbent was collected with a magnet, and the solution was examined for the presence of Cr(VI) and Congo red. Experiments were performed in triplicate.

2.4.3. The Effect of Solution pH

A pH range of 3.0 to 7.0 was used to investigate the pH effect on the removal of Cr(VI) and Congo red using the PPy/Fe3O4/SiO2 nanocomposite. The adsorbent mass used was 0.05 g, the time was 12.0 h, and the solution volume was 100.0 mL. After each experiment, magnetic separation was used to collect the adsorbent, and the solution was examined for any residual pollutants using ICP and a UV–Vis spectrophotometer (Hach, CO, USA). Experiments were performed in triplicate.

2.4.4. The Adsorption Isotherm

To study the maximum adsorption capacity and the type of the adsorption process of Cr(VI) and Congo red over the synthesized PPy/Fe3O4/SiO2 nanocomposite, the contaminated solution (100.0 mL) was mixed with 0.02 g of the composite for 24.0 h at pH 4.0. The adsorption mechanism was represented by three isotherm models: Freundlich, Langmuir, and Temkin.

2.4.5. The Regeneration Study

To assess the commercial application of an adsorbent, it is necessary to determine the reusability of the material for the effective removal of contaminants several times. This was achieved by performing the adsorption experiment 6.0 successive times by mixing the contaminated water (100 mg.L−1, 100 mL, pH 4.0) with adsorbent (0.05 g) for 12.0 h. After each cycle, the adsorbent was collected with an external magnet, washed with distilled water extensively, and then dried at 50.0 °C for 5.0 h to be used in the next cycle. Experiments were performed in triplicate.

2.4.6. Water Type Effect

Different types of water including wastewater, groundwater, tap water, and distilled water were used to determine the water type effect on the removal of Cr(VI) and Congo red using the synthesized PPy/Fe3O4/SiO2 nanocomposite. Polluted solution prepared using target water with a concentration of 100.0 mg.L−1 and volume of 100.0 mL was mixed with 0.05 g of adsorbent for 12 h. After each experiment, the solution was examined for the presence of contaminants.

3. Results and Discussion

3.1. Nanocomposite Characterization

The synthesized materials including Fe3O4, Fe3O4/SiO2, and PPy/Fe3O4/SiO2 nanomaterials were investigated using different techniques such as X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FT-IR), and scanning electron microscope (SEM). The XRD spectra of Fe3O4 nanoparticles, Fe3O4/SiO2, and PPy/Fe3O4/SiO2 are shown in Figure 1a. The XRD of Fe3O4 nanoparticles alone indicated the cubic phase structure, and the peaks correspond to the (511), (422), (400), (311), and (220) crystalline planes [30]. The XRD of the Fe3O4/SiO2 composite showed the appearance of all Fe3O4 peaks with the reduction in the peak intensity and the disappearance of the peak at 2θ = 20 [31] indicating the crystallinity decrease in Fe3O4 by the combination with amorphous SiO2. The XRD of the PPy/Fe3O4/SiO2 nanocomposite showed the appearance of all Fe3O4 peaks, and there was an overlap between the two peaks of Fe3O4 and PPy at 2θ = 20 that made this peak wider [32]. The diffraction intensity of this peak showed a clearly increase linked to the marked increase in the thickness of the polypyrrole shell, indicating the existence of amorphous polypyrrole in the sample. Additionally, the amorphous structure of SiO2 and PPy caused a reduction in the intensity of Fe3O4 peaks. This indicates the formation of the composite PPy/Fe3O4/SiO2. The XRD results are in agreement with previous studies of Fe3O4, Fe3O4/SiO2, and PPy [31,32,33]. For the investigation of the functional groups present, the FT-IR spectra of Fe3O4 nanoparticles, Fe3O4/SiO2, and PPy/Fe3O4/SiO2 are shown in Figure 1b. The FT-IR spectrum of Fe3O4 nanoparticles showed the appearance of an Fe-O band at 570 cm−1 with the appearance of a hydroxyl group band at 3418 cm−1, due to the existence of H2O in the nanoparticle structure [34]. In the spectra of Fe3O4/SiO2, the peaks appeared at 796 cm−1 and 1074 cm−1 and correspond to an Fe-O-Si bond and a Si-O bond, respectively. Additionally, the same peaks of Fe3O4 nanoparticles corresponding to hydroxyl group and Fe-O are observed with a small shift due to the interaction between Fe3O4 and SiO2, while the FT-IR spectrum of the nanocomposite PPy/Fe3O4/SiO2 showed the existence of all peaks corresponding to Si-O-Fe, Si-O, and Fe-O with a small shift due to the interaction between Fe3O4 and PPy [35]. Additionally, the NOH peak appeared at 934 cm−1 and the C-N peak appeared at 1192 cm−1. Additionally, the C-H vibrations were represented by the peaks at 1052, 1300, and 1298 cm−1 [36]. The pyrrole ring vibrations were represented by the peaks at 1552 and 1547 cm−1. The wide peak observed at the range of 3500–3000 is characteristic of PPy composites, due to the N-H bonds’ large number [28]. Furthermore, the N-H and O-H vibrations were represented by the peaks at 3447 and 3434 cm−1 [28,36]. The results of FT-IR are in excellent agreement with previous studies of Fe3O4, Fe3O4/SiO2 [29], and PPy and its compounds [37].
An advantage of the present nanocomposite is its magnetic properties that allow its separation from the experimental media. Therefore, the magnetization curves of the synthesized materials Fe3O4, Fe3O4/SiO2, and PPy/Fe3O4/SiO2 at room temperature are shown in Figure 1c. According to Figure 1c, the highest saturation magnetization was observed at 68 emu.g−1 for Fe3O4 nanoparticles. This value dropped to 35 emu.g−1 after the modification of Fe3O4 with silica, indicating the effective interaction between Fe3O4 and SiO2 [38]. Moreover, the successful modification of the composite Fe3O4/SiO2 with PPy was indicated by the dropped value of saturation magnetization to 8 emu.g−1, indicating that the SiO2 layer was thinner than the PPy layer [39]. It can be noticed that all synthesized materials have good magnetic properties, allowing magnetic separation using an external magnet, as shown in (Figure 1c, inset).
The thermal stability of the synthesized nanomaterials (Fe3O4, Fe3O4/SiO2, and PPy/Fe3O4/SiO2) was investigated using thermogravimetric analysis (TGA), as shown in Figure 1d. According to the TGA curve of Fe3O4, there is a minor weight loss (in the temperature range of room temperature to 10,000 °C) equal to about 3.0% that is attributed to the evaporation of the chemical or physical attached water or hydroxyl groups on the surface of Fe3O4 nanoparticles. The TGA curve of Fe3O4/SiO2 indicated the successful deposition of thermally stable SiO2 on the Fe3O4 surface as there was only 3.5% weight loss in the temperature from room temperature to 161 °C, which is attributed to the evaporation of adsorbed water. The TGA curve of the PPy/Fe3O4/SiO2 nanocomposite showed 5.0% weight loss at the temperature range from room temperature to 300.0 °C, which is attributed to the removal of the PPy monomer and the evaporation of adsorbed water. Ppy chains have good thermal stability and can be destroyed at high temperatures only. This fact is indicated by the sharp peak at 640.0 °C and weak peak at 400.0 to 500.0 °C. The low thermal stability of the PPy layer resulted in higher weight loss of PPy/Fe3O4/SiO2 compared to Fe3O4/SiO2. The TGA results indicate the good incorporation of PPy and silica in the synthesized PPy/Fe3O4/SiO2 nanocomposite.
For determining the surface structure and morphology of the synthesized materials, the SEM images were assessed (Figure 2).
The SEM image of Fe3O4 nanoparticles indicated their granular shape with a size range of 40.0 to 100.0 nm (in Figure 1a). Similar granular shapes with no big differences in porosity, morphology, and structure were observed between the Fe3O4/SiO2 composite (Figure 2b) and Fe3O4 nanoparticles. However, the morphology of the particles was changed by the addition of PPy, as shown in Figure 2c. The particles became cabbage in shape with an increase in their size as expected due to the deposition of pyrrole over Fe3O4 firstly, followed by the synthesis of PPy, causing the size increase. The present results are in agreement with the previous studies of Fe3O4 nanoparticles, Fe3O4/SiO2 composites [29], and PPy composites [40] from the view of particle shape, morphology, and porosity.

3.2. Adsorption Studies

3.2.1. Effect of Contact Time

The effect of the contact time on the removal of Cr(VI) and Congo red dye from water using the magnetic PPy/Fe3O4/SiO2 nanocomposite was studied, as shown in Figure 3a. According to Figure 3a, there were two stages of removal of both pollutants. The first stage showed a rapid removal of Cr(VI) and Congo red dye that appeared in the plot as a linear curve. The second stage showed stable removal of Cr(VI) and Congo red dye, which appeared in the plot as a plateau curve, meaning that the adsorption process reached equilibrium.
This behavior can be explained in that the first stage occurred due to the existence of a large number of empty adsorption sites on the adsorbent surface, allowing the rapid capture of pollutants within a short period. After that, the adsorption sites became filled, and equilibrium was achieved [41,42]. According to Figure 3a, the Cr(VI) and Congo red dye removal by the magnetic PPy/Fe3O4/SiO2 nanocomposite increased from 8.0 to 193.0 mg.g−1 and from 17.0 to 213.0 mg.g−1, respectively, when the contact time increased from 5.0 to 1440.0 min. As clearly shown in the contact time plot, 480.0 min was the point at which equilibrium was reached for the removal of Cr(VI) and Congo red dye using the PPy/Fe3O4/SiO2 nanocomposite.

3.2.2. Effect of Adsorbent Dosage

Figure 3b shows the effect of adsorbent dosages on the removal of Cr(VI) and Congo red dye from water using the PPy/Fe3O4/SiO2 nanocomposite. This was achieved using different masses of the nanocomposite as follows: 0.06, 0.05, 0.04, 0.03, 0.02, and 0.01 g. According to Figure 3b, the increased dose of the nanocomposite led to an increase in the removal percentage of both pollutants. This behavior is attributed to the existence of accessible and sufficient adsorption sites for the uptake of Cr(VI) and Congo red, which are increased by adding more quantities of the adsorbent, as reported in previous studies [43]. According to the Figure 3b, the removal efficiencies of Cr(VI) and Congo red were increased from 16.0% to 98.0% and from 19.0% to 99.5%, respectively, when the mass of the PPy/Fe3O4/SiO2 nanocomposite increased from 0.01 to 0.06 g. When the adsorbent mass increased from 0.05 to 0.06 g, there was no significant increase in the removal efficiencies. Therefore, 0.05 g was considered the optimum adsorbent dose for the removal of both contaminants using the PPy/Fe3O4/SiO2 nanocomposite. This behavior may be attributed to the saturation of the adsorption sites or the agglomeration of adsorbent particles when their quantity increased, leading to the constant removal efficiency [44].

3.2.3. Effect of Solution pH

The pH effect on the removal of Cr(VI) and Congo red using the PPy/Fe3O4/SiO2 nanocomposite was investigated in the range of 3.0 to 8.0, as shown in Figure 3c. According to Figure 3c, the highest removal efficiencies were achieved at the lowest pH (3.0 to 4.0) with removal values of 99.0% and 99.7% for Cr(VI) and Congo red, respectively. Moreover, the pH increase allowed a removal efficiency decrease for both Cr(VI) and Congo red. This behavior is attributed to many factors such as the charge in adsorbates and the adsorbent surface charge in the water environment. The increased removal efficiency at very low pH values is attributed to the high electrostatic attractions between the negatively charged adsorbates (acidic CR dye and Cr(VI)) and the protonated adsorbent functional groups. The decreased removal efficiency at a higher pH value is attributed to the existence of negatively charged hydroxyl groups that compete with adsorbate molecules for adsorption active sites. Additionally, Cr2O72− and HCrO4 are the predominant ions of chromium in the pH range of (2.0 to 6.0), while this form changes by a pH increase to HCrO42−. It is well known that the HCrO4 form of chromium ions has the lowest adsorption energy equal to (0.60 to 2.50 kcal.mol−1) [45] that explains their high removal at a lower pH. Moreover, at a low pH, H+ ions neutralize the negatively charged adsorbent surface, thereby reducing the barrier to diffusion of dichromate ions. However, the pH effect may be controlled by the development of an electric double layer over the adsorbent. With the increase in pH, the concentration of H+ ions changes from acidic to basic, and, consequently, the polarity of the double layer at the adsorbent surface may be changed from positive to negative. At a lower pH, the system attained equilibrium faster and also the percentage of chromium adsorbed increased, as explained by Verma et al. [46]. Finally, we can conclude that the adsorption behavior at different pH values is dependent on the protonation or deprotonation of the adsorbent functional groups which detect the attraction or repulsion forces between these functional groups and metallic ions and dyes. Therefore, the adsorbent active binding sites and surface chemistry of the material are clearly influenced by the pH, which affects the efficiency of the adsorption.

3.2.4. Initial Concentration Effect

Figure 3d shows the effect of using different Congo red dye and Cr(VI) initial concentrations on their adsorption over the PPy/Fe3O4/SiO2 nanocomposite. This was achieved by using a concentration range of 25.0–300.0 mg.L−1 for both Congo red dye and Cr(VI). According to Figure 3d, the adsorption capacity increased by increasing the initial concentration of Congo red dye and Cr(VI), and at a certain point, equilibrium was reached (the adsorption increase became slower). Equilibrium was reached at 100.0 mg.L−1 for both Congo red dye and Cr(VI), which is considered the optimum initial concentration. The maximum adsorption capacities reached 207.0 mg.L−1 and 257.0 mg.L−1 for Cr(VI) and Congo red dye, respectively. These highest values of pollutant removal at higher concentrations could be explained on the basis of the high attraction forces between pollutants and the adsorbent surface at high concentrations.

3.2.5. Adsorption Isotherm

Freundlich, Langmuir, and Temkin adsorption isotherms were used to investigate the adsorption isotherm of Congo red dye and Cr(VI) ions over the surface of the PPy/Fe3O4/SiO2 nanocomposite. The main purpose of studying the adsorption isotherm is to understand the adsorption mechanism over the studied adsorbent (i.e., assess the adsorption limit and the distribution of the pollutants on the adsorbent sites). The fitting of isotherms is presented in Figure 4.
The first studied isotherm was the Freundlich model that suggested heterogeneous adsorption of contaminants over the surface of the adsorbent, and the adsorption was achieved via the construction of multilayers from the pollutant ions [47]. The Freundlich model can be represented by the equation
qe = KF + CeN,
where qe denotes the amount of uptake pollutants after equilibration (mg.g−1), Ce denotes the rest concentrations (mg.L−1), and N and KF are the normal constants of the Freundlich model and related to the adsorption intensities and capacities, respectively.
The second studied isotherm was the Langmuir model that suggested homogenous adsorption of contaminants over the surface of the adsorbent, and the adsorption was achieved via the construction of a monolayer from the pollutant ions [48]. The Langmuir model can be represented by the equation
qe = KLqm/(1 + KLCe),
where KL denotes the Langmuir constant and qm denotes maximum adsorption after reaching saturation.
The last model is the Temkin isotherm that indicates the effect of sorbate ions on each other during the adsorption process. The Temkin model can be represented by the equation
qe = B lnA + B lnCe,
where B and A denote Temkin model constants. Table 1 summarizes the Freundlich and Langmuir parameters for the removal of Congo red dye and Cr(VI) ions over the PPy/Fe3O4/SiO2 nanocomposite.
According to the data in Table 1, there was a lower fit of the Freundlich and Temkin isotherms than the Langmuir isotherm for the adsorption of Congo red dye and Cr(VI) metal on the surface of the PPy/Fe3O4/SiO2 nanocomposite, as indicated from the values of R2. The maximum adsorption capacities of the PPy/Fe3O4/SiO2 nanocomposite calculated using the Langmuir model were found to be 298.22 mg.g−1 and 361.43 mg.g−1 for Cr(VI) and Congo red dye, respectively. Subsequently, the adsorption of Cr(VI) and Congo red dye on the surface of the PPy/Fe3O4/SiO2 nanocomposite was achieved in the form of monolayer uptake on the energetically symmetrical (homogeneous) adsorption sites. Additionally, the values of N obtained from fitting of the experimental data with the Freundlich isotherm were higher than zero, indicating the favorable adsorption of both pollutants. Additionally, the experimental data also fitted the Temkin model, indicating that the adsorption process was possibly affected by pollutant ion interactions.

3.2.6. The Regeneration Study

The application of any studied adsorbent at a large scale for wastewater treatment essentially requires an investigation of its regeneration [49,50]. Concerning this, the reusability of the PPy/Fe3O4/SiO2 nanocomposite for the removal of Cr(VI) ions and Congo red dye was studied for up to six successive cycles, as shown in Figure 5a.
After each cycle, the adsorbent was separated using an external magnet, washed several times extensively with distilled water to remove any adhered pollutants, and then dried in an oven to be used in the next adsorption–desorption cycle. The studied PPy/Fe3O4/SiO2 nanocomposite showed good reusability for the removal of Cr(VI) and Congo red dye during the studied cycles, with a decrease in the efficiencies by increasing the cycles. The composite’s efficiency to remove Congo red dropped from 98.0% to 74.0%, while for Cr(VI) removal, it dropped from 93.0% to 72.0% after six reuse cycles. This behavior is attributed to the destroyed adsorption sites over the nanocomposite surface after each cycle, causing the decrease in the material efficiency to capture the pollutants.

3.2.7. The Water Type Effect

The removal efficiency of Cr(VI) ions and Congo red dye on the surface of the PPy/Fe3O4/SiO2 nanocomposite using different types of water including sewage water, groundwater, tap water, and distilled water was investigated to determine the applicability of the synthesized nanocomposite for an effective real water treatment (Figure 5b). The order of the removal efficiency for both pollutants (Cr(VI) and Congo red dye) was distilled water > tap water > groundwater > sewage water. This resulting behavior is attributed to the existence of other contaminants in each water type that compete with the studied pollutants for the adsorption active sites over the nanocomposite surface. Furthermore, the selectivity of any innovative adsorbent should be studied toward different competing co-ions including bicarbonate, sulfate, and chloride ions that may be present in water. For example, it was reported that sulphate and chloride negatively affect the adsorption of Cr(VI) on polypyrrole@magnetic chitosan nanocomposites [51]. Additionally, the similarity of the ionic charge between Cr(VI) and bicarbonate ions showed a competitive effect on the adsorption of Cr(VI) [51,52]. Therefore, it is very important to determine the chemical composition of the water before starting the adsorption experiments. Matched to the reported concentrations in nature, these removal efficiencies were achieved for a pollutant concentration of 100.0 mg.L−1, which is considered high. This fact approves the applicability of the PPy/Fe3O4/SiO2 nanocomposite for real water treatment.

3.2.8. Comparative Study

The assessment of the synthesized nanocomposite for the removal of Congo red dye and Cr(VI) ions required its comparison with previously reported adsorbents. In this regard, we summarized the previously studied adsorbents for the removal of Congo red dye and Cr(VI) ions in Table 2. According to Table 2 [53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68], it is noticeable that the present synthesized PPy/Fe3O4/SiO2 nanocomposite showed a good removal capacity for the removal of Cr(VI) ions and Congo red dye when compared to previously reported materials. However, the difference in the adsorbent efficiency between materials may be controlled by the variability of the interactions between pollutants and the material functional groups. Strong interactions with specific functional groups present on the adsorbent surface allow a high adsorption capacity [67]. Furthermore, the competitive potential application of these innovative adsorbents should be evaluated taking into consideration different parameters associated with the material characteristics (regeneration, degradation, life cycle, etc.) and with the wastes resulting from the process (loaded pollutant disposal, chemicals used for the adsorption–desorption, etc.). Moreover, a cost comparison between these materials should be conducted.

4. Conclusions

In this study, a promising polymer-based nanocomposite, PPy/Fe3O4/SiO2, was synthesized for the elimination of Congo red dye and Cr(VI) from water. The synthesized nanocomposite was characterized using different techniques (XRD, FT-IR, SEM, etc.). XRD results confirmed the crystalline structure of Fe3O4 nanoparticles that decreased by the addition of PPy and silica. Additionally, SEM images confirmed the size of the synthesized materials. FT-IR bands approved the successful combination between three components of the hybrid Fe3O4, SiO2, and PPy. The synthesized nanocomposite was examined for the removal of Cr(VI) and Congo red dye from water. The effects of different factors including the contact time, adsorbent dosage, solution pH, and initial concentration on the adsorption process were studied to detect the optimum conditions. The adsorption data were found to be more fitting to the Langmuir model with maximum adsorption of 298.22 mg.g−1 and 361.43 mg.g−1 for the removal of Cr(VI) and Congo red, respectively. This indicated the monolayer adsorption of Cr(VI) and Congo red on the energetically homogeneous active sites of the PPy/Fe3O4/SiO2 nanocomposite. The regeneration study indicated the ability to reuse the nanocomposite several times for the removal of Cr(VI) and Congo red, which may reduce the overall cost of the treatment. Finally, we can conclude that the PPy/Fe3O4/SiO2 nanocomposite is a promising material for water treatment and must be examined in the future for additional pollutant removal.

Author Contributions

Conceptualization, F.M.A., N.S.A., K.M.K., A.A., F.B.R., and M.A.T.; methodology, F.M.A., N.S.A., K.M.K., A.A., F.B.R., and M.A.T.; formal analysis, F.M.A., N.S.A., K.M.K., A.A., F.B.R., and M.A.T.; data curation, F.M.A., N.S.A., K.M.K., A.A., F.B.R., and M.A.T.; writing—original draft preparation, F.M.A., N.S.A., K.M.K., A.A., F.B.R., and M.A.T.; writing—review and editing, F.M.A., N.S.A., K.M.K., A.A., F.B.R., and M.A.T.; visualization, F.B.R. 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

Data sharing is not applicable.

Acknowledgments

The authors extend their appreciation to the Deanship of Scientific Research at King Khalid University, Abha-KSA, for funding this work through the General Research Project under grant number: GRP/144/42. 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. Siddeeg, S.M.; Tahoon, M.A.; Alsaiari, N.S.; Shabbir, M.; Ben Rebah, F. Application of Functionalized Nanomaterials as Effective Adsorbents for the Removal of Heavy Metals from Wastewater: A Review. Curr. Anal. Chem. 2021, 17, 4–22. [Google Scholar] [CrossRef]
  2. 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]
  3. Amari, A.; Alzahrani, F.M.; Mohammedsaleh Katubi, K.; Alsaiari, N.S.; Tahoon, M.A.; Ben Rebah, F. Clay-Polymer Nanocomposites: Preparations and Utilization for Pollutants Removal. Materials 2021, 14, 1365. [Google Scholar] [CrossRef] [PubMed]
  4. Shindhal, T.; Rakholiya, P.; Varjani, S.; Pandey, A.; Ngo, H.H.; Guo, W.; Ng, H.Y.; Taherzadeh, M.J. A critical review on advances in the practices and perspectives for the treatment of dye industry wastewater. Bioengineered 2021, 12, 70–87. [Google Scholar] [CrossRef] [PubMed]
  5. Lin, J.; Zhang, X.; Li, Z.; Lei, L. Biodegradation of Reactive blue 13 in a two-stage anaerobic/aerobic fluidized beds system with a Pseudomonas sp. isolate. Bioresour. Technol. 2010, 101, 34–40. [Google Scholar] [CrossRef] [PubMed]
  6. Yoo, J.-W.; Cho, H.; Lee, K.-W.; Won, E.-J.; Lee, Y.-M. Combined effects of heavy metals (Cd, As, and Pb): Comparative study using conceptual models and the antioxidant responses in the brackish water flea. Comp. Biochem. Physiol. Part C Toxicol. Pharmacol. 2021, 239, 108863. [Google Scholar] [CrossRef] [PubMed]
  7. Suljević, D.; Sulejmanović, J.; Fočak, M.; Halilović, E.; Pupalović, D.; Hasić, A.; Alijagic, A. Assessing hexavalent chromium tissue-specific accumulation patterns and induced physiological responses to probe chromium toxicity in Coturnix japonica quail. Chemosphere 2021, 266, 129005. [Google Scholar] [CrossRef]
  8. Shaban, M.; Abukhadra, M.R.; Khan, A.A.P.; Jibali, B.M. Removal of Congo red, methylene blue and Cr (VI) ions from water using natural serpentine. J. Taiwan Inst. Chem. Eng. 2018, 82, 102–116. [Google Scholar] [CrossRef]
  9. Liu, J.; Xiong, J.; Tian, C.; Gao, B.; Wang, L.; Jia, X. The degradation of methyl orange and membrane fouling behavior in anaerobic baffled membrane bioreactor. Chem. Eng. J. 2018, 338, 719–725. [Google Scholar] [CrossRef]
  10. Qiao, X.-Q.; Zhang, Z.-W.; Li, Q.-H.; Hou, D.; Zhang, Q.; Zhang, J.; Li, D.-S.; Feng, P.; Bu, X. In situ synthesis of n–n Bi 2 MoO 6 & Bi 2 S 3 heterojunctions for highly efficient photocatalytic removal of Cr (VI). J. Mater. Chem. A 2018, 6, 22580–22589. [Google Scholar]
  11. Feng, L.; Chen, W.-M.; Li, J.-L.; Day, G.; Drake, H.; Joseph, E.; Zhou, H.-C. Biological antagonism inspired detoxification: Removal of toxic elements by porous polymer networks. ACS Appl. Mater. Interfaces 2019, 11, 14383–14390. [Google Scholar] [CrossRef]
  12. Wang, R.; Ng, D.H.; Liu, S. Recovery of nickel ions from wastewater by precipitation approach using silica xerogel. J. Hazard. Mater. 2019, 380, 120826. [Google Scholar] [CrossRef] [PubMed]
  13. Pan, C.; Troyer, L.D.; Catalano, J.G.; Giammar, D.E. Dynamics of chromium (VI) removal from drinking water by iron electrocoagulation. Environ. Sci. Technol. 2016, 50, 13502–13510. [Google Scholar] [CrossRef]
  14. Katubi, K.M.M.; Alsaiari, N.S.; Alzahrani, F.M.; M Siddeeg, S.; A Tahoon, M. Synthesis of Manganese Ferrite/Graphene Oxide Magnetic Nanocomposite for Pollutants Removal from Water. Processes 2021, 9, 589. [Google Scholar] [CrossRef]
  15. Amari, A.; Katubi, K.M.; Alsaiari, N.S.; Alzahrani, F.M.; Ben Rebah, F.; Tahoon, M.A. Magnetic metal organic framework immobilized laccase for wastewater decolorization. Processes 2021, 9, 774. [Google Scholar] [CrossRef]
  16. Delpiano, G.R.; Tocco, D.; Medda, L.; Magner, E.; Salis, A. Adsorption of Malachite Green and Alizarin Red S Dyes Using Fe-BTC Metal Organic Framework as Adsorbent. Int. J. Mol. Sci. 2021, 22, 788. [Google Scholar] [CrossRef]
  17. Ihlenburg, R.B.; Lehnen, A.-C.; Koetz, J.; Taubert, A. Sulfobetaine Cryogels for Preferential Adsorption of Methyl Orange from Mixed Dye Solutions. Polymers 2021, 13, 208. [Google Scholar] [CrossRef] [PubMed]
  18. Luca, P.D.; Chiodo, A.; Macario, A.; Siciliano, C.; B Nagy, J. Semi-Continuous Adsorption Processes with Multi-Walled Carbon Nanotubes for the Treatment of Water Contaminated by an Organic Textile Dye. Appl. Sci. 2021, 11, 1687. [Google Scholar] [CrossRef]
  19. Alsaiari, N.S.; Katubi, K.M.M.; Alzahrani, F.M.; Siddeeg, S.M.; Tahoon, M.A. The Application of Nanomaterials for the Electrochemical Detection of Antibiotics: A Review. Micromachines 2021, 12, 308. [Google Scholar] [CrossRef] [PubMed]
  20. Amari, A.; Al Mesfer, M.K.; Alsaiari, N.S.; Danish, M.; Alshahrani, A.M.; Tahoon, M.A.; Ben Rebah, F. Electrochemical and Optical Properties of Tellurium Dioxide (TeO2) Nanoparticles. Int. J. Electrochem. Sci. 2021, 16, 210235. [Google Scholar] [CrossRef]
  21. Amari, A.; Alalwan, B.; Siddeeg, S.M.; Tahoon, M.A.; Alsaiari, N.S.; Ben Rebah, F. Biomolecules Behavior on a Surface of Boron Doped/un-doped Graphene Nanosheets. Int. J. Electrochem. Sci. 2020, 15, 11427–11436. [Google Scholar] [CrossRef]
  22. Siddeeg, S.M.; Alsaiari, N.S.; Tahoon, M.A.; Ben Rebah, F. 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]
  23. Wang, X.-Y.; Jiang, X.-P.; Li, Y.; Zeng, S.; Zhang, Y.-W. Preparation Fe3O4 chitosan magnetic particles for covalent immobilization of lipase from Thermomyces lanuginosus. Int. J. Biol. Macromol. 2015, 75, 44–50. [Google Scholar] [CrossRef] [PubMed]
  24. Jeelani, P.G.; Mulay, P.; Venkat, R.; Ramalingam, C. Multifaceted application of silica nanoparticles. A review. Silicon 2020, 12, 1337–1354. [Google Scholar] [CrossRef]
  25. Cho, Y.; Shi, R.; Ivanisevic, A.; Borgens, R.B. A mesoporous silica nanosphere-based drug delivery system using an electrically conducting polymer. Nanotechnology 2009, 20, 275102. [Google Scholar] [CrossRef] [PubMed]
  26. Dong, J.; Hu, Y.; Xu, J.; Qu, X.; Zhao, C. Nanocomposite with Polypyrrole Encapsulated within SBA-15 Mesoporous Silica: Preparation and Its Electrochemical Application. Electroanal. Int. J. Devoted Fundam. Pract. Asp. Electroanal. 2009, 21, 1792–1798. [Google Scholar] [CrossRef]
  27. Ansari, R.; Fahim, N.K. Application of polypyrrole coated on wood sawdust for removal of Cr (VI) ion from aqueous solutions. React. Funct. Polym. 2007, 67, 367–374. [Google Scholar] [CrossRef]
  28. Zhao, J.; Zhang, S.; Liu, W.; Du, Z.; Fang, H. Fe3O4/PPy composite nanospheres as anode for lithium-ion batteries with superior cycling performance. Electrochim. Acta 2014, 121, 428–433. [Google Scholar] [CrossRef]
  29. Pirsa, S.; Aghbolagh Sharifi, K. A review of the applications of bioproteins in the preparation of biodegradable films and polymers. J. Chem. Lett. 2020, 1, 47–58. [Google Scholar]
  30. Pirsa, S.; Asadzadeh, F.; Sani, I.K. Synthesis of Magnetic Gluten/Pectin/Fe3O4 Nano-hydrogel and Its Use to Reduce Environmental Pollutants from Lake Urmia Sediments. J. Inorg. Organomet. Polym. Mater. 2020, 30, 3188–3198. [Google Scholar] [CrossRef]
  31. Wang, Y.; Peng, X.; Shi, J.; Tang, X.; Jiang, J.; Liu, W. Highly selective fluorescent chemosensor for Zn 2+ derived from inorganic-organic hybrid magnetic core/shell Fe 3 O 4@ SiO 2 nanoparticles. Nanoscale Res. Lett. 2012, 7, 1–13. [Google Scholar]
  32. Wang, J.-G.; Wei, B.; Kang, F. Facile synthesis of hierarchical conducting polypyrrole nanostructures via a reactive template of MnO 2 and their application in supercapacitors. Rsc Adv. 2014, 4, 199–202. [Google Scholar] [CrossRef]
  33. Ayad, M.; Salahuddin, N.; Fayed, A.; Bastakoti, B.P.; Suzukic, N.; Yamauchi, Y. Chemical design of a smart chitosan-polypyrrole–magnetite nanocomposite toward efficient water treatment. Phys. Chem. Chem. Phys. 2014, 16, 21812. [Google Scholar] [CrossRef]
  34. Cheng, Y.; Gao, F.; An, L.; Li, X.; Wang, G. Different combinations of Fe3O4 microsphere, Polypyrrole and silver as core–shell nanocomposites for adsorption and photocatalytic application. Adv. Powder Technol. 2014, 25, 1600–1607. [Google Scholar] [CrossRef]
  35. Wang, G.; Chang, Y.; Wang, L.; Liu, C. Synthesis, characterization and microwave absorption properties of Fe3O4/Co core/shell-type nanoparticles. Adv. Powder Technol. 2012, 23, 861–865. [Google Scholar] [CrossRef]
  36. Aigbe, U.; Ho, W.; Maity, A.; Khenfouch, M.; Srinivasu, V. Removal of hexavalent chromium from wastewater using PPy/Fe3O4 magnetic nanocomposite influenced by rotating magnetic field from two pole three-phase induction motor. J. Phys. Conf. Ser. 2018, 948, 012008. [Google Scholar] [CrossRef] [Green Version]
  37. Fu, Y.; Su, Y.-S.; Manthiram, A. Sulfur-polypyrrole composite cathodes for lithium-sulfur batteries. J. Electrochem. Soc. 2012, 159, A1420. [Google Scholar] [CrossRef]
  38. Abbas, M.; Abdel-Hamed, M.; Chen, J. Efficient one-pot sonochemical synthesis of thickness-controlled silica-coated superparamagnetic iron oxide (Fe 3 O 4/SiO 2) nanospheres. Appl. Phys. A 2017, 123, 775. [Google Scholar] [CrossRef]
  39. Abbas, M.; Torati, S.R.; Kim, C. A novel approach for the synthesis of ultrathin silica-coated iron oxide nanocubes decorated with silver nanodots (Fe 3 O 4/SiO 2/Ag) and their superior catalytic reduction of 4-nitroaniline. Nanoscale 2015, 7, 12192–12204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Pirsa, S.; Alizadeh, N. Nanoporous conducting polypyrrole gas sensor coupled to a gas chromatograph for determination of aromatic hydrocarbons using dispersive liquid–liquid microextraction method. IEEE Sens. J. 2011, 11, 3400–3405. [Google Scholar] [CrossRef]
  41. Bahador, F.; Foroutan, R.; Esmaeili, H.; Ramavandi, B. Enhancement of the chromium removal behavior of Moringa oleifera activated carbon by chitosan and iron oxide nanoparticles from water. Carbohydr. Polym. 2021, 251, 117085. [Google Scholar] [CrossRef] [PubMed]
  42. Rambabu, K.; Bharath, G.; Banat, F.; Show, P.L. Biosorption performance of date palm empty fruit bunch wastes for toxic hexavalent chromium removal. Environ. Res. 2020, 187, 109694. [Google Scholar] [CrossRef] [PubMed]
  43. Hsini, A.; Benafqir, M.; Naciri, Y.; Laabd, M.; Bouziani, A.; Ez-zahery, M.; Lakhmiri, R.; El Alem, N.; Albourine, A. Synthesis of an arginine-functionalized polyaniline@ FeOOH composite with high removal performance of hexavalent chromium ions from water: Adsorption behavior, regeneration and process capability studies. Colloids Surf. A Physicochem. Eng. Asp. 2021, 617, 126274. [Google Scholar] [CrossRef]
  44. Hsini, A.; Naciri, Y.; Laabd, M.; El Ouardi, M.; Ajmal, Z.; Lakhmiri, R.; Boukherroub, R.; Albourine, A. Synthesis and characterization of arginine-doped polyaniline/walnut shell hybrid composite with superior clean-up ability for chromium (VI) from aqueous media: Equilibrium, reusability and process optimization. J. Mol. Liq. 2020, 316, 113832. [Google Scholar] [CrossRef]
  45. Tahar, L.B.; Oueslati, M.H.; Abualreish, M.J.A. Synthesis of magnetite derivatives nanoparticles and their application for the removal of chromium (VI) from aqueous solutions. J. Colloid Interface Sci. 2018, 512, 115–126. [Google Scholar] [CrossRef]
  46. Verma, A.; Chakraborty, S.; Basu, J.K. Adsorption study of hexavalent chromium using tamarind hull-based adsorbents. Sep. Purif. Technol. 2006, 50, 336–341. [Google Scholar] [CrossRef]
  47. Hsini, A.; Naciri, Y.; Benafqir, M.; Ajmal, Z.; Aarab, N.; Laabd, M.; Navío, J.; Puga, F.; Boukherroub, R.; Bakiz, B. Facile synthesis and characterization of a novel 1, 2, 4, 5-benzene tetracarboxylic acid doped polyaniline@ zinc phosphate nanocomposite for highly efficient removal of hazardous hexavalent chromium ions from water. J. Colloid Interface Sci. 2021, 585, 560–573. [Google Scholar] [CrossRef]
  48. Dognani, G.; Hadi, P.; Ma, H.; Cabrera, F.C.; Job, A.E.; Agostini, D.L.; Hsiao, B.S. Effective chromium removal from water by polyaniline-coated electrospun adsorbent membrane. Chem. Eng. J. 2019, 372, 341–351. [Google Scholar] [CrossRef]
  49. Siddeeg, S.M.; Amari, A.; Tahoon, M.A.; Alsaiari, N.S.; Ben Rebah, F. 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]
  50. Chigondo, M.; Paumo, H.K.; Bhaumik, M.; Pillay, K.; Maity, A. Magnetic arginine-functionalized polypyrrole with improved and selective chromium (VI) ions removal from water. J. Mol. Liq. 2019, 275, 778–791. [Google Scholar] [CrossRef]
  51. Alsaiari, N.S.; Amari, A.; Katubi, K.M.; Alzahrani, F.M.; Ben Rebah, F.; Tahoon, M.A. Innovative magnetite based polymeric nanocomposite for water treatment. Processes 2021, 9, 576. [Google Scholar] [CrossRef]
  52. Karthikeyan, P.; Elanchezhiyan, S.S.; Preethi, J.; Meenakshi, S.; Park, C.M. Mechanistic performance of polyaniline-substituted hexagonal boron nitride composite as a highly efficient adsorbent for the removal of phosphate, nitrate, and hexavalent chromium ions from an aqueous environment. Appl. Surf. Sci. 2020, 511, 145543. [Google Scholar] [CrossRef]
  53. Sun, W.; Zhou, Y.; Su, Q.; Chen, L.; Wang, Y.; Liu, J.; Sun, Y.; Ma, H. Removal of chromium (VI) from aqueous solutions using polypyrrole-based magnetic composites. Polym. Bull. 2017, 74, 1157–1174. [Google Scholar] [CrossRef]
  54. Wang, H.; Yuan, X.; Wu, Y.; Chen, X.; Leng, L.; Wang, H.; Li, H.; Zeng, G. Facile synthesis of polypyrrole decorated reduced graphene oxide–Fe3O4 magnetic composites and its application for the Cr (VI) removal. Chem. Eng. J. 2015, 262, 597–606. [Google Scholar] [CrossRef]
  55. Ballav, N.; Choi, H.J.; Mishra, S.B.; Maity, A. Polypyrrole-coated halloysite nanotube clay nanocomposite: Synthesis, characterization and Cr (VI) adsorption behaviour. Appl. Clay Sci. 2014, 102, 60–70. [Google Scholar] [CrossRef]
  56. Bhaumik, M.; Maity, A.; Srinivasu, V.; Onyango, M.S. Removal of hexavalent chromium from aqueous solution using polypyrrole-polyaniline nanofibers. Chem. Eng. J. 2012, 181, 323–333. [Google Scholar] [CrossRef]
  57. Ballav, N.; Maity, A.; Mishra, S.B. High efficient removal of chromium (VI) using glycine doped polypyrrole adsorbent from aqueous solution. Chem. Eng. J. 2012, 198, 536–546. [Google Scholar] [CrossRef]
  58. Bhaumik, M.; Maity, A.; Srinivasu, V.; Onyango, M.S. Enhanced removal of Cr (VI) from aqueous solution using polypyrrole/Fe3O4 magnetic nanocomposite. J. Hazard. Mater. 2011, 190, 381–390. [Google Scholar] [CrossRef]
  59. Ballav, N.; Choi, H.; Mishra, S.; Maity, A. Synthesis, characterization of Fe3O4@ glycine doped polypyrrole magnetic nanocomposites and their potential performance to remove toxic Cr (VI). J. Ind. Eng. Chem. 2014, 20, 4085–4093. [Google Scholar] [CrossRef]
  60. Amalraj, A.; Selvi, M.K.; Rajeswari, A.; Pius, A. Preparation and characterization of aspartic acid doped polypyrrole for the efficient removal of Cr (VI) from aqueous solution. J. Water Process Eng. 2016, 11, 162–173. [Google Scholar] [CrossRef]
  61. Zhang, Z.; Zhu, W.; Wang, R.; Zhang, L.; Zhu, L.; Zhang, Q. Ionothermal confined self-organization for hierarchical porous magnesium borate superstructures as highly efficient adsorbents for dye removal. J. Mater. Chem. A 2014, 2, 19167–19179. [Google Scholar] [CrossRef]
  62. Lorenc-Grabowska, E.; Gryglewicz, G. Adsorption characteristics of Congo Red on coal-based mesoporous activated carbon. Dye. Pigment. 2007, 74, 34–40. [Google Scholar] [CrossRef]
  63. Zhao, J.; Tan, Y.; Su, K.; Zhao, J.; Yang, C.; Sang, L.; Lu, H.; Chen, J. A facile homogeneous precipitation synthesis of NiO nanosheets and their applications in water treatment. Appl. Surf. Sci. 2015, 337, 111–117. [Google Scholar] [CrossRef]
  64. Li, X.; Xiao, W.; He, G.; Zheng, W.; Yu, N.; Tan, M. Pore size and surface area control of MgO nanostructures using a surfactant-templated hydrothermal process: High adsorption capability to azo dyes. Colloids Surf. A Physicochem. Eng. Asp. 2012, 408, 79–86. [Google Scholar] [CrossRef]
  65. Bhattacharyya, K.G.; Sharma, A. Azadirachta indica leaf powder as an effective biosorbent for dyes: A case study with aqueous Congo Red solutions. J. Environ. Manag. 2004, 71, 217–229. [Google Scholar] [CrossRef]
  66. Zhai, Y.; Zhai, J.; Zhou, M.; Dong, S. Ordered magnetic core–manganese oxide shell nanostructures and their application in water treatment. J. Mater. Chem. 2009, 19, 7030–7035. [Google Scholar] [CrossRef]
  67. Wang, L.; Wang, A. Adsorption characteristics of Congo Red onto the chitosan/montmorillonite nanocomposite. J. Hazard. Mater. 2007, 147, 979–985. [Google Scholar] [CrossRef] [PubMed]
  68. Li, Z.; Hanafy, H.; Zhang, L.; Sellaoui, L.; Netto, M.S.; Oliveira, M.L.S.; Seliem, M.K.; Dotto, G.L.; Bonilla-Petriciolet, A.; Li, Q. Adsorption of congo red and methylene blue dyes on an ashitaba waste and a walnut shell-based activated carbon from aqueous solutions: Experiments, characterization and physical interpretations. Chem. Eng. J. 2020, 388, 124263. [Google Scholar] [CrossRef]
  69. Cseri, L.; Topuz, F.; Abdulhamid, M.A.; Alammar, A.; Budd, P.M.; Szekely, G. Electrospun Adsorptive Nanofibrous Membranes from Ion Exchange Polymers to Snare Textile Dyes from Wastewater. Adv. Mater. Technol. 2021, 2000955. [Google Scholar] [CrossRef]
Figure 1. XRD (a), FT-IR (b), magnetization curves (c) (inset: separation of magnetic nanocomposite after adsorption with an external magnet), and thermogravimetric analysis (TGA) (d) of synthesized Fe3O4 nanoparticles, Fe3O4/SiO2 nanocomposite, and PPy/Fe3O4/SiO2 nanocomposite.
Figure 1. XRD (a), FT-IR (b), magnetization curves (c) (inset: separation of magnetic nanocomposite after adsorption with an external magnet), and thermogravimetric analysis (TGA) (d) of synthesized Fe3O4 nanoparticles, Fe3O4/SiO2 nanocomposite, and PPy/Fe3O4/SiO2 nanocomposite.
Polymers 13 01742 g001
Figure 2. SEM images of Fe3O4 nanoparticles (a), Fe3O4/SiO2 nanocomposite (b), and PPy/Fe3O4/SiO2 nanocomposite (c).
Figure 2. SEM images of Fe3O4 nanoparticles (a), Fe3O4/SiO2 nanocomposite (b), and PPy/Fe3O4/SiO2 nanocomposite (c).
Polymers 13 01742 g002
Figure 3. The contact time effect (a), adsorbent dosage effect (b), pH effect (c), and initial concentration effect (d) on the removal of Congo red dye and Cr(VI) from water using PPy/Fe3O4/SiO2 nanocomposite.
Figure 3. The contact time effect (a), adsorbent dosage effect (b), pH effect (c), and initial concentration effect (d) on the removal of Congo red dye and Cr(VI) from water using PPy/Fe3O4/SiO2 nanocomposite.
Polymers 13 01742 g003
Figure 4. Plotting of Freundlich isotherm (a), Langmuir isotherm (b), and Temkin isotherm (c) for the removal of Congo red dye and Cr(VI) metal ions over the PPy/Fe3O4/SiO2 nanocomposite.
Figure 4. Plotting of Freundlich isotherm (a), Langmuir isotherm (b), and Temkin isotherm (c) for the removal of Congo red dye and Cr(VI) metal ions over the PPy/Fe3O4/SiO2 nanocomposite.
Polymers 13 01742 g004aPolymers 13 01742 g004b
Figure 5. The regeneration study of PPy/Fe3O4/SiO2 nanocomposite for the removal of Congo red dye and Cr(VI) up to six successive cycles (a) and the effect of water type (b).
Figure 5. The regeneration study of PPy/Fe3O4/SiO2 nanocomposite for the removal of Congo red dye and Cr(VI) up to six successive cycles (a) and the effect of water type (b).
Polymers 13 01742 g005
Table 1. The calculated Langmuir and Freundlich isotherm parameters for the removal of Cr(VI) and CR dye over PPy/Fe3O4/SiO2 nanocomposite.
Table 1. The calculated Langmuir and Freundlich isotherm parameters for the removal of Cr(VI) and CR dye over PPy/Fe3O4/SiO2 nanocomposite.
PollutantLangmuirFreundlichTemkin
qmax (mg.g−1)KL (L.mg.g−1)R2NKF (L.mg.g−1)R2B (J.mol−1)A (L.g−1)R2
CR dye361.430.0170.9460.4872.13600.84269.444.330.931
Cr(VI)298.220.0130.8920.5371.27200.78557.124.520.835
Table 2. Comparative assessment of CR dye and Cr(VI) ions adsorption capacity of PPy/Fe3O4/SiO2 nanocomposite with previously reported adsorbents.
Table 2. Comparative assessment of CR dye and Cr(VI) ions adsorption capacity of PPy/Fe3O4/SiO2 nanocomposite with previously reported adsorbents.
AdsorbentPollutantqm (mg.g−1)Ref.
PPy/Fe3O4/SiO2CR dye and Cr(VI)361.43 and 298.22This study
PPy/Fe3O4/AgClCr(VI)111[52]
PPy-rGO/Fe3O4Cr(VI)227[53]
PPy-coated halloysite nanotubesCr(VI)149[54]
PPy-PANI fibersCr(VI)227[55]
Glycine-doped PPyCr(VI)217[56]
PPy/Fe3O4Cr(VI)169[57]
Fe3O4 glycine-doped PPyCr(VI)238[58]
Aspartic acid-doped PPyCr(VI)177[59]
Hierarchical porous MgBO2(OH) microspheresCR dye228[60]
Mesoporous activated carbonCR dye189[61]
NiO nanosheetsCR dye168[62]
MgO powdersCR dye105[63]
Neem leaf powder2CR dye41[64]
Magnetic core–manganese oxide shellCR dye42[65]
Chitosan/montmorillonite nanocompositeCR dye55[66]
Ashitaba waste-based activated carbonsCR dye289–381[67]
Walnut shell-based activated carbonsCR dye314–400[68]
Nanofibrous membranes from ion polymersCR dye70.8[69]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Alzahrani, F.M.; Alsaiari, N.S.; Katubi, K.M.; Amari, A.; Ben Rebah, F.; Tahoon, M.A. Synthesis of Polymer-Based Magnetic Nanocomposite for Multi-Pollutants Removal from Water. Polymers 2021, 13, 1742. https://doi.org/10.3390/polym13111742

AMA Style

Alzahrani FM, Alsaiari NS, Katubi KM, Amari A, Ben Rebah F, Tahoon MA. Synthesis of Polymer-Based Magnetic Nanocomposite for Multi-Pollutants Removal from Water. Polymers. 2021; 13(11):1742. https://doi.org/10.3390/polym13111742

Chicago/Turabian Style

Alzahrani, Fatimah Mohammed, Norah Salem Alsaiari, Khadijah Mohammedsaleh Katubi, Abdelfattah Amari, Faouzi Ben Rebah, and Mohamed A. Tahoon. 2021. "Synthesis of Polymer-Based Magnetic Nanocomposite for Multi-Pollutants Removal from Water" Polymers 13, no. 11: 1742. https://doi.org/10.3390/polym13111742

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