Next Article in Journal
Remarkable Separation of Carbofuran Pesticide from Aqueous Solution Using Free Metal Ion Variation on Aluminum-Based Metal-Organic Framework
Next Article in Special Issue
Boosting Water Oxidation Activity via Carbon–Nitrogen Vacancies in NiFe Prussian Blue Analogue Electrocatalysts
Previous Article in Journal
Emulsifiers from White Beans: Extraction and Characterization
Previous Article in Special Issue
An Overview of Coacervates: The Special Disperse State of Amphiphilic and Polymeric Materials in Solution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fabrication of Polystyrene/AlOOH Hybrid Material for Pb(II) Decontamination from Wastewater: Isotherm, Kinetic, and Thermodynamic Studies

Department of Environmental Sciences, Faculty of Meteorology, Environment and Arid Land Agriculture, King Abdulaziz University, Jeddah 21589, Saudi Arabia
Colloids Interfaces 2022, 6(4), 72; https://doi.org/10.3390/colloids6040072
Submission received: 21 October 2022 / Revised: 16 November 2022 / Accepted: 21 November 2022 / Published: 24 November 2022
(This article belongs to the Special Issue Colloids Science in Asia)

Abstract

:
The nanomaterials’ toxicity to aquatic life is a big issue due to improper handling or incomplete separation after use. The immobilization of the nanomaterials in the polymeric matrix could be a practical approach to developing an efficient hybrid composite for wastewater purification. In this study, AlOOH nanoparticles were immobilized in the polystyrene polymeric matrix to prepare an effective adsorbent to scavenge the Pb(II) from the aqueous solution. The synthesized polystyrene/AlOOH (PS/AlOOH) hybrid was characterized using microscopic techniques coupled with elemental mapping and EDX, X-ray diffraction, and a furrier-transformed infrared spectrometer. The results revealed that the Pb(II) adsorption onto the polystyrene/AlOOH composite depends on the solution pH, the Pb(II) concentrations in the solution, the adsorption time, and the solute temperature. The maximum scavenging of Pb(II) occurs at pH 6 in 90 min. The adsorption of Pb(II) onto PS/AlOOH decreases from 97.7% to 58.5% with the increase in the Pb(II) concentration from 20 mg g−1 to 100 mg g−1. The kinetics and isotherm modeling demonstrated that Pb(II) adsorption is well suited for the pseudo-second-order kinetics and Toth isotherm models, suggesting that the chemisorption occurs at the heterogeneous surface of PS/AlOOH. The PS/AlOOH composite could be used multiple times without a significant loss in the adsorption efficiency. These results demonstrated that the polystyrene/AlOOH composite is an effective material for the purification of wastewater and can be used on a large scale.

1. Introduction

Water pollution from industrial activities is a severe environmental problem that encircles several organic and inorganic pollutants. However, organic pollutants could be biodegraded, while metals accumulate in living organisms, are non-biodegradable, and have adverse effects. Lead is widely used in battery manufacturing, printing, and painting and is a significant source of lead pollution. Lead is risky in low doses (the WHO recommended amount is 10 ppb), as it accumulates in body parts such as the brain, liver, bones, and kidneys [1]. Therefore, influential technology for Pb(II) scavenging from wastewater is needed to protect the environment.
In recent decades, chemical precipitation, phytoremediation, ion exchange, membrane-based methods, and many other techniques have been explored to scavenge contaminants [2]. The applications of these methods are limited or have some shortcomings, including the high energy consumption, money investment, and large volumes of sludge. Among them, adsorption is one of the most promising technologies for removing Pb(II) because of the ease of operation and comparatively low application cost. In the last few years, a different class of materials has been studied for the scavenging of organic and inorganic pollutants, such as carbons [3], agricultural waste [4,5], polymeric adsorbents [6], metal oxides [1,7], etc. Very few of them expressed good adsorption properties. Inorganic solids such as AlOOH, silica, iron oxides, etc. are well-characterized adsorbents widely used to remove charged metallic ions. AlOOH is well known for its environment-friendly nature. Various natural materials such as mica, kaolin, kyanite, and so forth have been investigated for the decontamination of the pollutants in wastewater [8,9]. The straightforward synthesis, high stability, high surface area, and low cost of AlOOH make it one of the most demanding materials for decontaminating heavy metals from wastewater. Keshtkar [10] investigated the adsorption capability of γ-alumina for the Ni(II) and reported 99.6% scavenging. The recovery of the Cd(II) using γ-alumina/β-cyclodextrin has been investigated by Esfanjani [11]. The authors declare that experimental conditions are essential for Cd(II) recovery, and 94.15% of Cd(II) was recovered at pH 9. The existing –OH groups on the alumina surface are the key factor for the high adsorption of the metal ions [12]. The hydrophilic nature and fine or nano size of AlOOH particles are difficult to handle in an aqueous medium. During the fixed bed column study or high-pressure flow system experiments, nanoparticles that are not correctly fixed on the bed may show inconsistent results. The problem can be fixed by preparing the hybrid adsorbents by irreversibly dispersing inorganic nanoparticles within different polymeric matrices [13].
Polystyrene-based polymeric ion exchangers such as Amberlite and Dowex have been widely applied to decontaminate toxic metallic pollutants from aqueous solutions for environmental pollution control. [14,15]. The high thermal stability, reusability, and recycling nature of polystyrene-based materials make them the most promising material for water purification applications. Heavy metal removal by polystyrene-based ion-exchangers is usually accomplished via ion exchange, chelation, or complex formation with the metal ions [16]. Few recent studies have been published on the polystyrene microplastic adsorption capabilities of metallic contaminants from the aquatic medium [17,18]. Thus, polystyrene as a polymer support for preparing an organic–inorganic hybrid adsorbent could be a good option for removing toxic metal ions from industrial wastewater.
The present work prepared a polystyrene/AlOOH (PS/AlOOH) hybrid adsorbent using simple immobilizing AlOOH nanoparticles in the polystyrene matrix. The application of the PS/AlOOH hybrid was investigated for the scavenging of Pb(II) from an aqueous medium. The adsorption kinetics, isotherm models, and thermodynamics studies were performed to identify the Pb(II) adsorption mechanism onto the PS/AlOOH hybrid.

2. Materials and Method

2.1. Materials

Lead nitrate Pb(NO3)2 (98.2%) and aluminum nitrate Al(NO3)3·9H2O (98%) were bought from Panreac Quimica S.A. Acetone (99.5%) was obtained from PanReac Applichem, ITW reagents. Polystyrene beads were supplied by BDH Ltd. A fixed amount of lead nitrate was dissolved in de-ionized water to make the standard Pb(II) solution.

2.2. Instrumentation

Scanning electron microscopy (SEM) images of AlOOH and PS/AlOOH were recorded on JSM7600F, JEOL. The X-ray diffraction (XRD) patterns of AlOOH, PS/AlOOH, and Pb(II)-adsorbed PS/AlOOH were recorded on ALTIMA-IV, RIGAKU spectroscopy. The transmission electron microscopy (TEM) imaginings of the PS/AlOOH hybrid adsorbent were recorded on a JEOL 200CX electron microscope. The Fourier transform infrared spectroscopy (FTIR) spectra of PS/AlOOH and Pb(II)-adsorbed PS/AlOOH in the wavelength range of 400–4000 cm−1 were acquired on a 100 FTIR Perkin Elmer Spectrum spectrophotometer. The amount of Pb(II) in the solution was analyzed by a DR6000 UV-visible spectrophotometer using a HACH LCK 306 standard kit.

2.3. Preparation of Polystyrene/AlOOH Hybrid

Herein, a sol-gel method was used to synthesize AlOOH from aluminum nitrate as a source of aluminum. Initially, 10 g of aluminum nitrate was dissolved in 100 mL of de-ionized water. After that, 9 mL of ammonia solution (28%) was added dropwise under stirring. A white precipitate was formed and left to stir for 16 h. The white product was thoroughly washed with de-ionized water and ethanol and dried at 105 °C for 18 h.
The PS/AlOOH hybrid adsorbent was synthesized using the modified method reported elsewhere [19] by mixing the PS and AlOOH in an equal ratio (W/W). Initially, 2.5 g of AlOOH was dispersed in 75 mL of acetone and sonicated for 30 min. After that, 2.5 g of PS beads were mixed in an acetone solution and stirred for 24 h. The PS beads became soft and formed a gel-like structure, and AlOOH was blended with the PS. After that, the PS-blended AlOOH was isolated from the acetone and dried at 70 °C for 4 h to evaporate the acetone. The obtained bulk PS/AlOOH hybrid was ground to make the fine powder and sieved (50–150 mesh size). The prepared PS/AlOOH hybrid was used for the Pb(II) adsorption studies.

2.4. Batch Adsorption Studies

The adsorption efficiency of PS/AlOOH for Pb(II) was explored in the batch experiments. A fixed mass of PS/AlOOH (0.05 g) was dispersed in the 20 mL Pb(II) solutions of known concentration ranges from 20 to 100 mg/L at a continual shaking speed of 200 rpm in a water bath shaker at 32 °C. The experiment was performed in the interaction time range from 0 to 150 min for the equilibrium time analysis by taking 0.05 g of PS/ALOOH mixed with 20 mL of Pb(II) solution of 100 mg L−1 concentration. The lead ions present as Pb(II) at a low concentration and a pH of less than 6. Therefore, the solution pH was adjusted from 2.0 to 6.0 to avoid the hydrolysis product formation of lead, and 0.01 M HCl and/or 0.1 M NaOH were used to set the desired pH. In contrast, the formation of Pb(II) hydrolysis products, such as PbOH+ and Pb(OH)2, occurs at a pH greater than 6.0, which might lead to its precipitation [20]. The Pb(II) concentration in the supernatant solution was analyzed using a DR-6000 spectrophotometer. The Pb(II) uptake capacity per unit mass of PS/ALOOH at any time was estimated as follows:
qt = (C0 − Ce)V/M
where qt is the Pb(II) adsorption capacity of PS/ALOOH in (mg g−1) at time t. C0 (initial) and Ce (equilibrium) are the Pb(II) concentrations in mg L−1. V and M represent the used Pb(II) solution volume (L) and PS/ALOOH mass (g), respectively.

2.5. Desorption Studies

The spent PS/ALOOH adsorbent was regenerated by the chemical treatment method to make the process more economical. For the regeneration, 0.05 g of completely exhausted PS/AlOOH was mixed with 20 mL of an eluent. Water, 0.1N NaCl, and 0.1M HCl were used as the eluent. After the Pb(II) desorption in the eluent, the used PS/AlOOH was washed with water and used to scavenge the Pb(II). This cycle was repeated three times.

3. Results and Discussion

3.1. Synthesis and Characterization PS/AlOOH Hybrid Adsorbent

The sol-del method successfully synthesized nanosized AlOOH particles and immobilized them in the polystyrene (PS) matrix. The hypothesis for AlOOH immobilization with PS was to recover the AlOOH nanoparticles after adsorption. The immobilization of the AlOOH nanoparticles in the PS may slow down the motion of the polymeric chain and form an interfacial layer around the AlOOH nanoparticles, which prevents the agglomeration of the hydrophilic AlOOH nanoparticles [19]. A schematic diagram for preparing the PS/AlOOH hybrid is displayed in Figure 1. The synthesized AlOOH and PS/AlOOH hybrid adsorbent was characterized using several instrumental analyses such as XRD, SEM, TEM, elemental mapping, EDX, and FTIR.
The XRD patterns of AlOOH, PS/AlOOH, and Pb(II)-adsorbed PS/AlOOH are presented in Figure 2. The XRD pattern of the AlOOH shows the major characteristic peaks at 18.69°(001), 20.17°(020), 27.81°(120), 40.83°(201), 53.17°(024), and 62.21°(071), indicating the successful synthesis of the AlOOH with slight impurities of the Al2O3, which is formed due to the dehydration of few AlOOH particles. The crystallite size of AlOOH was observed in the range of 20.8 to 48.10 nm. The XRD pattern of the PS/AlOOH hybrid shows a broad peak indicating the amorphous nature of the prepared hybrid material. A similar pattern for the Pb(II)-adsorbed PS/Al₂O₃ hybrid was observed. The amorphous nature of the PS/AlOOH hybrid is due to the polystyrene, a highly amorphous polymer [21].
The surface morphologies of the AlOOH, PS/AlOOH, and Pb(II)-adsorbed PS/Al₂O were recorded on the scanning electron microscopy, and the SEM images are shown in Figure 3. The SEM image of the AlOOH (Figure 3a) shows the agglomerated particles of various shapes and sizes. The AlOOH particles immobilized in the PS matrix show a flake-like morphology (Figure 3b), and embedded AlOOH particles are visible. However, Pb(II)-adsorbed PS/AlOOH (Figure 3c) shows a similar morphology because adsorbed Pb(II) cannot be seen in the microscopy. TEM analysis further explored the morphology and interface between PS and AlOOH. TEM images of the PS/AlOOH (Figure 4a) show that the polymeric matrix contains the AlOOH nanoparticle with a porous morphology. The AlOOH particles are well distributed in the PS polymeric network (Figure 4b).
EDX analysis was performed to confirm the Pb(II) adsorption onto the PS/AlOOH surface, and the EDX spectra are presented in Figure 5a. The EDX analysis shows the existence of the C, O, Al, and Pb elements, with 67.99, 21.51, 8.63, and 1.87% (wt%), respectively. These results reveal the successful adsorption of Pb(II) onto the PS/AlOOH surface. Moreover, to observe the distribution of the elements, elemental mapping was performed, and the mapping images are shown in Figure 5b–f. The mapping images show the good distribution of each element, indicating the good mixing of the PS and AlOOH. The elemental mapping shows the presence of carbon (Figure 5b), aluminum (Figure 5c), oxygen (Figure 5d), and surfaced adsorbed lead (Figure 5e), while Figure 5f shows all the observed elements in the Pb(II)-adsorbed PS/AlOOH hybrid adsorbent. The mapping of the Pb(II)-adsorbed PS/AlOOH shows the good distribution of the Pb(II) particles all over the surface, indicating that all the active sites contributed to the adsorption process.
FTIR analysis of the PS/AlOOH after Pb(II) adsorption was performed to find the presence of the functional groups and their interaction with Pb(II). Figure 6 illustrates the FTIR spectrum of PS/AlOOH before and after Pb(II) adsorption with several adsorption peaks. The peaks at 3446 cm−1 are the hydroxyl group peak of AlOOH and the adsorbed moisture. The peaks that appeared around the wavelength 3059–2851 cm−1 are the aromatic C-H stretching vibrations of PS [22]. The aromatic C=C stretching vibration for PS is recorded at 1601, 1492, and 1453 cm−1, while the C-H bending vibrations are observed at 750 and 698 cm−1. The peak for the AlO-OH showed absorbance at 1384, 1027, and 755 cm−1 [23]. A peak appeared at 1154 cm−1, belonging to the O-H bending in AlO-OH. Moreover, the peaks appearing in the lower wavelength region belong to the Al-O groups in AlOOH [24]. As shown in Figure 6b, the AlO-OH bands shifted after the Pb(II) adsorption, indicating the interaction between the AlO-OH and Pb(II). These results revealed that hydroxyl groups were the leading active site on the PS/AlOOH hybrid, which is involved in Pb(II) adsorption [25].

3.2. Adsorption Kinetics

The rate of Pb(II) uptake from the solution and diffusion to the PS/AlOOH surface is generally explored with the kinetics of the adsorption process. The Pb(II) adsorption rate onto the PS/AlOOH hybrid adsorbent as a function of the interaction time is presented in Figure 7. The results in Figure 7 show that the adsorption of Pb(II) onto PS/AlOOH was fast during the initial few minutes due to the empty adsorption sites on the hybrid polymer. The active sites become occupied with the Pb(II) ions as the interaction time increases, and the complete saturation of PS/AlOOH active sites occurred in 90 min, indicating the equilibrium establishment and that no place would remain vacant on the PS/AlOOH [26].
The Pb(II) adsorption rate onto PS/AlOOH can be investigated by fitting the equilibrium data to non-linear kinetics equations such as pseudo-first-order, pseudo-second-order, and Elovich models. The equation for the Lagergren pseudo-first-order model [27] is as follows:
q t = q e ( 1 e ( k 1 t ) )
where qe and qt are the Pb(II) uptake capacities onto the PS/AlOOH (mg g−1) at saturation and at time t, respectively, while the first-order kinetic model rate constant is represented by k1 min−1. The values of the pseudo-first-order parameters were obtained from the plot of qt against t (Figure 7).
The Pb(II) equilibrium data were also fitted to the non-linear pseudo-second-order model [28], which can be stated as:
q t = q e 2 k 2 t [ k 2 ( q e ) t + 1 ]
where k2 (g mg−1 min−1) is the pseudo-second-order rate constant.
The adsorption kinetic data for Pb(II) removal were also fitted to the Elovich kinetics model, and the non-linear equation for the applied model is as follows:
q t = 1 β ln α β t
where α and β are the Elovich model rate constants.
The analysis of the best-fitted model was evaluated using the error function assessment by measuring the difference between the theoretical and experimental data. The validation of the applied kinetic models was assessed by fitting the root-mean-square error (RMSE) and chi-square (χ2) statistical error functions. The equations for the applied statistical error functions are as follows:
R S M E =   1 N M   i = 1 N ( q e e x p q e c a l ) 2  
Χ 2 = i = 1 N [ ( q e e x p q e c a l ) 2 q e c a l ]
where qeexp and qecal are the experimental and theoretical Pb(II) uptake capacity (mg g−1).
The non-linear fitting of the experimental data to the pseudo-first-order, pseudo-second-order, and Elovich models is displayed in Figure 7a. Kinetics parameters and statistical error values are included in Table 1. The superior values of the R2 indicate the fitting of all applied kinetics models to the data [29]. However, the values of the RMSE and χ2 are the lowest for the pseudo-second-order kinetic model. Moreover, the slightly greater value of the R2 for the pseudo-second-order model indicates that the pseudo-second-order model fits the experimentally investigated data more than the pseudo-first-order and Elovich models. The pseudo-second-order model was best suited to the Pb(II) adsorption data, indicating that the Pb(II) adsorption rate-limiting step is chemisorption via the exchange or/and sharing of the electrons between the metal ions and PS/AlOOH, surface chelation, or ion-exchange. Similar Pb(II) adsorption behavior was also reported on allophane [26] and polystyrene microplastics adsorbent [18].
Since the pseudo-first-order, pseudo-second-order, and Elovich kinetic models provide details only about the rate of Pb(II) adsorption, the Weber and Morris [30] intra-particle mass transfer diffusion model was used to find the distribution of Pb(II) ions on/in the PS/AlOOH surface. The linear equation for the intra-particle diffusion model is as follows
qt = kid t1/2 + C
where C (mg g−1) and kid (mg(g min)−1) represent the intercept and intra-particle diffusion rate constant. A plot of qt versus t1/2 is used to find values of the C and kid. The qt vs. t1/2 (Figure 7b) demonstrates the two straight-line portions. The first part of the curves signifies the boundary layer distribution of the Pb(II) ions on the PS/AlOOH surface, whereas the second line results from the intra-particle diffusion. The intercept value details the distribution of Pb(II) onto the PS/AlOOH surface. The higher intercept values lead to boundary layer adsorption. The intercept value for Pb(II) is 16.33 mg g−1. The result designated that the scavenging of Pb(II) onto PS/AlOOH was controlled through a boundary layer effect [17,18].

3.3. Effect of Solution pH

The scavenging of Pb(II) onto PS/AlOOH was performed at pH 2.0 to 6.0 to investigate the connection between solution pH and uptake capacity. The pH study results shown in Figure 8 indicate that the adsorption of Pb(II) improved with the rise in the solution pH, and the highest Pb(II) adsorption onto PS/AlOOH was noted at pH 6. The adsorbed Pb(II) amount increased rapidly from pH 2.0 to 4.0, and, after that, very slow adsorption was observed. In a highly acidic medium, Pb(II) scavenging was low due to competition between positively charged excess H+ and Pb(II) toward the same binding sites of PS/AlOOH. With the increasing solution pH, the negative charge density on the PS/AlOOH surface increases, which favors the binding of positively charged Pb(II) ions [31,32]. The possible ion exchange and complexation reactions on the surface of PS/AlOOH may occur as follows:
AlO–OH + H → AlO–OH2+ (low pH)
AlO–OH + Pb(II) → (AlO–O–Pb)+ + H+
AlO–O + Pb(II) → (AlO–O–Pb)+
AlO–OH + Pb(II) + H2O → (AlO–O–PbOH) + 2H+

3.4. Effect of Concentration and Adsorption Isotherm

The effect of Pb(II) concentrations in the solution on its scavenging onto PS/AlOOH was investigated in ranges from 20 to 100 mg L–1 to ensure the adsorption equilibrium concentration. As seen in Figure 9, the adsorption improved from 7.6 to 23.2 mg g–1 with the increased metal ion concentration. The adsorption capacity of Pb(II) onto the PS/AlOOH polymeric hybrid increased with the rise in the initial Pb(II) concentration, which might be attributed to an increase in the concentration gradient driving force with the rise in the initial Pb(II) ion concentration [33].
Langmuir, Freundlich, Redlich–Peterson, and Toth adsorption isotherms were applied to find the PS/AlOOH hybrid adsorbent’s adsorption behaviors and surface properties. The non-linear equations for the Langmuir, Freundlich, Redlich–Peterson (R–P), and Toth isotherm models, respectively, can be expressed as follows:
q e = q m k L C e 1 + k L C e
q e = k F C e 1 n
q e = K R P   C e 1 + α R P   C e β
q e = q m   C e ( K T o + C e t n ) 1 / t n
where Ce: Pb(II) at equilibrium concentration, qe: Pb(II) at equilibrium capacity, qm: monolayer adsorption capacity, kF and 1/n: Freundlich constant associated with adsorption intensity and uptake capacity, K R P (L/g) and α R P : Redlich–Peterson isotherm constants, and β: exponent reflecting the heterogeneity of the PS/AlOOH hybrid adsorbent. KTO and tn are the Toth model constant (dimensionless) and exponent (0 < n ≤ 1), respectively. The non-linear plots for the applied isotherm models are included in Figure 9, and the obtained isotherm parameters and their corresponding error function analysis RMSE and χ2 are included in Table 2. The values of the R2, RMSE, and χ2 are very close for the applied models. However, the value of R2 for the Toth model is slightly higher than that for the Langmuir, Freundlich, and R–P isotherm models. Upon further comparison of the error function RMSE and χ2 values, Toth gave the lowest values, while the Langmuir model showed the lowest χ2 values. Empirically, the Toth model is the modified Langmuir equation, which suggests the heterogeneous adsorption of Pb(II) onto the PS/AlOOH hybrid adsorbent [34]. The values of R2, RMSE, and χ2 indicate that the Toth isotherm model predicts the scavenging of Pb(II) onto the PS/AlOOH hybrid adsorbent through the chemisorption.

3.5. Pb(II) Adsorption Thermodynamics

The impact of temperatures on the scavenging of Pb(II) from the aqueous solution onto PS/AlOOH was performed at a temperature range from 305 to 323 K, as shown in Figure 10a. It was noticed that the uptake capacity of PS/AlOOH enhanced very slight from 23.2 to 24 mg g−1 with the increase in the Pb(II) solution temperature. The reduction in solution viscosity and the increase in the Pb(II) ions mobility may cause a slight increase in the metal ions scavenging. Moreover, the deformation of the active sites and diffusion of Pb(II) ions may be increased in PS/ALOOH at higher temperatures. Furthermore, ions are readily dissolved at a higher temperature, so their adsorption becomes more favorable [35].
Thermodynamic examinations were performed to observe the energetic changes related to Pb(II) adsorption onto PS/AlOOH. The following were applied to the thermodynamic data to obtain the values of standard free energy change (ΔG°), enthalpy change (ΔH°), and entropy change (ΔS°):
Δ G° = −RT ln Kc
lnKc = (ΔS°/R) − (ΔH°/RT)
where, Kc, T, and R are the distribution coefficient, temperature (K), and gas constant (8.314 J mol−1 K−1), respectively. The thermodynamic parameters (ΔH° and ΔS°) values were determined from the slope and intercept of the ln Kc vs. 1/T plot, and the values are presented in Table 3. The adsorption of Pb(II) from the aqueous solution onto PS/AlOOH shows negative ΔG° values, implying that Pb(II) scavenging was spontaneous. The positive value of ΔH° confirmed that adsorption was endothermic in nature. Furthermore, Table 3, demonstrating a positive value for ΔS°, specifies that the degrees of freedom enhanced at the solid–liquid interface during Pb(II) adsorption onto PS/AlOOH [36].

3.6. Desorption and Reusability Studies

The Pb(II) desorption from the exhausted PS/AlOOH has been investigated using pure de-ionized water, 0.1M NaCl, and 0.1M HCl. The results demonstrated that de-ionized water and 0.1M NaCl could not break the bonds between Pb(II) and PS/AlOOH, while 0.1M HCl liberated the Pb(II) in the solution. The desorption of the Pb(II) in HCl is due to the ion-exchange process between the H+ ions and Pb(II) [37]. Therefore, 0.1M HCl was used to regenerate the used PS/AlOOH. The results revealed that the uptake capacity of the PS/AlOOH changed from 23.2 mg/g to 18 mg/g after three cycles of reuse, as displayed in Figure 10b. The decline in the uptake capacity might be due to the deactivation of the same active sites and undesorbed Pb(II) ions from the pore of the PS/AlOOH.

3.7. Comparison of Adsorption Performance

The adsorption capacities of various adsorbents have been compared with the PS/AlOOH used for decontaminating Pb(II) from an aqueous medium. Table 4 summarizes the adsorption capacities of multiple adsorbents used to scavenge Pb(II). Comparing the adsorption capacity shows that PS/AlOOH is a moderate hybrid adsorbent material for Pb(II) sequestration from an aqueous solution.

4. Conclusions

A facile strategy has been used herein to synthesize the PS/AlOOH hybrid adsorbent. The prepared PS/AlOOH hybrid adsorbent has been successfully utilized for the sequestration of Pb(II) from an aquatic medium. The XRD, EDX, and elemental mapping analysis evidence the successful binding of Pb(II) onto the PS/AlOOH hybrid adsorbent. The findings demonstrated the fast scavenging of Pb(II) onto the PS/AlOOH hybrid within 90 min, and the most significant adsorption was observed at pH 6. The Pb(II) adsorption kinetic data were best suited for the pseudo-second-order kinetic model, and inter-particle diffusion was the rate-limiting step. The Pb(II) adsorption equilibrium data were applied to the Langmuir, Freundlich, Redlich–Petersen, and Toth isotherm models, but the Toth isotherm model was suited. The maximum monolayer adsorption capacity was 23.611 mg g−1. The kinetic and isotherm results revealed that the chemisorption and the ion-exchanged mechanism were involved in the Pb(II) scavenging. The thermodynamic study demonstrated that Pb(II) removal by the PS/AlOOH hybrid was not affected much by the change in the solution temperature. Regeneration studies proved that HCl was the best eluent to desorb the Pb(II) from the saturated PS/AlOOH hybrid, and after three times of reuse, the PS/AlOOH hybrid showed a reasonable adsorption capacity. Based on the findings, the PS/AlOOH hybrid can be used as an adsorbent for scavenging pollutants from wastewater.

Funding

This research received no external funding.

Data Availability Statement

All the data included in the article or available on demand.

Acknowledgments

Author gratefully acknowledge the support from the Deanship of Scientific Research and Department of Environmental Sciences, King Abdulaziz University, Jeddah, Saudi Arabia.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Asencios, Y.J.O.; Sun-Kou, M.R. Synthesis of high-surface-area γ-Al2O3 from aluminum scrap and its use for the adsorption of metals: Pb(II), Cd(II) and Zn(II). App. Sur. Sci. 2012, 258, 10002–10011. [Google Scholar] [CrossRef]
  2. Somma, S.; Reverchon, E.; Baldino, L. Water Purification of Classical and Emerging Organic Pollutants: An Extensive Review. ChemEngineering 2021, 5, 47. [Google Scholar] [CrossRef]
  3. Abdelkhalek, A.; El-Latif, M.A.; Ibrahim, H.; Hamad, H.; Showman, M. Controlled synthesis of graphene oxide/silica hybrid nanocomposites for removal of aromatic pollutants in water. Sci. Rep. 2022, 12, 7060. [Google Scholar] [CrossRef] [PubMed]
  4. Bayuo, J.; Rwiza, M.; Mtei, K. A comprehensive review on the decontamination of lead(II) from water and wastewater by low-cost biosorbents. RSC Adv. 2022, 12, 11233. [Google Scholar] [CrossRef] [PubMed]
  5. Afroze, S.; Sen, T.K. A Review on Heavy Metal Ions and Dye Adsorption from Water by Agricultural Solid Waste Adsorbents. Water Air Soil Pollut. 2018, 229, 225. [Google Scholar] [CrossRef]
  6. Wang, Q.; Zhu, S.; Xi, C.; Zhang, F. A Review: Adsorption and Removal of Heavy Metals Based on Polyamide-amines Composites. Front. Chem. 2022, 10, 814643. [Google Scholar] [CrossRef]
  7. Gupta, K.; Joshi, P.; Gusain, R.; Khatri, O.P. Recent advances in adsorptive removal of heavy metal and metalloid ions by metal oxide-based nanomaterials. Coord. Chem. Rev. 2021, 445, 214100. [Google Scholar] [CrossRef]
  8. Nosrati, S.; Seifpanahi-Shabani, K.; Karamoozian, M. Novel polymorphous aluminosilicate nano minerals: Preparation, characterization and dyes wastewater treatment. Korean J. Chem. Eng. 2017, 34, 2406–2417. [Google Scholar] [CrossRef]
  9. Alhassan, S.I.; Hung, L.; He, Y.; Yan, L.; Wu, B.; Wang, H. Fluoride removal from water using alumina and aluminum-based composites: A comprehensive review of progress. Cri. Rev. Environ. Sci. Technol. 2021, 51, 2051–2085. [Google Scholar] [CrossRef]
  10. Keshtkar, Z.; Tamjidi, S.; Vaferi, B. Intensifying nickel (II) uptake from wastewater using the synthesized γ-alumina: An experimental investigation of the effect of nano-adsorbent properties and operating conditions. Environ. Technol. Innov. 2021, 22, 101439. [Google Scholar] [CrossRef]
  11. Esfanjani, L.; Farhadyar, N.; Shahbazi, H.R.; Fathi, F. Development of a method for cadmium ion removal from the water using nano γ-alumina/β-cyclodextrin. Toxicol. Rep. 2021, 8, 1877–1882. [Google Scholar] [CrossRef] [PubMed]
  12. Koju, N.K.; Song, X.; Wang, Q.; Hu, Z.; Colombo, C. Cadmium removal from simulated groundwater using alumina nanoparticles: Behaviors and mechanisms. Environ. Pollut. 2018, 240, 255–266. [Google Scholar] [CrossRef] [PubMed]
  13. Adeola, A.O.; Nomngongo, P.N. Advanced Polymeric Nanocomposites for Water Treatment Applications: A Holistic Perspective. Polymers 2022, 14, 2462. [Google Scholar] [CrossRef] [PubMed]
  14. Mohammadi, L.; Rahdar, L.; Khaksefidi, A.; Ghamkhari, R.; Fytianos, A.; Kyazas, G.Z. Polystyrene Magnetic Nanocomposites as Antibiotic Adsorbents. Polymers 2020, 12, 1313. [Google Scholar] [CrossRef] [PubMed]
  15. Erol, H.B.U.; Hestekin, C.N.; Hestekin, J.A. Effects of Resin Chemistries on the Selective Removal of Industrially Relevant Metal Ions Using Wafer-Enhanced Electrodeionization. Membranes 2021, 11, 45. [Google Scholar] [CrossRef]
  16. Pan, B.; Zhang, W.; Lv, L.; Zhang, Q.; Zheng, S. Development of polymeric and polymer-based hybrid adsorbents for pollutants removal from waters. J. Chem. Eng. 2009, 151, 19–29. [Google Scholar] [CrossRef]
  17. Lin, Z.; Hu, Y.; Yuan, Y.; Hu, B.; Wang, B. Comparative analysis of kinetics and mechanisms for Pb(II) sorption onto three kinds of microplastics. Ecotoxicol. Environ. 2021, 208, 111451. [Google Scholar] [CrossRef]
  18. Lu, X.; Zeng, F.; Wei, S.; Hu, B.; Wang, B. Effects of humic acid on Pb2+ adsorption onto polystyrene microplastics from spectroscopic analysis and site energy distribution analysis. Sci. Rep. 2022, 12, 8932. [Google Scholar] [CrossRef]
  19. Mortezaei, M.; Famili, M.H.N.; Kalaee, M.R. Effect of immobilized interfacial layer on the maximum filler loading of polystyrene/silica nanocomposites. J. Reinf. Plast. Compos. 2011, 30, 593–599. [Google Scholar] [CrossRef]
  20. Gupta, V.K.; Agarwal, S.; Saleh, T.A. Synthesis and characterization of alumina-coated carbon nanotubes and their application for lead removal. J. Hazard. Mater. 2011, 185, 17–23. [Google Scholar] [CrossRef]
  21. Suresh, K.; Kumar, R.V.; Pugazhenthi, G. Processing and characterization of polystyrene nanocomposites based on CoAl layered double hydroxide. J. Sci. Adv. Mater. Dev. 2016, 1, 351–361. [Google Scholar] [CrossRef] [Green Version]
  22. Fang, J.; Xuan, Y.; Li, Q. Preparation of polystyrene spheres in different particle sizes and assembly of the PS colloidal crystals. Sci. China Technol. Sci. 2010, 53, 3088–3093. [Google Scholar] [CrossRef]
  23. Prabhakar, R.; Samadder, S.R. Use of adsorption-influencing parameters for designing the batch adsorber and neural network–based prediction modelling for the aqueous arsenate removal using combustion synthesised nano-alumina. Environ. Sci. Pollut. Res. 2020, 27, 26367–26384. [Google Scholar] [CrossRef]
  24. Zhang, H.; Li, P.; Cui, W.; Liu, C.; Wang, S.; Zheng, S.; Zhang, Y. Synthesis of nanostructured γ-AlOOH and its accelerating behavior on the thermal decomposition of AP. RSC Adv. 2016, 6, 27235–27241. [Google Scholar] [CrossRef]
  25. Gohari, R.J.; Lau, W.J.; Matsuura, T.; Halakoo, E.; Ismail, A.F. Adsorptive removal of Pb(II) from aqueous solution by novel PES/HMO ultrafiltration mixed matrix membrane. Sep. Puri. Tech. 2013, 120, 59–68. [Google Scholar] [CrossRef]
  26. Akartasse, N.; Azzaoui, K.; Mejdoubi, E.; Elansari, L.L.; Hammouti, B.; Siaj, M.; Jodeh, S.; Hanbali, G.; Hamed, R.; Rhazi, L. Chitosan-Hydroxyapatite Bio-Based Composite in Film Form: Synthesis and Application in Wastewater. Polymers 2022, 14, 4265. [Google Scholar] [CrossRef]
  27. Lagergen, S. Zur theorie der sogenannten adsorption geloster stoffe. K. Sven. Vetensk. Handl. 1898, 24, 1–39. [Google Scholar]
  28. Ho, Y.S.; McKay, G. Pseudo-second-order model for sorption processes. Proc. Biochem. 1999, 34, 451–465. [Google Scholar] [CrossRef]
  29. Kostoglou, M.; Karapantsios, T.D. Why Is the Linearized Form of Pseudo-Second Order Adsorption Kinetic Model So Successful in Fitting Batch Adsorption Experimental Data? Colloids Interfaces 2022, 6, 55. [Google Scholar] [CrossRef]
  30. Weber, W.J.; Morris, J.C. Kinetics of adsorption on carbon from solution. J. Sanit. Eng. Div. Am. Soc. Civ. Eng. 1963, 89, 31–60. [Google Scholar] [CrossRef]
  31. Sun, B.; Li, X.; Zhao, R.; Yin, M.; Wang, Z.; Jiang, Z.; Wang, C. Hierarchical aminated PAN/γ–AlOOH electrospun composite nanofibers and their heavy metal ion adsorption performance. J. Taiwan Institute Chem. Eng. 2016, 62, 219–227. [Google Scholar] [CrossRef]
  32. Bowen, C.; Yhui, W.; Fan, Z.; Yujie, W.; Cheng, C.; Xiaojuan, J.; Xin, X. Synthesis of porous AlOOH nanospindle via hydrothermal route and its application in Pb(II) removal. Mater. Res. Express 2019, 6, 125057. [Google Scholar]
  33. Wu, F.C.; Tseng, R.L.; Juang, R.S. Kinetic modeling of liquid-phase adsorption of reactive dyes and metal ions on chitosan. Water Res. 2001, 35, 613–618. [Google Scholar] [CrossRef]
  34. Ayawei, N.; Ebelegi, A.N.; Wankasi, D. Modelling and Interpretation of Adsorption Isotherms. J. Chem. 2017, 11, 3039817. [Google Scholar] [CrossRef] [Green Version]
  35. Chen, C.; Wang, X. Adsorption of Ni(II) from Aqueous Solution Using Oxidized Multiwall Carbon Nanotubes. Ind. Eng. Chem. Res. 2006, 45, 9144–9149. [Google Scholar] [CrossRef]
  36. Ma, D.; Zou, X.; Li, R.; Chen, P.; Wang, Y.; Chen, T.; Zhang, Q.; Liu, H.; Chen, Y.; Lv, W.; et al. Highly efficient adsorption of Pb(II) by cubic nanocrystals in aqueous solution: Behavior and mechanism. Arab. J. Chem. 2020, 13, 5229–5240. [Google Scholar] [CrossRef]
  37. Patel, H. Review on solvent desorption study from exhausted adsorbent. J. Saudi Chem. Soc. 2021, 25, 101302. [Google Scholar] [CrossRef]
  38. Momčilović, M.; Purenović, M.; Bojić, A.; Zarubica, A.; Ranđelović, M. Removal of lead(II) ions from aqueous solutions by adsorption onto pine cone activated carbon. Desalination 2011, 276, 53–59. [Google Scholar] [CrossRef]
  39. Depci, T.; Kul, A.R.; Önal, Y. Competitive adsorption of lead and zinc from aqueous solution on activated carbon prepared from Van apple pulp: Study in single- and multi-solute systems. Chem. Eng. J. 2012, 200–202, 224–236. [Google Scholar] [CrossRef]
  40. Ibrahim, M.N.M.; Ngah, W.S.W.; Norliyana, M.S.; Daud, W.R.W.; Rafatullah, M.; Sulaiman, O.; Hashim, R. A novel agricultural waste adsorbent for the removal of lead (II) ions from aqueous solutions. J. Hazard. Mater. 2010, 182, 377–385. [Google Scholar] [CrossRef]
  41. Motsa, M.M.; Mamba, B.B.; Thwala, J.M.; Msagati, T.A.M. Preparation, characterization, and application of polypropylene–clinoptilolite composites for the selective adsorption of lead from aqueous media. J. Colloid Interface Sci. 2011, 359, 210–219. [Google Scholar] [CrossRef] [PubMed]
  42. Cherono, F.; Mburu, N.; Kakoi, B. Adsorption of lead, copper and zinc in a multi-metal aqueous solution by waste rubber tires for the design of single batch adsorber. Heliyon 2021, 7, e08254. [Google Scholar] [CrossRef]
  43. Zand, A.D.; Abyaneh, M.R. Adsorption of Lead, manganese, and copper onto biochar in landfill leachate: Implication of non-linear regression analysis. Sustain. Environ. Res. 2020, 30, 18. [Google Scholar] [CrossRef]
  44. Nyairo, W.N.; Eker, Y.R.; Kowenje, C.; Akin, I.; Bingol, H.; Tor, A.; Ongeri, D.M. Efficient adsorption of lead (II) and copper (II) from aqueous phase using oxidized multiwalled carbon nanotubes/polypyrrole composite. Sep. Sci. Technol. 2018, 53, 1498–1510. [Google Scholar] [CrossRef]
  45. Li, Y.-H.; Wang, S.; Wei, J.; Zhang, X.; Xu, C.; Luan, Z.; Wu, D.; Wei, B. Lead adsorption on carbon nanotubes. Chem. Phys. Lett. 2002, 357, 263–266. [Google Scholar] [CrossRef]
  46. Alghamdi, A.A.; Al-Odayni, A.-B.; Saeed, W.S.; Al-Kahtani, A.; Alharthi, F.A.; Aouak, T. Efficient Adsorption of Lead (II) from Aqueous Phase Solutions Using Polypyrrole-Based Activated Carbon. Materials 2019, 12, 2020. [Google Scholar] [CrossRef] [Green Version]
  47. Ali, I.H.; Al Mesfer, M.K.; Khan, M.I.; Danish, M.; Alghamdi, M.M. Exploring Adsorption Process of Lead (II) and Chromium (VI) Ions from Aqueous Solutions on Acid Activated Carbon Prepared from Juniperus procera Leaves. Processes 2019, 7, 217. [Google Scholar] [CrossRef]
Figure 1. A schematic diagram for preparing the PS/AlOOH hybrid adsorbent.
Figure 1. A schematic diagram for preparing the PS/AlOOH hybrid adsorbent.
Colloids 06 00072 g001
Figure 2. XRD pattern of the AlOOH, PS/AlOOH, and Pb(II)-adsorbed PS/AlOOH.
Figure 2. XRD pattern of the AlOOH, PS/AlOOH, and Pb(II)-adsorbed PS/AlOOH.
Colloids 06 00072 g002
Figure 3. SEM images of (a) AlOOH, (b) PS/AlOOH, and (c) Pb(II)-adsorbed PS/AlOOH.
Figure 3. SEM images of (a) AlOOH, (b) PS/AlOOH, and (c) Pb(II)-adsorbed PS/AlOOH.
Colloids 06 00072 g003
Figure 4. TEM images of the PS/AlOOH hybrid adsorbent.
Figure 4. TEM images of the PS/AlOOH hybrid adsorbent.
Colloids 06 00072 g004
Figure 5. (a) EDX analysis and elemental mapping of the Pb-adsorbed PS/AlOOH hybrid adsorbent; (b) C, (c) Al, (d) O, (e) Pb, and (f) mixed elements.
Figure 5. (a) EDX analysis and elemental mapping of the Pb-adsorbed PS/AlOOH hybrid adsorbent; (b) C, (c) Al, (d) O, (e) Pb, and (f) mixed elements.
Colloids 06 00072 g005
Figure 6. FTIR spectrum of the PS/AlOOH (a) before adsorption and (b) after Pb(II) adsorption.
Figure 6. FTIR spectrum of the PS/AlOOH (a) before adsorption and (b) after Pb(II) adsorption.
Colloids 06 00072 g006
Figure 7. (a) Adsorption kinetic plots for Pb(II) removal; (b) intra-particle diffusion plot (pH—5.6, solution volume—20 mL, Pb(II) conc.—100 mg/L, temp.—32 °C, PS/AlOOH mass—0.05 g).
Figure 7. (a) Adsorption kinetic plots for Pb(II) removal; (b) intra-particle diffusion plot (pH—5.6, solution volume—20 mL, Pb(II) conc.—100 mg/L, temp.—32 °C, PS/AlOOH mass—0.05 g).
Colloids 06 00072 g007
Figure 8. Adsorption of Pb(II) at varying solution pH values (time—150 min, solution volume—20 mL, Pb(II) conc.—100 mg/L, temp.—32 °C, PS/AlOOH mass—0.05 g).
Figure 8. Adsorption of Pb(II) at varying solution pH values (time—150 min, solution volume—20 mL, Pb(II) conc.—100 mg/L, temp.—32 °C, PS/AlOOH mass—0.05 g).
Colloids 06 00072 g008
Figure 9. Adsorption isotherm models plot for Pb(II) removal (pH—5.6, solution volume—20 mL, temp.—32 °C, PS/AlOOH mass—0.05 g).
Figure 9. Adsorption isotherm models plot for Pb(II) removal (pH—5.6, solution volume—20 mL, temp.—32 °C, PS/AlOOH mass—0.05 g).
Colloids 06 00072 g009
Figure 10. (a) Adsorption of Pb(II) onto PS/AlOOH at varying solution temperatures, and (b) reusability study of PS/AlOOH.
Figure 10. (a) Adsorption of Pb(II) onto PS/AlOOH at varying solution temperatures, and (b) reusability study of PS/AlOOH.
Colloids 06 00072 g010
Table 1. The values of kinetic parameters for the adsorption of Pb(II).
Table 1. The values of kinetic parameters for the adsorption of Pb(II).
Pseudo-First-OrderPseudo-Second-OrderElovich Model
qe (mg g−1)22.934qe (mg g−1)25.813a (mg g−1 min−1/2)5.192
k1 (min−1)0.0541k2 (g mg−1 min−1)0.0029β (mg g−1)0.204
R20.9544R20.9815R20.9711
RMSE1.414RMSE1.040RMSE1.125
χ22.605χ20.820χ20.567
Table 2. The values of isotherm parameters for Pb(II) adsorption onto the PS/AlOOH adsorbent.
Table 2. The values of isotherm parameters for Pb(II) adsorption onto the PS/AlOOH adsorbent.
LangmuirFreundlichRedlich–PetersenToth
qm23.611 mg g−1n4.962KRP30.555 L g−1qm21.355 mg g−1
KL1.168 L mg−1 Kf2.003 mg g−1 (mg L−1)−1/n aRP1.412 L mg−1KTo0.699 L mg−1
β0.972tn0.984
R20.9760R20.9438R20.9863R20.9864
RMSE1.036RMSE2.047RMSE1.009RMSE1.004
χ20.390χ23.457χ20.411χ20.409
Table 3. Thermodynamic parameters for Pb(II) adsorption on PS/AlOOH.
Table 3. Thermodynamic parameters for Pb(II) adsorption on PS/AlOOH.
Temperature −ΔG° (kJ mol−1)ΔH° (kJ mol−1)ΔS° (Jmol−1 K−1)
3050.818
3130.9473.75314.998
3231.088
Table 4. Pb(II) adsorption capacity of various adsorbents.
Table 4. Pb(II) adsorption capacity of various adsorbents.
Adsorbent Adsorption Capacity (mg/g)Ref
Alumina13.11[1]
Pinecone-activated carbon27.53[38]
Activated carbon17.77[39]
Modified soda lignin46.72[40]
Polypropylene–clinoptilolite1.01[41]
Activated carbon from waste rubber tire9.6805[42]
Biochar0.44[43]
Polypyrrole/oMWCNT 26.32[44]
Acidified CNTs 17.44[45]
Polypyrrole-based activated carbon50[46]
Acid-activated carbon30.3[47]
PS/AlOOH23.61This study
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kumar, R. Fabrication of Polystyrene/AlOOH Hybrid Material for Pb(II) Decontamination from Wastewater: Isotherm, Kinetic, and Thermodynamic Studies. Colloids Interfaces 2022, 6, 72. https://doi.org/10.3390/colloids6040072

AMA Style

Kumar R. Fabrication of Polystyrene/AlOOH Hybrid Material for Pb(II) Decontamination from Wastewater: Isotherm, Kinetic, and Thermodynamic Studies. Colloids and Interfaces. 2022; 6(4):72. https://doi.org/10.3390/colloids6040072

Chicago/Turabian Style

Kumar, Rajeev. 2022. "Fabrication of Polystyrene/AlOOH Hybrid Material for Pb(II) Decontamination from Wastewater: Isotherm, Kinetic, and Thermodynamic Studies" Colloids and Interfaces 6, no. 4: 72. https://doi.org/10.3390/colloids6040072

Article Metrics

Back to TopTop