Next Article in Journal
From Nano-Crystals to Periodically Aggregated Assembly in Arylate Polyesters—Continuous Helicoid or Discrete Cross-Hatch Grating?
Next Article in Special Issue
Removal of Lead from Wastewater Using Synthesized Polyethyleneimine-Grafted Graphene Oxide
Previous Article in Journal
Nitrogen Adsorption and Characteristics of Iron, Cobalt, and Nickel Oxides Impregnated on SBA-15 Mesoporous Silica
Previous Article in Special Issue
Effective Thallium(I) Removal by Nanocellulose Bioadsorbent Prepared by Nitro-Oxidation of Sorghum Stalks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile Synthesis of the Polyaniline@Waste Cellulosic Nanocomposite for the Efficient Decontamination of Copper(II) and Phenol from Wastewater

Department of Environmental Sciences, Faculty of Meteorology, Environment and Arid Land Agriculture, King Abdulaziz University, Jeddah 21589, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(6), 1014; https://doi.org/10.3390/nano13061014
Submission received: 13 February 2023 / Revised: 8 March 2023 / Accepted: 9 March 2023 / Published: 11 March 2023
(This article belongs to the Special Issue The Application of Nanomaterials in Heavy Metal Detection and Removal)

Abstract

:
The existence of heavy metals and organic pollutants in wastewater is a threat to the ecosystem and a challenge for researchers to remove using common technology. Herein, a facile one-step in situ oxidative polymerization synthesis method has been used to fabricate polyaniline@waste cellulosic nanocomposite adsornt, polyaniline-embedded waste tissue paper (PANI@WTP) to remove copper(II) and phenol from the aqueous solution. The structural and surface properties of the synthesized materials were examined by XRD, FTIR, TEM, and a zeta potential analyzer. The scavenging of the Cu(II) and phenol onto the prepared materials was investigated as a function of interaction time, pollutant concentration, and solution pH. Advanced kinetics and isotherms modeling is used to explore the Cu(II) ion and phenol adsorption mechanisms. The synthesized PANI@WTP adsorbent showed a high intake capacity for Cu(II) than phenol, with the maximum calculated adsorption capacity of 605.20 and 501.23 mg g−1, respectively. The Langmuir equilibrium isotherm model is well-fitted for Cu(II) and phenol adsorption onto the PANI@WTP. The superior scavenging capability of the PANI@WTP for Cu(II) and phenol could be explained based on the host–guest interaction forces and large active sites. Moreover, the efficiency of the PANI@WTP for Cu(II) and phenol scavenging was excellent even after the five cycles of regeneration.

1. Introduction

Wastewater containing pollutants such as heavy metals and organic compounds reduces water quality and threatens the ecosystem and human health. Removing and recovering organic contaminants and metal ions from aqueous effluents is difficult. This is due to their varied physical and chemical characteristics, which hinder the complete removal and recovery. Industrial and mining activity and other sources introduce heavy metals and organic pollutants into water systems [1]. Water containing harmful metals such as copper, iron, lead, mercury etc., and organic debris such as phenol has health-related consequences for humans and animals. Copper is a micronutrient trace element required for human and animal nutrition. Copper is toxic and carcinogenic and can accumulate in the liver, leading to stomach cramps, breathing problems, and liver and kidney loss [2]. On the other hand, phenol and its derivatives are highly toxic compounds that significantly impact human health and aquatic species. Approximately 6 million tons of phenol are generated each year globally, with a marked upward tendency [3,4].
Among the technologies reported to remove hazardous heavy metal cations and phenolic compounds from polluted water are chemical precipitation [5], membrane processes [6], flotation [7], solvent extraction [8], oxidation [9], adsorption [10], ion exchange [11], and electrochemical methods [12]. Each technology has some advantages and disadvantages. However, most of these treatment methods have shortcomings such as expense, secondary pollution, or being poor/ineffective etc. For instance, chemical precipitation generates more sludge since it uses more chemicals than necessary to treat the water, while other technologies need high capital investment and energy. However, the adsorption technique is considered relatively cost effective, practical, and simple to use [5,13,14]. The applications of convectional adsorbent materials, such as activated carbon (AC) [15], zeolite [16,17], and resins [18], and so forth, are widely investigated to decontaminate organic as well as inorganic pollutants from water streams. Scientists have recently been searching for the fabrication of low-cost adsorbents, mainly made from biopolymers such as alginate, cellulose, waste biomass, etc., as they are low-cost, abundantly available, and efficient adsorbents [19].
Cellulose (C6H11O5) is the most abundant organic biopolymer on Earth [20,21]. Cellulose can be used to produce a variety of nanomaterials, cellulose nanofibrils, and oxidized cellulose nanoparticles (CNFs) etc. Organic moieties such as carboxylic and hydroxyl groups can be changed or grafted with other functional groups to create new functional groups that can be adapted explicitly for application. These functional groups ensure that the addition of nanoparticles can modify cellulose biopolymers. The cellulose materials’ functional groups are crucial to the adsorption of metal ions [22]. Chemically modified cellulose has a greater adsorption capacity for various aquatic contaminants than its unmodified counterparts. Organic compounds, acids, bases, minerals, oxidizing agents, and other substances have been employed to modify cellulose [23,24]. Tissue paper (TP) is a cellulose fiber-rich material widely used for multiple purposes. It produces a massive amount of waste after its use. Waste tissue paper (WTP) has been used to prepare carbon aerogel and activated carbon (AC), and is utilized as an ideal candidate for the adsorption of organic chemicals and heavy metals from effluent. Products made of spongy tissue absorb water quickly while maintaining a high absorption capacity throughout the task. The ability to absorb liquid relies on enough surface tension to suck out the liquid and a high permeability. Due to the high hydrophilicity, TP absorbs a large amount of water and fewer pollutants. Therefore, to enhance pollutant adsorption and reduce its hydrophilicity, TP should be altered with low surface energy materials [25].
Among conductive polymers, polyaniline has gained huge research attention because of its important characteristics, including a simple synthesis route, environmental stability, good electrical conductivity, etc. Various applications related to active functional groups, such as amine, imine, and secondary amino groups on the polyaniline structure, make it suitable for removing dyes, toxic metals, and organic chemicals and dyes from the aqueous phase [26]. However, PANI’s significance is constrained by several drawbacks, including insolubility or low/partial solubility in common solvents, long-chain polymer aggregates, poor processability, and infusibility. Moreover, after a prolonged cycle time, its electrical conductivity also declines. Polyaniline has been combined with an additional adsorbent with more excellent adsorption capabilities, good regeneration ability, and selectivity utilizing straightforward synthesis procedures to develop novel nanocomposite adsorbents. The most attractive PANI/cellulose-based nanocomposites exhibit electrical conductivity because of the combined use of cellulose nanofillers and PANI matrix. Such nanomaterials, which have enhanced conductive, electrical, mechanical, and adsorbing capabilities, are widely used in the water treatment and electronics industries, as well as in the biomedical and electronics industries [24,26]. Both chemical and electrochemical oxidative polymerization in acidic media can be used to synthesize PANI. Ammonium persulfate (APS) and potassium persulfate (KPS) are the most often utilized initiators or oxidants for the chemical polymerization of aniline because they provide better conversion and yield. Several different macromolecular PANI structures exist due to the variance in experimental synthesis circumstances. In situ polymerization is the most popular technique for developing PANI nanocomposites [26]. Therefore, fabricating a WTP composite with polyaniline could be an excellent strategy to enhance pollutant adsorption ability and lower its wettability.
In this article, waste tissue paper and polyaniline (PANI@WTP) nanocomposite were synthesized via a facile single-step in situ oxidative polymerization of aniline in the presence of ammonium persulfate (NH4)2S2O8 as an oxidizing agent. The fabricated nanoporous polyaniline waste cellulosic tissue paper-based adsorbent was utilized for scavenging Cu(II) and phenol from synthetic wastewater solution. Variable adsorption parameters such as pollutant concentration, solution pH, and reaction time were investigated. Advanced adsorption kinetics and isotherm models were also fitted to the experimental data.

2. Materials and Methods

2.1. Chemicals

Ammonium persulfate extra pure (NH4)2S2O8) was purchased by SD Fine Chemicals India. Cupric Sulfate (CuSO4.5H2O), phenol (C6H5OH), and aniline were obtained from BHD Chemical Ltd., Poole, UK, and for the analysis of the copper solution, HACH Permachem reagents (HACH LANGE Gmbh, Duesseldorf, Germany) CuVer 1 copper reagent (category no. 2105869) was used. The DR/6000 UV-visible spectrophotometer was utilized to analyze the phenol as well as Cu(II) solutions.

2.2. Pulverization of WTP

The waste tissue paper used in the present study was collected from the industrial waste treatment laboratory of King Abd ulaziz University in Jeddah, Saudi Arabia, and sterilized in UV light for 2 h. The sterilized WTP was then washed with hot and cold water and heated in the oven drier for 24 h at 105 °C. Then, the dried WTP was soaked in deionized water for 30 min and exfoliated in a blender. This process was repeated several times until exfoliated WTP fibers were obtained. The exfoliated fibrous WTP was dried at 105 °C.

2.3. Synthesis of PANI@WTP Nanocomposite

The PANI@WTP composite was synthesized with aniline polymerization on the exfoliated WTP. Initially, 1 g of exfoliated WTP was suspended in 30 mL ethanol and sonicated for 10 min. After that, 150 mL 0.5 M HCl containing 1.35 mL aniline was added to the WTP suspended and stirred for half an hour in an ice bath. Ammonium persulfate (0.5 M) solution prepared in 1 M HCl (50 mL) was added dropwise for polymerizing the aniline onto WTP. The solution was agitated for 18 h. Blue-green PANI@WTP composite was then filtered and cleaned with water, ethanol, and acetone until the color was removed, then dried for 24 h at 60 °C in the oven. A similar method was used to synthesize polyaniline in the absence of WTP.

2.4. Characterization

To examine the surface structural and chemical compositions of waste tissue paper and the synthesized PANI@WTP composite, X-ray diffraction (XRD), Bruker D8 Advance X-ray diffractometer (Bruker Inc., Bremen, Germany) was used. A Fourier-transform infrared spectrometer (FTIR) model Agilent Cary 630 spectrometer was used to analyze the functional groups on WTP, PANI@WTP composite, PANI@WTP -Cu(II), and PANI@WTP-phenol before and after adsorption. The zeta potential was analyzed using Malvern panalytical serial no. MA1070389 in pH range of 2–10.

2.5. Adsorption of Cu(II) and Phenol

Batch-adsorption studies were completed using 0.02 g PANI@WTP composite dosage in 20 mL of Cu(II) and phenol solutions at various concentrations (10–800 mg/L), time (0–480 min), and solution pH (2–9). The adsorption tests were conducted in triplet, and the average results were reported. Reagents of 0.1 M HCl and 0.1 M NaOH were utilized to change the pH of the solution. After the adsorption of phenol onto the PANI@WTP composite, the residual concentrations of phenol in solutions were analyzed at wavelength λ-271 nm. CuVer 1 copper reagent (cat 2105869) was used to analyze the Cu(II) on the HACH DR-6000 spectrophotometer. The following equations determined the adsorption at equilibrium:
q e = ( C 0   C e ) v m  
where, qe is the uptake capacity of Cu(II) and phenol in (mg/g) onto PANI@WTP composite at equilibrium, Ce is Cu(II) and phenol concentrations in solution (mg/L) at equilibrium, and C0, the initial Cu(II) ion and phenol solutions (mg/L). The volume (v) of the solution is in liters, and the dry weight (m) of PANI@WTP composite is in grams.

2.6. Desorption and Regeneration Studies

The desorption of both pollutants and regeneration of the spent PANI@WTP composite adsorbent was carried out using 0.1 M NaOH and 0.1 HCl. A fixed amount (0.02 g) of spent PANI@WTP composite was mixed with 20 mL of 0.1 M NaOH or 0.1 HCl for pollutant desorption in the solution for 3 h under shaking conditions. Then, PANI@WTP was filtered and thoroughly washed with deionized water and dried at 70 °C for 16 h. The dried PANI@WTP was again used as an adsorbent to remove Cu(II) and phenol at the optimum adsorption conditions. A similar adsorption-desorption process was repeated for up to five cycles.

2.7. Synthetic Tap and Groundwater Purification

The adsorption performance of the PANI@WPT was also tested for the removal of the Cu(II) and phenol from the synthetic tap and groundwater. Therefore, a solution of Cu(II) and phenol was prepared by adding the appropriate amount to the tap and groundwater. A fixed amount of CuSO4.5H2O and phenol were added to prepare the 20 mg/L and 100 mg/L concentration solution in tap and groundwater. The adsorption studies were performed by mixing 0.02 g PANI@WPT in 20 mL prepared solution at 30 °C for 3 h at pH 5.2 for Cu((I) and pH 5 for phenol.

3. Results and Discussions

3.1. Synthesis and Characterization

Tissue paper (TP) is a flexible, soft, lightweight absorbent made of cellulose fibers. The paper pulp containing cellulose fiber is used to make the TP sheets. The WTP sheet was pulverized into a fluffy fibrous texture utilizing the blender. The idea to immobilize polyaniline onto the WTP was to reduce the hydrophilicity of the cellulose fibers and enhance the adsorption affinity toward pollutants. The existence of the hydroxyl functional groups on the cellulose makes the TP absorb more water. The interaction of the cellulose’s hydroxyl groups with the amine groups of the polyaniline can reduce the wettability of the TP and add more active sites for the interaction with the pollutants. The proposed scheme for the PANI@WTP synthesis is shown in Figure 1. To confirm successful synthesis, PANI@WTP was characterized by XRD, TEM, and FTIR spectroscopy.
The XRD pattern of the WTP and PANI@WTP is shown in Figure 2. The XRD spectrum of WTP displayed three major peaks at 2θ = 16.43°, 22.66°, and 34.47° corresponding to reflection lattice planes (110), (200), and (004) of the cellulose fiber structures. The highly crystalline structure of the WTP indicates the cellulose nanocrystal structure [25]. The XRD pattern of the PANI is shown in Figure S1, which shows two characteristic peaks centered at 19.35° and 25.65°. In comparison to the WTP and PANI, the XRD pattern of PANI@WTP nanocomposite showed a slight change in the peak positions (low) and peak intensity (increases) at 2θ = 15.29°, 22.33°, 31.57°, and 34.49°. The slight reduction in peak positions can be explained based on the polyaniline interaction with the WTP. The increase in peak height can be due to the solubilization of hemicellulose in acidic conditions during the PANI@WTP synthesis. The TEM images of WTP and PANI@WTP composite material are shown in Figure 3. The TEM image of WTP Figure 3a,b shows a fiber-like elongated network mainly due to the cellulosic structure. The TEM image of the PANI@WTP composite in Figure 3b shows the porous PANI particles deposited on the surface of WTP. The PANI particles are porous and of different shapes and sizes. However, the PANI particle alone shows various shapes and size particles, as shown in Figure S2. These results indicate the WTP has an effect on the porosity of the PANI in the PANI@WTP composite.
The FTIR spectrums of WTP and PANI@WTP before and after phenol and Cu(II) adsorption are displayed in Figure 4. The FTIR spectra of the PANI showing the characteristic peaks at 3439, 1575, 1475 1297, 1113, and 796 cm¹ belong to the amine group, nitrogen quinine, benzene ring, C–N, and C=N stretching vibrations. The peak that appeared at nearly 3340 cm−1 from WTP was formed as a result of –OH stretching vibrations of the hydroxyl group on the cellulose chemical structure [27]. This peak shifted to 3323 cm−1 after aniline polymerization with WTP, indicating the interaction between the –OH group of cellulose and –NH group of PANI. The peak observed at 2900 cm−1 was due to the vibrational stretching of C−H group on the surface of the biopolymer [28]. The narrow peaks at 1428 cm−1 in the lower frequency region of WTP were due to the –CH2 group on the adsorbent. The N–H and –OH stretching vibration of the amino group of polyaniline and hydroxyl group of WTP were assigned to a characteristic peak at 3323 cm−1 in PANI@WTP composite [29]. However, the absorption peak observed at 3323 cm−1 was shifted to 3339 cm−1 in PANI@WTP–Cu(II) due to the complexation of the hydroxyl and amide group with Cu(II) ion after adsorption [29]. The aromatic compound (C–N) on the composite adsorbent formed a tiny peak at 1312 cm−1 [30]. The peak at 1024 cm−1 is attributed to the C–O–C ring in WTP [31]. This peak was shifted to 1026 cm−1 after adsorption with phenol. The shift of the peak band from 2900 cm−1 to 2809 cm−1 in PANI@WTP was attributed to the stretching vibration of the hydrocarbons (C–H) and hydrogen bonding between WTP and PANI during polymerization [27]. The same peak value (2809 cm−1) was also reported for polyaniline-impregnated nano cellulose adsorbent for the adsorption of hexavalent chromium [32]. The peak region between 1000 and 1150 cm−1 designated to the C–O–C bond vibration and this peak was like the peak band at 1158 cm−1 as illustrated in Figure 4. It is also noted that after adsorption with the metal ion Cu(II) and phenol, the adsorption peak corresponding to the phenolic –OH and NH+ functional groups changed. The change in wavenumber is an the implication of the participation of different functional group on Cu(II) and phenol adsorption from the aqueous media and the formation of bonds with PANI@WTP [33].

3.2. Scavenging of Cu(II) and Phenol

In the comparative adsorption tests for removing the Cu(II) and phenol from the aqueous solution onto WTP, PANI, and PANI@WTP nanocomposite, 0.02 g of material was mixed with a 20 mL (100 mg/L) solution at a pH of 5. The results (Figure S3) indicated that the adsorption efficiency of the Cu(II) and phenol onto WPT, PANI, and PANI@WTP nanocomposite was 46, 38, and 96.23 mg/g and 30, 38, and 74.3 mg/g, respectively. The PANI@WTP nanocomposite showed a higher adsorption of Cu(II) and phenol than WPT and PANI due to a large number of functional groups present on the surface of the PANI@WTP nanocomposite. These findings were used to evaluate the effectiveness of PANI@WTP nanocomposite for Cu(II) and phenol as a function of the adsorption parameters such as solution pH, pollutant concentration, and contact time.

3.3. Effect of Solution pH on Adsorption Process

The pH of a solution plays a significant role in determining the surface charge of the adsorbent during the adsorption mechanism and the extent of ionization of pollutants [34]. The zeta potential analysis of PANI@WTP nanocomposite at various pHs is shown in Figure 5a. As can be shown from Figure 5a, the point of zero charge (PZC) of the PANI@WTP nanocomposite is 9.76, which indicates that the adsorbent becomes positively charged at a pH < 9.76 and negatively charged at a pH > 9.76. The impact of the pH on the adsorption of Cu(II) and phenol was evaluated in the range of 2–5.5 and 2–10, respectively, The solution pH was adjusted between 2 and 5.5 for Cu(II) to avoid precipitation [35]. The highest scavenging of Cu(II) and phenol was achieved at pH 5.2 and 3, respectively. Cu(II) adsorption capability at the low pH range can be related to competition for similar adsorption sites between hydrogen ions and divalent Cu(II) ions [35,36]. The pH of the solution also has a significant influence on the phenol elimination as shown in Figure 5b. In the acidic media (pH 3–5), the removal of phenol was high. The high adsorption capacity at a lower pH for phenol might be due to the protonation of the organic functional groups such as carbonyl found both on the adsorbate and surface of the adsorbent and, as a result decreasing the electrostatic repulsive force [37]. A similar phenol adsorption has been reported by Fe–nano zeolite [38].

3.4. Effect of Contact Time on Adsorption and Kinetic Models

The Cu(II) and phenol adsorption onto the PANI@WTP composite was examined at various time intervals (0–420 min). The adsorption analysis was carried out using 20 mL of Cu(II) and phenol solutions with initial pollutant concentrations of 500 mg/L, pH 5.2 for Cu(II), and pH 5 for phenols at 30 °C, and the outcomes are shown in Figure 6. A quick removal of Cu(II) and phenol was detected during the first 5 min of the adsorption process, and subsequently, scavenging gradually increased [39,40]. It was discovered that the rate of phenol adsorption was slightly lower than Cu(II) solution. The higher adsorption of the Cu(II) onto PANI@WTP can be explained based on its smaller size than the phenol. The Cu(II) ionic radius is 0.73 A (0.074 nm), while phenol’s effective molecular diameter is 0.75 nm. The small-size Cu(II) can easily diffuse into the porous structure of the PANI@WTP [41,42]. Moreover, the fast removal of the Cu(II) ion at the beginning of the adsorption process might be due to the availability of more adsorption sites of PANI@WTP composite, and as time passed, the active sites were occupied [43].
The rate of the adsorption process for the uptake of Cu(II) and phenol was determined using adsorption kinetics. There are two main processes: physical adsorption and chemical adsorption. The physical adsorption is caused by weak attraction forces (van der Waals). In contrast, chemisorption necessitates the formation of a strong connection between the solvent and the substrate to permit the activation of atoms [44]. The rate of adsorption kinetics for Cu(II) and phenol in liquid–solid interactions has been modeled using pseudo-first-order, pseudo-second-order, and Elovich kinetic models [45]. The nonlinear pseudo-first, pseudo-second-order, and Elovich kinetic models are described in Equations (2)–(4), respectively.
q t = q e ( 1 e k 1 t ) ,
q t = k 2 q e 2 t 1 + k 2 q e t  
q t = 1 β + l n ( α β ) + 1 β l n ( t )
where qe and qt are the amounts of adsorbate uptake per mass of adsorbent at equilibrium and at any time t (min), respectively, and k1 (min¹) is the rate constant of the pseudo-first-order equation, and k2 (g mg−1 min−1), is the pseudo-second-order equation constant rate. Elovich constant α (mg g−1 min−1) is the adsorption rate, and β (mg g−1) is the desorption coefficient.
The nonlinear plots for applied kinetic models for Cu(II) adsorption and phenol onto PANI@WTP are shown in Figure 6a,b, respectively. The computed parameters of the kinetic models are illustrated in Table 1. The correlation coefficient (R2) values for pseudo-first-order kinetic, pseudo-second-order, and Elovich kinetic models are examined to assess the suitability of the experimental data, as tabulated in Table 1. The calculated kinetic parameter correlation coefficient (R2) value suggested the Elovich kinetic model had a high (R2) value (0.9485) for Cu(II) compared to the pseudo-first and pseudo-second-order kinetic model. The order of kinetic model best fitted the removal of Cu(II) by PANI@WTP was Elovich kinetic model > pseudo-second order > pseudo-first-order kinetic model, respectively. The scavenging of phenol onto PANI@WTP was best fitted to the pseudo-second-order kinetic model with a high R2 value (0.9746). The order of the best-suited kinetic model for phenol scavenging onto PANI@WTP was in the following order: pseudo-second-order kinetic > pseudo-first-order kinetic > Elovich kinetic model. Furthermore, the values of computed adsorption capacity qecal.: 396.4154 mg/g for the scavenging of Cu(II) from the pseudo-second-order kinetic model was consistent to the data obtained from the experimental value qeexp: 397 mg/g, except for a negligible difference. The calculated adsorption capacity was qecal.: 328.3705 mg/g for the phenol from the pseudo-second-order kinetic model was also close to the experimental data (320 mg/g).

3.5. Effect of Concentration on Adsorption and Isotherm Models

The adsorption capacity is significantly influenced by the initial Cu(II) and phenol concentrations [26]. The impact of the initial feed concentration of Cu(II) and phenol on adsorption is shown in Figure 7a,b, respectively. As seen from Figure 7, adsorption capacity increases from 8.86 mg/g to 397 mg/g for Cu (II) and 2.45 to 320 mg/g for phenol. The capacity of the PANI@WTP increases as the initial amount of Cu(II)/phenol rises. As the initial pollutant concentration rises, the number of Cu(II) and phenol-filled adsorptive sites of PANI@WTP increases, resulting in increased adsorption capacity. Cu(II) ions and phenol competed for a fixed number of dynamic locations on the adsorbent at greater concentrations. As a result, the contaminant molecules have insufficient binding sites on PANI@WTP at higher concentrations [46].
The well-known adsorption isotherms, namely the Langmuir, Freundlich, and Temkin adsorption isotherms, were applied to assess the interaction among PANI@WTP and adsorbate. According to the Langmuir adsorption concept, molecules are adsorbed at a definite number of well-defined catalyst surfaces, evenly distributed across the adsorbent’s surface. There is no connection between the adsorbate species because these binding sites have a similar potential for the adsorption of a monomolecular layer [31]. The Freundlich isotherm is useful when working with heterogeneous sorbent media to determine the sorption phenomenon. The Freundlich isotherm is assumed from the principle that adsorptive sites disperse exponentially in relation to the heat of the sorption model. The Temkin isotherm model describes the interactions of the adsorbent and adsorbate during adsorption progression. The assumptions of the Temkin model are based on the concept that during the adsorption of the sorbate and sorbent, neglecting the effect of extremely low and large values of concentrations, the heat of the adsorption will not be changed. Rather, the heat declines with coverage because of the interactions of the sorbate and sorbent during adsorption [47]. The following equations, respectively, represent the Langmuir, Freundlich, and Temkin isotherms.
q e = q m K L C e 1 + K L C e
q e = K F C e 1 / n  
q e = B t   ln ( K t C e )
where KF (the strength factor) and n are the Freundlich adsorption constants that were determined from the intercept and slope of the linear plots of log qe vs. log Ce, respectively and Langmuir constants, KL (L/mg) are constants that are related to adsorption capacity and energy or net enthalpy of adsorption, respectively. qe is the corresponding adsorption capacity (mg g−1), qm represents maximum adsorption capacity (mg g−1) and Ce is the Cu(II) and phenol equilibrium concentration (mg/L) respectively. Kt and Bt are the Temkin constants that refer to the heat of adsorption. The Langmuir, Freundlich, and Temkin isotherm models for the adsorption of Cu(II) and phenol onto PANI@WTP are illustrated in Figure 7a,b, respectively. The computed parameters of the Langmuir, Freundlich, and Temkin isotherm models computed by nonlinear regression of qe vs. Ce are tabulated in Table 2. According to the higher R2 values in Table 2, the Langmuir isotherm model (R2, 0.9439) followed by the Freundlich isotherm model (R2, 0.8699) best described the experimental results for the adsorption of Cu(II) onto PANI@WTP. However, the R2 value (0.9647) estimated for the adsorption of phenol on PANI@WTP nanocomposite, calculated from the Langmuir isotherm model equation, followed by the Temkin isotherm model, was well suited to the experimental results.

3.6. Cu(II) and Phenol Adsorption Mechanism

The scavenging Cu(II) and phenol onto the PANI@WTP nanocomposite was influenced by several parameters, which include physical and chemical features as well as the surface features of the adsorbent and the nature of the pollutant itself. The higher adsorption capacity of the PANI@WTP nanocomposite than the WTP was due to the large number of active sites. The WPT only has the cellulose-based active sites (–OH and =O), while after coupling with PANI, the PANI@WTP nanocomposite has –OH, =O, NH, =N, and NH2 adsorption sites for Cu(II) and phenol. The effective adsorption of Cu(II) ions onto PANI@WTP could be described as the interaction between Cu(II) ions and the hydroxyl and amine-functional groups [30]. The increase of adsorption capacity with increasing pH for Cu(II) ion suggested that the adsorption mechanism might be due to the ion exchange properties of positively charged Cu(II) and H+ on +NH2 of the composite [31]. Moreover, as the solution pH increases, the surface charge of the adsorbent becomes more negative, which favours the adsorption of the Cu(II) at the higher pH. Furthermore, the adsorption mechanism of Cu(II) ions might be described based on the FTIR analysis. The change in wave number indicated the interaction of Cu(II) ion via complexation with –OH and amide functional groups on the surface of the material [35]. Moreover, the adsorption of phenol onto PANI@WTP may be due to the strong interactions of the NH+ from the polyaniline and delocalized oxygen atoms from the phenol [48]. Furthermore, the pseudo-second-order kinetic model fitting for phenol adsorption onto PANI@WTP indicated that the phenol adsorption process might be due to the interactions of π-π electrons from the aromatic ring on phenol and PANI [49]. In general, the adsorption is a complex process with more mechanisms; hence, hydrogen bonding, physical adsorption, and π-π interactions might take part in the adsorption process of Cu(II) and phenol removal onto the PANI@WTP composite [50,51].

3.7. Regeneration

The regeneration of the used adsorbent material is essential to make the adsorption process cost effective [52]. The reusability of the PANI@WTP nanocomposite was studied for five cycles, as shown in Figure 8. The adsorption–desorption experimental results depicted that the PANI@WTP nanocomposite adsorbent can be regenerated several times with a slight loss in adsorption capability. This might be due to the deactivation of the active sites and incomplete desorption of the pollutant species from the inner structure of the adsorbent.

3.8. Comparison of the Adsorption Efficiency

The PANI@WTP nanocomposite efficiency was compared with other adsorbents utilized for Cu(II) and phenol adsorption. Table 3 shows the comparative adsorption capacity of the other adsorbents for Cu(II) and phenol from literature. The findings of the comparison demonstrate that the PANI@WTP nanocomposite is an effective material for scavenging Cu(II) and phenol from wastewater.

3.9. Synthetic Tap and Groundwater Purification

The adsorption performance of PANI@WTP was tested to remove Cu(II) and phenol from the tap and groundwater, and results are included in Figure 9. As seen in Figure 9a, PANI@WTP’s ability to adsorb the Cu(II) was approximately 14 mg/g for groundwater and 11.3 mg/g for tap water, while the phenol adsorption capacity was 5.19 mg/g and 5.14 mg/g at 20 mg/L. As shown in Figure 9b, a removal capacity of 43.23 mg/g and 21 mg/g was recorded for Cu(II) and 13.23 mg/g and 27 mg/g for the phenol adsorption onto the PANI@WTP composite at a 100 mg/L concentration for ground and tap water, respectively. The adsorption capacity of the adsorbent was also compared with deionized wastewater samples. It was clearly seen that the solution preparations in the tap, groundwater, or deionized water affect the adsorption process. The low adsorption of the pollutants from tap and groundwater compared to deionized water might be related to ionic species such as minerals, anions, halides etc., in the ground and tap water, which hindered the treatment process [66].

4. Conclusions

In this article, waste tissue paper and a polyaniline (PANI@WTP) nanocomposite were synthesized using a single-step in situ oxidative polymerization of aniline in the presence of ammonium persulfate (NH4)2S2O8 as an oxidizing agent. The fabricated PANI@WTP composite was examined for the decontamination of Cu(II) and phenol from wastewater. It has been found that the optimum pH values for both Cu(II) and phenol adsorption onto PANI@WTP composite were 5.2 and 3, respectively. The Elovich kinetic model and pseudo-second-order kinetic model fitted the experimental kinetic data for Cu(II) and phenol adsorption, respectively, whereas the Langmuir equilibrium isotherm model was well fitted for adsorption of both Cu(II) and phenol onto the PANI@WTP composite. In general, the study results suggested that the fabricated composite material is recyclable and still a promising candidate adsorbent for treating pollutants from wastewater.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13061014/s1, Figure S1: XRD pattern of the polyaniline, Figure S2: SEM image of the polyaniline, Figure S3: The comparison- of adsorption capacities of WPT, PANI and PANI@WTP for Cu(II) and phenol removal. (concentration; 100 mg/L, adsorbent dosage; 0.02 g, volume; 20 mL, pH 5.2 for Cu(II), pH, 5 for phenol, Temp.; 30 °C).

Author Contributions

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

Funding

This research work was funded by Institutional Fund Projects under grant no. (IFPRC-182-155-2020).

Data Availability Statement

Data is available on request.

Acknowledgments

This research work was funded by Institutional Fund Projects under grant no. (IFPRC-182-155-2020). Therefore, the authors gratefully acknowledge technical and financial support from the Ministry of Education and King Abdulaziz University, Jeddah, Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chai, W.S.; Cheun, J.Y.; Kumar, P.S.; Mubashir, M.; Majeed, Z.; Banat, F.; Ho, S.-H.; Show, P.L. A review on conventional and novel materials towards heavy metal adsorption in wastewater treatment application. J. Clean. Prod. 2021, 296, 126589. [Google Scholar] [CrossRef]
  2. Abbas, A.; Al-Amer, A.M.; Laoui, T.; Al-Marri, M.J.; Nasser, M.S.; Khraisheh, M.; Atieh, M.A. Heavy metal removal from aqueous solution by advanced carbon nanotubes: Critical review of adsorption applications. Sep. Purif. Technol. 2016, 157, 141–161. [Google Scholar]
  3. Fu, F.; Wang, Q. Removal of heavy metal ions from wastewaters: A review. J. Environ. Manag. 2011, 92, 407–418. [Google Scholar] [CrossRef]
  4. Mahdavi, M.; Amin, M.M.; Mahvi, A.H.; Pourzamani, H.; Ebrahimi, A. Metals, heavy metals and microorganism removal from spent filter backwash water by hybrid coagulation-UF processes. J. Water Reuse Desalin. 2018, 8, 225–233. [Google Scholar] [CrossRef] [Green Version]
  5. Barakat, M. New trends in removing heavy metals from industrial wastewater. Arab. J. Chem. 2011, 4, 361–377. [Google Scholar] [CrossRef] [Green Version]
  6. Gunatilake, S. Methods of removing heavy metals from industrial wastewater. Methods 2015, 1, 14. [Google Scholar]
  7. Han, H.; Rafiq, M.K.; Zhou, T.; Xu, R.; Mašek, O.; Li, X. A critical review of clay-based composites with enhanced adsorption performance for metal and organic pollutants. J. Hazard. Mater. 2019, 369, 780–796. [Google Scholar] [CrossRef]
  8. Burakov, A.E.; Galunin, E.V.; Burakova, I.V.; Kucherova, A.E.; Agarwal, S.; Tkachev, A.G.; Gupta, V.K. Adsorption of heavy metals on conventional and nanostructured materials for wastewater treatment purposes: A review. Ecotoxicol. Environ. Saf. 2018, 148, 702–712. [Google Scholar] [CrossRef] [PubMed]
  9. Bradl, H.B. Adsorption of heavy metal ions on soils and soils constituents. J. Colloid Interface Sci. 2004, 277, 1–18. [Google Scholar] [CrossRef] [PubMed]
  10. Uddin, M.K. A review on the adsorption of heavy metals by clay minerals, with special focus on the past decade. Chem. Eng. J. 2017, 308, 438–462. [Google Scholar] [CrossRef]
  11. Yu, F.; Li, Y.; Huang, G.; Yang, C.; Chen, C.; Zhou, T.; Zhao, Y.; Ma, J. Adsorption behavior of the antibiotic levofloxacin on microplastics in the presence of different heavy metals in an aqueous solution. Chemosphere 2020, 260, 127650. [Google Scholar] [CrossRef] [PubMed]
  12. Ajiboye, T.O.; Oyewo, O.A.; Onwudiwe, D.C. Simultaneous removal of organics and heavy metals from industrial wastewater: A review. Chemosphere 2021, 262, 128379. [Google Scholar] [CrossRef] [PubMed]
  13. Duan, C.; Ma, T.; Wang, J.; Zhou, Y. Removal of heavy metals from aqueous solution using carbon-based adsorbents: A review. J. Water Process. Eng. 2020, 37, 101339. [Google Scholar] [CrossRef]
  14. Gomaa, H.; Shenashen, M.A.; Elbaz, A.; Yamaguchi, H.; Abdelmottaleb, M.; El-Safty, S.A. Mesoscopic engineering materials for visual detection and selective removal of copper ions from drinking and waste water sources. J. Hazard. Mater. 2021, 406, 124314. [Google Scholar] [CrossRef] [PubMed]
  15. Bhatnagar, A.; Hogland, W.; Marques, M.; Sillanpää, M. An overview of the modification methods of activated carbon for its water treatment applications. Chem. Eng. J. 2013, 219, 499–511. [Google Scholar] [CrossRef]
  16. Hong, M.; Yu, L.; Wang, Y.; Zhang, J.; Chen, Z.; Dong, L.; Zan, Q.; Li, R. Heavy metal adsorption with zeolites: The role of hierarchical pore architecture. Chem. Eng. J. 2019, 359, 363–372. [Google Scholar] [CrossRef]
  17. Narayanan, S.; Tamizhdurai, P.; Mangesh, V.; Ragupathi, C.; Ramesh, A. Recent advances in the synthesis and applications of mordenite zeolite–review. RSC Adv. 2021, 11, 250–267. [Google Scholar] [CrossRef]
  18. Waheed, A.; Baig, N.; Ullah, N.; Falath, W. Removal of hazardous dyes, toxic metal ions and organic pollutants from wastewater by using porous hyper-cross-linked polymeric materials: A review of recent advances. J. Environ. Manag. 2021, 287, 112360. [Google Scholar] [CrossRef]
  19. Udayakumar, G.P.; Muthusamy, S.; Selvaganesh, B.; Sivarajasekar, N.; Rambabu, K.; Sivamani, S.; Sivakumar, N.; Maran, J.P.; Hosseini-Bandegharaei, A. Ecofriendly biopolymers and composites: Preparation and their applications in water-treatment. Biotechnol. Adv. 2021, 52, 107815. [Google Scholar] [CrossRef]
  20. Bhaladhare, S.; Das, D. Cellulose: A fascinating biopolymer for hydrogel synthesis. J. Mater. Chem. B 2022, 10, 1923–1945. [Google Scholar] [CrossRef]
  21. Liu, Y.; Ahmed, S.; Sameen, D.E.; Wang, Y.; Lu, R.; Dai, J.; Li, S.; Qin, W. A review of cellulose and its derivatives in biopolymer-based for food packaging application. Trends Food Sci. Technol. 2021, 112, 532–546. [Google Scholar] [CrossRef]
  22. Abdelhamid, H.N.; Mathew, A.P. Cellulose-based materials for water remediation: Adsorption, catalysis, and antifouling. Front. Chem. Eng. 2021, 3, 790314. [Google Scholar] [CrossRef]
  23. Klemm, D.; Heublein, B.; Fink, H.P.; Bohn, A. Cellulose: Fascinating biopolymer and sustainable raw material. Angew. Chem. Int. Ed. 2005, 44, 3358–3393. [Google Scholar] [CrossRef]
  24. Rana, A.K.; Scarpa, F.; Thakur, V.K. Cellulose/polyaniline hybrid nanocomposites: Design fabrication, and emerging multidimensional applications. Ind. Crops Prod. 2022, 187, 115356. [Google Scholar] [CrossRef]
  25. Akhlamadi, G.; Goharshadi, E.K. Sustainable and superhydrophobic cellulose nanocrystal-based aerogel derived from waste tissue paper as a sorbent for efficient oil/water separation. Process Saf. Environ. Prot. 2021, 154, 155–167. [Google Scholar] [CrossRef]
  26. Hajjaoui, H.; Soufi, A.; Boumya, W.; Abdennouri, M.; Barka, N. Polyaniline/Nanomaterial Composites for the Removal of Heavy Metals by Adsorption: A Review. J. Compos. Sci. 2021, 5, 233. [Google Scholar] [CrossRef]
  27. Md Salim, R.; Asik, J.; Sarjadi, M.S. Chemical functional groups of extractives, cellulose and lignin extracted from native Leucaena leucocephala bark. Wood Sci. Technol. 2021, 55, 295–313. [Google Scholar] [CrossRef]
  28. Bhatti, H.N.; Mahmood, Z.; Kausar, A.; Yakout, S.M.; Shair, O.H.; Iqbal, M. Biocomposites of polypyrrole, polyaniline and sodium alginate with cellulosic biomass: Adsorption-desorption, kinetics and thermodynamic studies for the removal of 2, 4-dichlorophenol. Int. J. Biol. Macromol. 2020, 153, 146–157. [Google Scholar] [CrossRef] [PubMed]
  29. Chen, J.H.; Xing, H.T.; Guo, H.X.; Li, G.P.; Weng, W.; Hu, S.R. Preparation, characterization and adsorption properties of a novel 3-aminopropyltriethoxysilane functionalized sodium alginate porous membrane adsorbent for Cr (III) ions. J. Hazard. Mater. 2013, 248, 285–294. [Google Scholar] [CrossRef] [PubMed]
  30. Li, R.; Liu, L.; Yang, F. Removal of aqueous Hg (II) and Cr (VI) using phytic acid doped polyaniline/cellulose acetate composite membrane. J. Hazard. Mater. 2014, 280, 20–30. [Google Scholar] [CrossRef]
  31. Kumar, N.; Kardam, A.; Jain, V.; Nagpal, S. A rapid, reusable polyaniline-impregnated nanocellulose composite-based system for enhanced removal of chromium and cleaning of waste water. Sep. Sci. Technol. 2020, 55, 1436–1448. [Google Scholar] [CrossRef]
  32. Liu, L.; Cao, J.; Huang, J.; Cai, Y.; Yao, J. Extraction of pectins with different degrees of esterification from mulberry branch bark. Bioresour. Technol. 2010, 101, 3268–3273. [Google Scholar] [CrossRef]
  33. Janaki, V.; Vijayaraghavan, K.; Oh, B.-T.; Ramasamy, A.; Kamala-Kannan, S. Synthesis, characterization and application of cellulose/polyaniline nanocomposite for the treatment of simulated textile effluent. Cellulose 2013, 20, 1153–1166. [Google Scholar] [CrossRef]
  34. Chen, B.; Zhao, H.; Chen, S.; Long, F.; Huang, B.; Yang, B.; Pan, X. A magnetically recyclable chitosan composite adsorbent functionalized with EDTA for simultaneous capture of anionic dye and heavy metals in complex wastewater. Chem. Eng. J. 2019, 356, 69–80. [Google Scholar] [CrossRef]
  35. Ben-Ali, S.; Jaouali, I.; Souissi-Najar, S.; Ouederni, A. Characterization and adsorption capacity of raw pomegranate peel biosorbent for copper removal. J. Clean. Prod. 2017, 142, 3809–3821. [Google Scholar] [CrossRef]
  36. Wang, X.; Liang, X.; Wang, Y.; Wang, X.; Liu, M.; Yin, D.; Xia, S.; Zhao, J.; Zhang, Y. Adsorption of Copper (II) onto activated carbons from sewage sludge by microwave-induced phosphoric acid and zinc chloride activation. Desalination 2011, 278, 231–237. [Google Scholar] [CrossRef]
  37. Beker, U.; Ganbold, B.; Dertli, H.; Gülbayir, D.D. Adsorption of phenol by activated carbon: Influence of activation methods and solution pH. Energy Convers. Manag. 2010, 51, 235–240. [Google Scholar] [CrossRef]
  38. Tri, N.L.M.; Thang, P.Q.; Van Tan, L.; Huong, P.T.; Kim, J.; Viet, N.M.; Phuong, N.M.; Al Tahtamouni, T. Removal of phenolic compounds from wastewaters by using synthesized Fe-nano zeolite. J. Water Process Eng. 2020, 33, 101070. [Google Scholar] [CrossRef]
  39. Abouzeid, R.E.; Khiari, R.; El-Wakil, N.; Dufresne, A. Current state and new trends in the use of cellulose nanomaterials for wastewater treatment. Biomacromolecules 2018, 20, 573–597. [Google Scholar] [CrossRef] [PubMed]
  40. Kumar, R.; Barakat, M.; Taleb, M.A.; Seliem, M.K. A recyclable multifunctional graphene oxide/SiO2@ polyaniline microspheres composite for Cu (II) and Cr (VI) decontamination from wastewater. J. Clean. Prod. 2020, 268, 122290. [Google Scholar] [CrossRef]
  41. Kayestha, R.; Hajela, K. ESR studies on the effect of ionic radii on displacement of Mn(II) bound to a soluble β-galactoside binding hepatic lectin. FEBS Lett. 1995, 368, 285–288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Lorenc-Grabowska, E. Effect of micropore size distribution on phenol adsorption on steam activated carbons. Adsorption 2016, 22, 599–607. [Google Scholar] [CrossRef] [Green Version]
  43. Ji, F.; Li, C.; Tang, B.; Xu, J.; Lu, G.; Liu, P. Preparation of cellulose acetate/zeolite composite fiber and its adsorption behavior for heavy metal ions in aqueous solution. Chem. Eng. J. 2012, 209, 325–333. [Google Scholar] [CrossRef]
  44. Rathi, B.S.; Kumar, P.S. Application of adsorption process for effective removal of emerging contaminants from water and wastewater. Environ. Pollut. 2021, 280, 116995. [Google Scholar] [CrossRef]
  45. Moussout, H.; Ahlafi, H.; Aazza, M.; Maghat, H. Critical of linear and nonlinear equations of pseudo-first order and pseudo-second order kinetic models. Karbala Int. J. Mod. Sci. 2018, 4, 244–254. [Google Scholar] [CrossRef]
  46. Saravanan, A.; Karishma, S.; Kumar, P.S.; Varjani, S.; Yaashikaa, P.; Jeevanantham, S.; Ramamurthy, R.; Reshma, B. Simultaneous removal of Cu (II) and reactive green 6 dye from wastewater using immobilized mixed fungal biomass and its recovery. Chemosphere 2021, 271, 129519. [Google Scholar] [CrossRef]
  47. Dada, A.; Olalekan, A.; Olatunya, A.; Dada, O. Langmuir, Freundlich, Temkin and Dubinin–Radushkevich isotherms studies of equilibrium sorption of Zn2+ unto phosphoric acid modified rice husk. IOSR J. Appl. Chem. 2012, 3, 38–45. [Google Scholar]
  48. Daraei, H.; Mittal, A.; Noorisepehr, M.; Daraei, F. Kinetic and equilibrium studies of adsorptive removal of phenol onto eggshell waste. Environ. Sci. Pollut. Res. 2013, 20, 4603–4611. [Google Scholar] [CrossRef]
  49. Zhang, D.; Huo, P.; Liu, W. Behavior of phenol adsorption on thermal modified activated carbon. Chin. J. Chem. Eng. 2016, 24, 446–452. [Google Scholar] [CrossRef]
  50. Bin-Dahman, O.A.; Saleh, T.A. Synthesis of carbon nanotubes grafted with PEG and its efficiency for the removal of phenol from industrial wastewater. Environ. Nanotechnol. Monit. Manag. 2020, 13, 100286. [Google Scholar] [CrossRef]
  51. Zhang, Q.; Han, Y.; Wu, L. Influence of electrostatic field on the adsorption of phenol on single-walled carbon nanotubes: A study by molecular dynamics simulation. Chem. Eng. J. 2019, 363, 278–284. [Google Scholar] [CrossRef]
  52. Dai, Y.; Zhang, N.; Xing, C.; Cui, Q.; Sun, Q. The adsorption, regeneration and engineering applications of biochar for removal organic pollutants: A review. Chemosphere 2019, 223, 12–27. [Google Scholar] [CrossRef]
  53. Khan, T.A.; Mukhlif, A.A.; Khan, E.A. Uptake of Cu2+ and Zn2+ from simulated wastewater using muskmelon peel biochar: Isotherm and kinetic studies. Egypt. J. Basic Appl. Sci. 2017, 4, 236–248. [Google Scholar] [CrossRef] [Green Version]
  54. Hossain, M.F.; Akther, N.; Zhou, Y. Recent advancements in graphene adsorbents for wastewater treatment: Current status and challenges. Chin. Chem. Lett. 2020, 31, 2525–2538. [Google Scholar] [CrossRef]
  55. Gupta, M.; Gupta, H.; Kharat, D. Adsorption of Cu (II) by low cost adsorbents and the cost analysis. Environ. Technol. Innov. 2018, 10, 91–101. [Google Scholar] [CrossRef]
  56. Ricou, P.; Lecuyer, I.; Le Cloirec, P. Removal of Cu2+, Zn2+ and Pb2+ by adsorption onto fly ash and fly ash/lime mixing. Water Sci. Technol. 1999, 39, 239–247. [Google Scholar] [CrossRef]
  57. Özer, A.; Özer, D.; Özer, A. The adsorption of copper (II) ions on to dehydrated wheat bran (DWB): Determination of the equilibrium and thermodynamic parameters. Process Biochem. 2004, 39, 2183–2191. [Google Scholar] [CrossRef]
  58. Ulmanu, M.; Marañón, E.; Fernández, Y.; Castrillón, L.; Anger, I.; Dumitriu, D. Removal of copper and cadmium ions from diluted aqueous solutions by low cost and waste material adsorbents. Water Air Soil Pollut. 2003, 142, 357–373. [Google Scholar] [CrossRef] [Green Version]
  59. Feng, N.; Guo, X.; Liang, S. Adsorption study of copper (II) by chemically modified orange peel. J. Hazard. Mater. 2009, 164, 1286–1292. [Google Scholar] [CrossRef] [PubMed]
  60. Rio, S.; Faur-Brasquet, C.; Le Coq, L.; Courcoux, P.; Le Cloirec, P. Experimental design methodology for the preparation of carbonaceous sorbents from sewage sludge by chemical activation—Application to air and water treatments. Chemosphere 2005, 58, 423–437. [Google Scholar] [CrossRef]
  61. Achak, M.; Hafidi, A.; Ouazzani, N.; Sayadi, S.; Mandi, L. Low cost biosorbent “banana peel” for the removal of phenolic compounds from olive mill wastewater: Kinetic and equilibrium studies. J. Hazard. Mater. 2009, 166, 117–125. [Google Scholar] [CrossRef] [PubMed]
  62. Mohammed, N.A.; Abu-Zurayk, R.A.; Hamadneh, I.; Al-Dujaili, A.H. Phenol adsorption on biochar prepared from the pine fruit shells: Equilibrium, kinetic and thermodynamics studies. J. Environ. Manag. 2018, 226, 377–385. [Google Scholar] [CrossRef]
  63. Gupta, A.; Balomajumder, C. Simultaneous adsorption of Cr (VI) and phenol onto tea waste biomass from binary mixture: Multicomponent adsorption, thermodynamic and kinetic study. J. Environ. Chem. Eng. 2015, 3, 785–796. [Google Scholar] [CrossRef]
  64. Allahkarami, E.; Dehghan Monfared, A.; Silva, L.F.O.; Dotto, G.L. Lead ferrite-activated carbon magnetic composite for efficient removal of phenol from aqueous solutions: Synthesis, characterization, and adsorption studies. Sci. Rep. 2022, 12, 10718. [Google Scholar] [CrossRef] [PubMed]
  65. Al Bsoul, A.; Hailat, M.; Abdelhay, A.; Tawalbeh, M.; Al-Othman, A.; Al-Taani, A.A. Efficient removal of phenol compounds from water environment using Ziziphus leaves adsorbent. Sci. Total Environ. 2021, 761, 143229. [Google Scholar] [CrossRef]
  66. Hasan, H.A.; Muhammad, M.H. A review of biological drinking water treatment technologies for contaminants removal from polluted water resources. J. Water Process Eng. 2020, 33, 101035. [Google Scholar] [CrossRef]
Figure 1. A proposed scheme for the PANI@WTP nanocomposite synthesis and application.
Figure 1. A proposed scheme for the PANI@WTP nanocomposite synthesis and application.
Nanomaterials 13 01014 g001
Figure 2. XRD pattern of WTP and PANI@WTP nanocomposite.
Figure 2. XRD pattern of WTP and PANI@WTP nanocomposite.
Nanomaterials 13 01014 g002
Figure 3. TEM images of (a,b) WTP and (c) PANI@WTP nanocomposite.
Figure 3. TEM images of (a,b) WTP and (c) PANI@WTP nanocomposite.
Nanomaterials 13 01014 g003
Figure 4. FTIR spectrum of WTP, PANI@WTP nanocomposite before and after pollutant adsorption.
Figure 4. FTIR spectrum of WTP, PANI@WTP nanocomposite before and after pollutant adsorption.
Nanomaterials 13 01014 g004
Figure 5. (a) Zeta potential analysis, (b) the effect of solution pH on adsorption of Cu(II) and phenol (concentration, 500 mg/L; adsorbent dosage, 0.02 g; volume, 20 mL; temp, 30 °C).
Figure 5. (a) Zeta potential analysis, (b) the effect of solution pH on adsorption of Cu(II) and phenol (concentration, 500 mg/L; adsorbent dosage, 0.02 g; volume, 20 mL; temp, 30 °C).
Nanomaterials 13 01014 g005
Figure 6. The kinetic plots for the adsorption of (a) Cu(II) and (b) phenol onto the PANI@WTP nanocomposite (concentration, 500 mg/L; adsorbent dosage, 0.02 g; volume, 20 mL; pH 5.2 for Cu(II), pH, 5 for phenol; temp, 30 °C).
Figure 6. The kinetic plots for the adsorption of (a) Cu(II) and (b) phenol onto the PANI@WTP nanocomposite (concentration, 500 mg/L; adsorbent dosage, 0.02 g; volume, 20 mL; pH 5.2 for Cu(II), pH, 5 for phenol; temp, 30 °C).
Nanomaterials 13 01014 g006
Figure 7. Adsorption isotherm plots for (a) Cu(II) and (b) phenol onto PANI@WTP composite (adsorbent dosage, 0.02 g; volume, 20 mL; contact time, 5 h; temp, 30 °C; pH 5.2 for Cu(II), pH 5 for phenol solution).
Figure 7. Adsorption isotherm plots for (a) Cu(II) and (b) phenol onto PANI@WTP composite (adsorbent dosage, 0.02 g; volume, 20 mL; contact time, 5 h; temp, 30 °C; pH 5.2 for Cu(II), pH 5 for phenol solution).
Nanomaterials 13 01014 g007
Figure 8. Regeneration of PANI@WTP nanocomposite (concentration, 500 mg/L; volume, 20 mL; adsorbent mass, 0.02 g; pH 5 for phenol, 5.2 for Cu(II); temp, 30 °C).
Figure 8. Regeneration of PANI@WTP nanocomposite (concentration, 500 mg/L; volume, 20 mL; adsorbent mass, 0.02 g; pH 5 for phenol, 5.2 for Cu(II); temp, 30 °C).
Nanomaterials 13 01014 g008
Figure 9. Adsorption of Cu(II) and phenol from synthetic tap and groundwater using PANI@WTP (a) 20 mg/L and (b) 100 mg/L (volume, 20 mL; adsorbent mass, 0.02 g; pH 5 for phenol, 5.2 for Cu(II); temp, 30 °C).
Figure 9. Adsorption of Cu(II) and phenol from synthetic tap and groundwater using PANI@WTP (a) 20 mg/L and (b) 100 mg/L (volume, 20 mL; adsorbent mass, 0.02 g; pH 5 for phenol, 5.2 for Cu(II); temp, 30 °C).
Nanomaterials 13 01014 g009
Table 1. Adsorption kinetics parameters for Cu(II) and Phenol onto PANI@WTP nanocomposite.
Table 1. Adsorption kinetics parameters for Cu(II) and Phenol onto PANI@WTP nanocomposite.
Kinetic ModelParametersCu(II)Phenol
Pseudo-first-orderqe(exp) (mg g−1):397320
qe(cal) (mg g−1):377.499308.525
k1(min−1):0.09200.0726.
R2:0.61950.9431
Pseudo-second-orderqe(cal) (mg g−1):396.41328.370
k2(g mg−1 min−1):0.339 × 10−30. 285 × 10−3
R2:0.78830.9746
Elovich modela(mg g−1 min−1):425.615132.691
β(mg g−1):0.01950.0206
R2:0.94850.9121
Table 2. Adsorption isotherm parameters for Cu(II) and phenol onto PANI@WTP nanocomposite.
Table 2. Adsorption isotherm parameters for Cu(II) and phenol onto PANI@WTP nanocomposite.
Isotherm ModelParametersCu(II)Phenol
Langmuirqm(mg g−1):605.204501.234
KL(L mg−1):0.01350.0062
R2:0.94360.9647
Freundlichn:2.20121.8580
Kf(mg g−1) (mg L−1)−1/n:37.15214.591
R2:0.86990.9061
TemkinBt(J mg−1)27.20329.083
Kt(L mg−1):0.35900.1136
R2:0.84110.9113
Table 3. Comparison of maximum adsorption capacity of Cu(II) and phenol with other materials.
Table 3. Comparison of maximum adsorption capacity of Cu(II) and phenol with other materials.
Experimental Conditions
AdsorbateAdsorbentspHConc. (mg/L)Contact Time (min)qe(mg g−1)Ref.
Cu(II)Musk melon75012078.74[53]
Banana peel6-144028.57[54]
Sugarcane bagasse5106088.9[55]
Fly ash5500-207.3[56]
Wheat bran51003051.5[57]
Cellulose pulp waste61001804.98[58]
Orange peel5100180289[59]
Carbon510014577-83[60]
PANI@WTP5.2500240605.20This study
PhenolBanana peel730180688.9[61]
Biochar6.55027026.738[62]
Tea waste biomass75014409.487[63]
Lead ferrite-MAC225060158.9[64]
Ziziphus leaves62024015[65]
PANI@WTP5500180501.23This study
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Doyo, A.N.; Kumar, R.; Barakat, M.A. Facile Synthesis of the Polyaniline@Waste Cellulosic Nanocomposite for the Efficient Decontamination of Copper(II) and Phenol from Wastewater. Nanomaterials 2023, 13, 1014. https://doi.org/10.3390/nano13061014

AMA Style

Doyo AN, Kumar R, Barakat MA. Facile Synthesis of the Polyaniline@Waste Cellulosic Nanocomposite for the Efficient Decontamination of Copper(II) and Phenol from Wastewater. Nanomaterials. 2023; 13(6):1014. https://doi.org/10.3390/nano13061014

Chicago/Turabian Style

Doyo, Ahmed N., Rajeev Kumar, and Mohamed A. Barakat. 2023. "Facile Synthesis of the Polyaniline@Waste Cellulosic Nanocomposite for the Efficient Decontamination of Copper(II) and Phenol from Wastewater" Nanomaterials 13, no. 6: 1014. https://doi.org/10.3390/nano13061014

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