Serious environmental deterioration caused by synthetic waste plastics, and the pollution of freshwater resources are the most alarming and marked challenges of the 21st century. Therefore, immense scientific efforts are being made towards the management of waste plastics and treatment of polluted water. The current study reports on the utilization of waste polyethylene terephthalate (wPET) and waste polystyrene (wPS) for fabrication of activated carbon (AC) and its application for the removal of hazardous polycyclic aromatic hydrocarbons (PAHs) pollutants from water. AC was prepared from wPET and wPS by carbonization under a N2 atmosphere followed by chemical activation with 1 M KOH and 1 M HCl. The AC was characterized by scanning electron microscopy, surface area analysis, and Fourier transform infrared spectroscopy. Adsorption of PAHs from aqueous solutions through AC was examined by batch adsorption tests. The optimum parameters for maximum adsorption of PAHs were found to be: initial PAHs concentration 40 ppm, 2 h contact time, pH 3, 5, and 7, 50 °C temperature and adsorbent dose of 0.8 g. Kinetic and isotherm models were applied to evaluate the adsorbent capacity for PAHs adsorption. The kinetic study shows that the adsorption of these PAHs onto AC follows pseudo-second-order kinetics. The experimental results demonstrated that the Langmuir isotherm model best fitted the data. The thermodynamic factors calculated such as entropy change (ΔS°), enthalpy change (ΔS°) and free energy change (ΔG°) show that the adsorption process is non-spontaneous and endothermic in nature. Results were also compared with the efficiencies of some commercial adsorbents used in practice. This examination revealed that the novel plastic-derived AC possesses a great potential for elimination and recovery of PAH elimination from industrial wastewater.

  • AC was prepared from wPET and wPS by carbonization followed by chemical activation.

  • Adsorption of PAHs from aqueous solutions through AC was examined by batch adsorption tests.

  • PAH adsorption from the aqueous solutions follows pseudo-second-order kinetics.

  • The experimental data are best fitted to the Langmuir isotherm model.

  • AC possesses a large potential for elimination and recovery of PAHs from wastewater.

Graphical Abstract

Graphical Abstract
Graphical Abstract

PAHs are recalcitrant and toxic organic pollutants, and consist of two or more fused benzene rings in linear, cluster or angular arrangements, and are associated with low water solubility (Johnsen & Karlson 2005). The PAHs are released from anthropogenic sources as well as from some natural sources. Compared with anthropogenic sources, the natural sources of PAHs in the environment are insignificant (Witt 1995; Charlesworth et al. 2002). Various natural sources include oil seepages, thermal geological reactions, volcanic eruptions, and prairie and natural forest fires (Wright & Welbourn 2002; Qamar et al. 2017; Hidayat et al. 2019). Among the different anthropogenic sources accountable for the PAH emissions, residential waste burning, agricultural processing, vehicular emissions, military operations, coal tar production, asphalt and shale oil production, coal conversion, metal working, petroleum processing, steel and iron foundry works, aluminum smelting and coke production and incomplete combustion are notable (Ravindra et al. 2008; Hall et al. 2009). PAHs are commonly used in lubricating materials, wood preserving, agricultural and pharmaceutical products, resins, dyes, pesticides, and the manufacture of pigments (Abdel-Shafy & Mansour 2016). PAHs are ubiquitous in tar deposits, coal and oil, and PAHs mostly in the aquatic system have originated from contaminated sediment, atmosphere deposition, and accidental leakage (USEPA 1994; Blanchard et al. 2004).

PAHs can be categorized into two major groups, including high-molecular-weight PAHs which contain six, five or four benzene rings, and low-molecular-weight PAHs containing three or two benzene rings (Wenzl et al. 2006; Waqas et al. 2014). Owing to their mutagenic and carcinogenic characteristics, high lipophilicity, persistence, and hazardous nature, PAHs are of great concern in the aquatic system (Sweetman & Jones 2000; Hidayat et al. 2019). Generally, the carcinogenicity of PAHs increases with rising molecular weight (Eisler 1987). PAHs can cause both long and transient wellbeing impacts, the long-term wellbeing impacts include symptoms of asthma, inhalation difficulties, lung abnormalities, damage to liver and kidney, cataracts and a suppressed immune system (Abdel-Shafy & Mansour 2016). The short-term health effects of PAHs in humans are not clear, in any case occupational exposure to PAHs in high levels may bring about transient impacts, for example diarrhea, eye irritation, vomiting, confusion and nausea (Abdel-Shafy & Mansour 2016). Various PAHs have been found to occur in plants as well as animal tissues (USEPA 2008), water bodies (Chang et al. 2004a, 2004b; Balati et al. 2015), sediments and soil (Vane et al. 2014; Wang et al. 2015) and in air (Liu et al. 2015), which signifies their presence in the environment. Since PAHs are non-biodegradable and persistent, they consequently enter the food chain via soil and move to the human body by dermal contact, inhalation, or ingestion (Abdel-Shafy & Mansour 2016). Once entering the body, PAHs can go crosswise over cell membranes and are effectively adsorbed into cells, where the immune system changes PAHs to epoxide hydrolase and diolepoxides, reacting with DNA and preventing its production (Kaya et al. 2013).

The European Union and USEPA have listed 16 PAHs among the concerned major pollutants. The occurrence and fate of major PAHs in conventional activated sludge in a wastewater treatment plant have been extensively studied by researchers (Manoli & Samara 1999). Lower levels of various PAHs have been found to occur in wastewaters treated by various processes (Busetti et al. 2006). Other mechanisms must be considered, such as air-stripping, chemical degradation, advection, or volatilization, which help to remove PAHs, depending on the system operating conditions, process design as well as wastewater characteristics (Manoli & Samara 1999). Over time, different methods have been developed to cater for the removal of PAHs from contaminated water as well as soil to reduce the possible risk of PAHs to human health and the environment. Phytoremediation, biological, thermal, chemical, and physical processes (including phytodegradation, phytostabilisation, phytoextraction, rhizofiltration, bioaugmentation, biostimulation, natural attenuation, chemical precipitation, electrolysis, adsorption, photolysis, ion exchange, oxidation, radio frequency heating, thermal desorption, and incineration methods) are the main treatment techniques for PAH contaminated water, sediment, and soil (Zhou et al. 2013; Hu et al. 2014; Li et al. 2014; Peng et al. 2014; Sun et al. 2014; Wang et al. 2014; Li et al. 2015). Although, most of these strategies have some disadvantages, for example, complicated operating procedures, high maintenance as well as investment costs. In addition, some of these treatment techniques result in the production of secondary by-products and some of them are mutagenic (e.g., haloacetic acid, trihalomethanes) and carcinogenic, which further increases the adverse effects to public health (Shih & Lederberg 1976; Comninellis et al. 2008; Oller et al. 2011).

Among all the treatment methods that have been developed, adsorption is one of the most relevant and promising methods to remove organic and inorganic micro-pollutants (Saucier et al. 2017; Umpierres et al. 2018; Mironyuk et al. 2019; Tatarchuk et al. 2019). This process has many advantages: (i) probability of adsorbent regeneration and reuse; (ii) environmentally benign; (iii) low investment cost; (iv) toxicity removal; (v) vigorous continuous and batch processes; (vi) handling microlevels of pollutants; and (vii) simple to operate and to design (Torrellas et al. 2016, Bhatnagar & Anastopoulos 2017). Various adsorbents used for this purpose include AC (Dabek & Ozimina 2009), zeolite (Chang et al. 2004a, 2004b), bottom ash (Gupta et al. 2012), Curcuma caesia-based AC (Arora et al. 2020), Chenopodium album ash (Arora et al. 2019a), ash from the black turmeric rhizome (Asha et al. 2021), iron-based metal organic frameworks (Arora et al. 2019b), Fe-benzene dicarboxylic acid metal organic frameworks (Soni et al. 2020), solid residues from agricultural activities such as pine bark and ash waste (Li et al. 2010; Pérez-Gregorio et al. 2010), rice husk (Hidayat et al. 2019), sugar cane bagasse, green coconut shells, chitin, and chitosan (Crisafully et al. 2008), as well as other organic-free hydrophilic minerals such as talc, silica and alumina (Su et al. 2006; Müller et al. 2007) and numerous materials such as plant-derived material, sewage sludge, shells, petroleum coke and bituminous coal have been used for the production of AC (Awoyemi 2011; Gupta 2015) through carbonization followed by activation processes. AC is produced depending on its particular purpose in the form of granules, powder, or briquette. The more common form of AC in practice includes powder and granular form for PAH sorption (Paune et al. 1998; Ania et al. 2007; Valderrama et al. 2008). The development of a low-cost and more efficient adsorbent has led to continued research interests in this field.

A tremendous increase in the usage of plastic over the past five decades has been observed worldwide. Due to its inappropriate dumping and waste management, the increased production and application of plastic is threatening the environment. The unintentional as well as unwarranted disposal of conventional plastic in water reservoirs has contributed to the decline in the population of marine organisms (O'Brine & Thompson 2010). Due to its non-biodegradable and stable chemical structure, plastics stay longer in the environment, constituting a large part of the solid waste. Some techniques have been suggested for handling the growing plastic solid wastes, such as its conversion into liquid or gaseous fuels, recycling, incineration and land filling, etc. (Ilyas et al. 2018). All these techniques have been critically reviewed, which revealed that each technique has its own shortcomings, such as in land filling, the non-biodegradable materials dumped on the land cause soil pollution, recycling is an expensive and energy-requiring process, whereas incineration and conversion involve the generation of more hazardous emissions, such as CO2 and persistent organic pollutants (POPs) etc. (Ilyas et al. 2018). However, a more practical and environmentally friendly approach is the utilization of plastic wastes in mitigating water pollution by utilization in carbonized form as adsorbents for the elimination of different toxic substances from wastewater (Ikram et al. 2020).

In the current study, AC was derived from waste polyethylene terephthalate (wPET-AC) and waste polystyrene (wPS-AC) by carbonization followed by its chemical activation, and was utilized for the adsorption of PAHs from water. This study aimed to investigate the potential of utilizing waste PET- and PS-derived AC for the removal of various PAHs from wastewater as a low-cost adsorbent and mitigate the issue of plastic solid waste management. The characterization of AC, optimization of adsorption parameters, adsorption kinetics, and adsorption isotherms have also been discussed.

Sample collection

Standard solutions of PAHs Mix-13 (USEPA 16) consisting of benzo(a)anthracene (BaA; 10 μg/L), benzo(a)pyrene (BaP; 10 μg/L), benzo(k)fluoranthene (BkF; 10 μg/L), benzo(b)fluoranthene (BbF; 10 g/L), dibenzo(ah)anthracene (D[ah]A; 10 μg/L), benzo(ghi)perylene (B[ghi]P; 10 μg/L),indeno(123-cd)pyrene (Ind; 10 μg/L), pyrene (Pyr; 10 μg/L), chrysene (Chr; 10 μg/L), fluoranthene (Flu; 10 μg/L), acenaphthene (Ace; 10 μg/L), acenaphthylene (Acy; 100 μg/L), anthracene (Ant; 10 μg/L), phenanthrene (Ph; 10 μg/L), fluorene (Fl; 10 μg/L) and naphthalene (Nap; 100 μg/L) purity >99% was purchased from the Supelco Company, USA. On a daily basis, the working standard solutions were prepared by diluting the PAHs Mix-13 standard solution in 2-propanol. Thus, to avert photochemical degradation, the sample extracts and standard solutions were stored in brown glass vials at 4 °C for further study.

Synthesis of AC from waste plastic by carbonization

Samples of wPET and wPS were collected from street waste bins. The wPET and wPS samples were washed, dried, and shredded in to small pieces. The waste plastic samples were converted into AC by carbonization in a specially designed steel reactor under an inert atmosphere at high temperature. About 5 g of wPET or wPS were taken in the steel reactor and inserted in the tube furnace, the nitrogen supply was connected to the reactor and heated to 600 °C for one hour, as shown in Figure 1. The AC was scrapped from the reactor and activated chemically by treatment with 1 M KOH and 1 M HCl alternately using a procedure reported in the literature (Lian et al. 2011). The solid carbon was finally ground to a fine powder and passed through a sieve with pore size 250 μm and then stored in glass vials for further study.

Figure 1

Diagrammatic presentation of the AC preparation in the laboratory.

Figure 1

Diagrammatic presentation of the AC preparation in the laboratory.

Close modal

Characterization of AC

The AC samples synthesized from wPET and wPS were characterized by various instrumental analyses. The surface morphology of AC was studied by a SEM (JSM-5910, JEOL, Japan). The analysis of the functional groups of AC was performed using FTIR (Spectrum Two, S.No: 103385) and the specific surface area was analyzed using a pore size and surface area analyzer (NOVA2200e, Quantachrome, USA). The zeta potential of the wPET-AC and wPS-AC was obtained using a Zeta Meter 3.0 (Zeta Meter Inc.).

Adsorption experiments

Adsorption of PAHs from water samples over wPET-AC and wPS-AC was studied in batch mode adsorption experiments. About 100 ml of water sample containing PAHs were taken into a 500 ml Erlenmeyer flasks at 24 ± 2 °C, definite amounts of AC adsorbent were added to it and stirred on a magnetic stirrer with 150 rpm speed until equilibrium was reached. Several variables were measured in the study, including dosage of AC (0.2, 0.4, 0.6, and 0.8 g/100 ml), contact time (0.25, 0.5, 1 and 2 h), PAHs concentrations (10, 20, 30, and 40 ppm) and pH of solutions (3, 5, 7, 9, and 11), pH was adjusted by adding 0.1 M HCl or NaOH to the solution as required.

Analyses and calculations

The PAHs concentration in the treated samples was examined by using a UV spectrometer (Mansouri et al. 2020). Adsorption efficiency at equilibrium (qe) was calculated as the following (Chang et al. 2004a, 2004b):
where, Co is the initial concentration of PAHs (μg/L), Ci is the remaining concentration of PAHs at equilibrium in the solution (μg/L), V represents the volume of the solution (L) and M is the AC mass (g).
Adsorption efficiency at a specific time (qt) was calculated as the following (Crisafully et al. 2008):
where, Ct is the liquid-phase concentration at that specific time t (μg/L). Percent removal efficiency (%MR) was calculated using the following relations (Björklund & Li 2015):

The adsorption kinetics was investigated by applying pseudo-first-order (Ho & McKay 1998), pseudo-second-order (Chen et al. 2010), and intraparticle diffusion kinetic models (Weber & Morris 1963). Likewise, the adsorption mechanism was studied by applying Freundlich (Yao et al. 2014), and Langmuir (Lawrinenko 2014) isotherms to the adsorption data. The equations for these models are given in Table 1. All calculations were conducted using Excel 2010 software (Microsoft Office).

Table 1

Adsorption models used in the study

Kinetic modelsEquation
Pseudo-first-order  where, 
 • At equilibrium the quantity of PAHs adsorbed is presented by qe (μg g−1
 • At time t the quantity of PAHs adsorbed is presented by qt, and 
 • The rate constant is given by k1 [1/h]. 
Pseudo-second-order  where, 
 • K2 is the rate constant [g/(μg·h)] for second-order kinetics. 
 • At equilibrium the sorption capacity is presented by qe. 
Intraparticle diffusion qt = Kid(t)0.5 + C where, 
 • Kid (h−1) is rate of reaction for intraparticle diffusion process, and 
 • C is the constant related to the thickness of boundary layer (μg g−1). 
Equilibrium models  
Freundlich  
Langmuir  where, 
 • qe represents the PAHs adsorbed concentration over the carbonaceous adsorbents (μg g−1), 
 • Ce is the PAHs equilibrium concentration (μg L−1), 
 • While Kf and n represent the adsorption coefficient and constant for the model, respectively. 
Kinetic modelsEquation
Pseudo-first-order  where, 
 • At equilibrium the quantity of PAHs adsorbed is presented by qe (μg g−1
 • At time t the quantity of PAHs adsorbed is presented by qt, and 
 • The rate constant is given by k1 [1/h]. 
Pseudo-second-order  where, 
 • K2 is the rate constant [g/(μg·h)] for second-order kinetics. 
 • At equilibrium the sorption capacity is presented by qe. 
Intraparticle diffusion qt = Kid(t)0.5 + C where, 
 • Kid (h−1) is rate of reaction for intraparticle diffusion process, and 
 • C is the constant related to the thickness of boundary layer (μg g−1). 
Equilibrium models  
Freundlich  
Langmuir  where, 
 • qe represents the PAHs adsorbed concentration over the carbonaceous adsorbents (μg g−1), 
 • Ce is the PAHs equilibrium concentration (μg L−1), 
 • While Kf and n represent the adsorption coefficient and constant for the model, respectively. 

Characterization of wPET-AC and wPS-AC

SEM analysis

The AC synthesized from wPET and wPS, i.e., wPET-AC and wPS-AC were characterized by SEM, FTIR, and SSA analysis. The SEM micrographs of wPET, wPET-AC, wPS and wPS-AC are given in Figure 2. In Figure 2(a) SEM images highlighting the porosity of wPET are shown. The surface features of wPET and the region of both external as well as internal structures and in addition the unfilled spaces between the particles can be seen (Ilyas et al. 2019; Ilyas et al. 2021). The SEM micrographs of wPET-AC (Figure 2(b)), show a stretched but uneven morphology, with some fine particles agglomerated on the surface. The surface also seems flaky, which indicates the highly porous structure of the AC. The micrograph of wPS displayed in Figure 2(c) shows a spongy structure with wide pores openings, formed from interconnected polymer films (Ilyas et al. 2019). The SEM micrograph of wPS-AC (Figure 2(d)) displays an irregular and rough morphology with numerous cavities and mounds on the surface. Some fine particles are agglomerated on the surface, these agglomerated particles may be spherical carbon black structures, whereas on the surface fissures and cracks are also visible. The SEM analysis suggests that wPET-AC exhibits better porosity than wPS-AC.

Figure 2

SEM images of (a) wPET, (b) wPET-AC, (c) wPS and (d) wPS-AC.

Figure 2

SEM images of (a) wPET, (b) wPET-AC, (c) wPS and (d) wPS-AC.

Close modal

FTIR spectra

Figure 3 shows FTIR spectra of wPS, wPS-AC, wPET, and wPET-AC. The FTIR spectrum of wPS shows major peaks centered at around 3,000 and 2,800 cm−1 corresponding to aromatic C-H and methylene C-H bonds. Multiple peaks positioned at 1,630 and 1,500 cm−1 indicate aromatic C = C vibrations, and a peak around 1,300 cm−1 indicates methylene C-H vibration, all these configurations are in conformity with the structure of polystyrene (Herman et al. 2015; Subramani & Sepperumal 2016). The FTIR spectrum of wPS-AC (Figure 3(a)), shows the major peaks positioned at 3,028 cm−1 (aromatic C-H, stretching), 1,599 cm−1 (aromatic C = C, stretching), 1,451 cm−1 (CH2 stretching), 1,017 cm−1 (C-O, stretching), 696 cm−1 (C-H, stretching) and 540 (C-H, stretching) cm−1, which indicates that the resultant AC consists of aromatic functionalities. The characteristic infrared absorption bands of wPET (Figure 3(b)) at 1,696 cm−1 (C = O, stretch), 1,260 cm−1 (C = C, stretch), 1,092 cm−1 (C-O, stretching vibration), 1,010 cm−1 (C–O–C, stretching), 720 cm−1 (C-H, stretch), and 521 cm−1 (C-H, stretch) are clearly observed (Tantawy et al. 2014). The FTIR spectrum of wPET-AC (Figure 3(b)) showed that the wPET-AC had functional groups of wPET. The noticeable absorption peaks were at 3,414 cm−1 which might be due to stretching vibrations of –OH, 1,696 cm−1 (C = O, stretch), 1,260 cm−1 (C = C, stretch), 1,092 cm−1 (C-O, stretching vibration), 1,010 cm−1 (C–O–C, stretching), 720 cm−1(C-H, stretch) and 521 cm−1 (C-H, stretch). These results suggest that the AC derived from wPET contains various oxy functionalities, which may have resulted from surface oxidation of AC.

Figure 3

FTIR spectra of (a) wPET and wPET-AC and (b) wPS and wPS-AC.

Figure 3

FTIR spectra of (a) wPET and wPET-AC and (b) wPS and wPS-AC.

Close modal

Surface area analysis

The surface characteristics such as total pore volume (Vt), average pore radius (RP), BET surface area (SBET), and BJH surface area (SBJH) of the AC are shown in Table 2. Results show that wPET-AC exhibits SBET of 65.35 m2 g−1, SBJH of 11.02 m2 g−1, Vt of 0.009 cm3 g−1 and an RP of 15.02 Å. In the case of wPS-AC, the SBJH was found to be 14.68 m2 g−1, the SBET 50.14 m2 g−1, with RP of 14.09 A0 and Vt of 0.010 cm3 g−1. It can be observed that the surface area of the wPS-AC is relatively smaller than that of wPET-AC, however, the surface areas of both the ACs are suitable for sufficient adsorption. The BET surface areas of the wPS-AC and wPET-AC were comparable with the literature reports, for example, the SBET values obtained for bentonites by other researchers, which include 31.5 m2 g−1 reported by Xifang et al. (2007), 56 m2 g−1 obtained by Ding et al. (2009) and 34.1 m2 g−1 reported by Guerra et al. (2013), wPET-AC was found to have a higher surface area, which is desirable for efficient sorption.

Table 2

Surface area and pore dimensions of wPET-AC and wPS-AC.

AdsorbentBJH surface area (m2 g−1)BET surface area (m2 g−1)RP (A0)Vt (cm3 g−1)
wPET-AC 11.02 65.35 15.02 0.009 
wPS-AC 14.68 50.14 14.09 0.010 
AdsorbentBJH surface area (m2 g−1)BET surface area (m2 g−1)RP (A0)Vt (cm3 g−1)
wPET-AC 11.02 65.35 15.02 0.009 
wPS-AC 14.68 50.14 14.09 0.010 

Adsorption of PAHs over AC

The adsorption of various 16 PAHs was studied over AC derived from wPET and wPS from aqueous solution. Initially, the PAHs adsorptions over the AC examinations were carried out for 90 min contact time and adsorbent to solution ratio of 0.3 g/100 mL at a temperature of 24 ± 2 °C, the % adsorptions of 16 PAHs are given in Figure 4. Results indicate that the wPET-AC showed a greater tendency for the 16 PAHs adsorption than that of the wPS-AC. This is reliable with its larger average pore radius observed by BET analysis. In the case of wPET-AC the values of % adsorption of the 16 PAHs ranged from 53.80 to 92.44. The lowest value of 53.80% was noted for Nap while the highest value of 92.44% was noted for Pyr. Similarly, the adsorption values of these 16 PAHs ranged 44.3–82.3% for wPS-AC. The wPS-AC has shown the highest removal efficiency for B[ghi]P, while the lowest value for Nap from aqueous media. The data were compared with the previous work done by researchers for PAHs removal while utilizing AC. The adsorption test of B[ghi]P, Inp, BaA, BkF, BbF, Chr, BaP, Pyr, Fl, Ant and Ph was carried out by Amstaetter et al. (2012) using AC and achieved better adsorption efficiency for Ph (95%), Ant (97.5%), Flu (96.5%), Pyr (98%), BaP (88%), Cry (94.5%), BbF (86%), BkF (88%), BaA (94.5%), InP (73%), B[ghi]P (69%). Similarly, a high adsorption capacity using powder AC for PAHs (B[ghi]P (84%), Inp (85%), BaA (93%), BkF (92%), BbF (93%), Chr (95%), BaP (88%), Pyr (96%), Fl (95%), Ant (89%) and Ph (82%)) was attained from the contaminated soil (pore water) (Hale et al. 2012). Furthermore, Oleszczuk et al. (2012) studied the adsorption capacity of PAHs (B[ghi]P, BaP, Phr, Flu, Ph) from sewage sludge, which was 95% using AC. Hence, in the current study, the performance of the AC is also better. Further experiments were carried out under various conditions to investigate the most favorable adsorption factors.

Figure 4

% Adsorption efficiency of wPET-AC and wPS-AC for 16 PAHs adsorption.

Figure 4

% Adsorption efficiency of wPET-AC and wPS-AC for 16 PAHs adsorption.

Close modal

Effect of process parameters

After studying the efficiency of wPET-AC and wPS-AC, it was found that they showed better adsorption efficiency. In order to investigate the optimum conditions required for maximum adsorption of PAHs, the adsorption was investigated under different conditions of concentration of PAHs, pH, contact time, temperature, and adsorbent dose.

Effect of adsorbent dosage

The availability and accessibility of adsorption sites is represented by the quantity of adsorbent provided for adsorption (Abbas et al. 2013). The effect of the adsorbent amounts was investigated with a 100 mL solution of PAHs and varying amounts of each adsorbent tanning from 0.2 to 0.8 g. Figure 5(a) and 5(b) shows the PAHs adsorption at varying amounts of wPET-AC and wPS-AC. The increased amount of AC increases the availability of surface active sites for adsorption, therefore a linear increase in the adsorption of PAHs is observed with raining AC dosage. The PAHs adsorption is not considerably affected by an extra increase in the adsorbent dose more than 0.8 g owing to the chance of aggregation, which restricts the number of active-site surfaces available for adsorption (Liu et al. 2014). It can be further observed from the data that, at low adsorbent dosage, the removal of all PAHs was lower. However, at optimum adsorbent dosage, the % removal of all PAHs was almost uniform. These results further show that the optimum dose of adsorbent in the current study is 0.8 g, which is higher than that reported by Awe et al. (2020) and Rad et al. (2014) and lower than those reported by Awoyemi (2011). However, this value was in agreement with Njoku et al. (2014).

Figure 5

Effect of adsorbent dosage (a, b) and temperature (c, d) on the % adsorption of 16 PAHs over the wPET-AC and wPS/AC.

Figure 5

Effect of adsorbent dosage (a, b) and temperature (c, d) on the % adsorption of 16 PAHs over the wPET-AC and wPS/AC.

Close modal

Effect of temperature

Variation in temperature can affect the adsorption of PAHs onto the adsorbents from aqueous solution. The removal of PAHs from aqueous solution by wPET-AC and wPS-AC at different temperatures ranging from 20 to 50 °C is displayed in Figure 5(c) and 5(d), which shows that the removal of all PAHs increases as the temperature increased. This also suggests that the adsorption of PAHs over the wPET-AC and wPS-AC surfaces is endothermic. Various other researchers have reported the endothermic nature of the adsorption process using different adsorbents such as hazelnut shell activated carbon (Kobya 2004), coconut shell AC (Qiu et al. 2012), agricultural by-products (Chen et al. 2010) and hen feathers (Mittal et al. 2007).

Effect of contact time

The effects of contact time on PAHs adsorption onto wPET-AC and wPS-AC were investigated by carrying out adsorption experiments for different contact times, i.e., 0.25, 0.5, 1 and 2 hours. For all 16 PAHs, the adsorption on the AC increases up to 2 hr and becomes constant thereafter (Figure 6(a) and 6(b)). This pattern might be ascribed to a large number of active sites available on the surface of the adsorbent initially. After a contact time of 2 hr, the majority of active sites present on AC have been occupied and equilibrium is attained. It was observed that prolonged contact time had no effect on the PAHs adsorption. Thus, the optimum contact time was considered to be 2 hr. The absence of any particular trend in the adsorption of the studied PAHs is presumably due to the fact that different factors, for example, the nature of the adsorbent, chemical properties of the adsorbate, molecular size, solubility and hydrophobic character influence the adsorption process (Canzano et al. 2014). The adsorption characteristics, for example, surface functional groups, number of active sites, etc. also influence the adsorption process (Gupta & Gupta 2016). Previous studies have also reported similar observations for the adsorption of various PAHs on different adsorbents in that the adsorption of PAHs increases with increasing contact time and after equilibrium is attained (Awoyemi 2011; Gupta 2015).

Figure 6

(a, b). Effect of contact time on the % adsorption of 16 PAHs over the wPET-AC and wPS-AC.

Figure 6

(a, b). Effect of contact time on the % adsorption of 16 PAHs over the wPET-AC and wPS-AC.

Close modal

Effect of initial concentration of PAHs

The adsorption of PAHs was studied as a function of initial concentration at varying initial PAH concentrations (10–40 ppm). Adsorption of PAHs was found to be increased when increasing the initial PAHs concentration (Figure 7(a) and 7(b)). This is because of the way that increasing of the initial PAHs concentration may result in the availability of PAHs to adsorb over the adsorbent surface active sites. Similar behavior has been observed in previous studies and arises from the effect of interactions between the initial PAHs concentration and adsorbent (Ping et al. 2006; Awoyemi 2011; Gupta 2015). Gong et al. (2007) indicated that a high initial PAH concentration has positive influences, which may be due to the availability of a larger number of PAHs to interact with the active sites of the AC.

Figure 7

Effect of initial concentration (a, b) and pH (c, d) on the % adsorption of 16 PAHs over the wPET-AC and wPS/AC.

Figure 7

Effect of initial concentration (a, b) and pH (c, d) on the % adsorption of 16 PAHs over the wPET-AC and wPS/AC.

Close modal

Effect of pH

pH is an important parameter that controls the PAHs uptake from aqueous solution. The influence of pH variation on the adsorption of PAHs using wPET-AC and wPS-AC as adsorbents from aqueous solution was studied at varying the pH (Figure 7(c) and 7(d)). The initial pH values of PAH mixture solutions were adjusted at 3, 5, 7, 9, and 11 with drops of diluted 0.1 M NaOH or HCl solution. The degree of adsorption limit is influenced by the pH of the solution due to the adsorbent surface charge distribution fluctuating the degree of adsorption, as indicated by the functional group of adsorbate (Putra et al. 2009). For example, Ph (89.44%), Fl (91.07%), BkF (88.67%), D[ah]A (92.47%), B[ghi]P (86.82%) adsorption over wPET-AC was recorded at pH 3.0 which is because the increment of positive charge on the surface of adsorbent at low pH resulted in a higher interaction between the PAH molecule and adsorbent surface. In agreement with this, an improved removal efficiency of Ph over soil comprising humic and fulvic acids, occurred at pH 3.0 (Ping et al. 2006). Similarly, the maximum adsorption of Nap (87.73%), Ace (84.00%), Acy (91%), Ant (91.63%), Flu (87.52%), Chr (89.49%) and InP (92.88%) occurs over the wPET-AC at pH 5.0. Results have shown that, under the optimum conditions, the maximum adsorption of BaA, Pyr, BaP, and BbF from aqueous medium attained over wPET-AC adsorbent was 90%, 87.27%, 86 and 88.23%, respectively at pH 7.0.

In the case of wPS-AC, the adsorption was Ph (55.33%), Fl (84.66%), BkF (86.23%), D[ah]A (85.09), B[ghi]P (86.54%) at pH 3.0, Nap (83.61%), Ace (79.40%), Acy (87.50%), Ant (89.64%), Flu (87.02%), Chr (82.30) and InP (99.25%) at pH 5.0 and BaA (84.70%), Pyr (84.10%), BaP (85.07%) and BbF (82.15%) at pH 7.0. At low pH values, due to the formation of positive charges on the surface of the adsorbent, higher adsorption occurs, and then the electrostatic interaction between the PAH and the surface charge of the adsorbent increases. Conversely, the positive charge on the surface of the adsorbent at higher pH is diminished and PAH molecules interact with OH ions for the adsorption on active sites, diminishing the adsorption efficiency of the adsorbent (Gupta 2015). As shown in Figure 7(c) and 7(d), the adsorptive efficiency of wPET-AC and wPS-AC decreased when the pH of the solution was increased from 7.0 to 11.0. This may be due to the increased number of surface functional groups. pH can change the chemical form, hydrophilicity, solubility, and protonation of PAHs in solution (Liao et al. 2008). These results are in agreement with the findings of Ping et al. (2006) and Balati et al. (2015). Furthermore, the zeta potential of the wPET-AC and wPS-AC decreased as the pH value was increased (as shown in Figure 8). The adsorption quantity decreased with increasing pH value due to the electrostatic repulsion force (Deng et al. 2013).

Figure 8

Zeta potentials of wPET-AC and wPS-AC at various pH values.

Figure 8

Zeta potentials of wPET-AC and wPS-AC at various pH values.

Close modal

Adsorption kinetics

In order to investigate the mechanism of sorption and potential rate-controlling steps such as mass transport and chemical reaction processes, kinetic models have been used to test the experimental data. These kinetic models included the pseudo-first-order equation and the pseudo-second-order equation. Based on these results, the pseudo-second-order kinetics equation provided a better fit to the adsorption data compared to the pseudo-first-order equation. The reaction fits well to a pseudo-second-order model as it indicates a trend towards chemisorption. In this kind of adsorption, the chemical reaction seems significant in the rate-controlling step and the pseudo-second-order chemical reaction kinetics provides the best correlation of the experimental data and the adsorption's mechanism is chemically rate controlling and because of this it is called chemisorption. According to Robati (2013), the removal from a solution is due to physico-chemical interactions between the two phases. The values of experimental adsorption capacity (qe,exp) were found to be in good agreement with the calculated qe values (qe,cal) using the pseudo-second-order model. The values of adsorption capacity, and correlation coefficient (R2) along with rate constants for pseudo-second-order kinetics models are given in Table 3. The adsorption data were also analyzed using an intra-particle diffusion kinetic model. The data from the Table (Table 4) about the regression factor and reaction rate are very accurate accordingly; however, the data manipulation via the pseudo-second-order equation is more reliable and thus shows the best fitting of data via the pseudo-second-order reaction rate.

Table 3

Pseudo-second-order kinetic parameters for adsorption of PAHs over wPET-AC and wPS-AC

wPET-AC
wPS-AC
PAHsk2 (mg g−1 min−1)qe2 (mg g−1)R2k2 (mg g−1 min−1)qe2 (mg g−1)R2
Nap 5 × 103 12.987 0.973 1 × 104 10.870 0.999 
Ace 5 × 103 13.158 0.999 7 × 104 10.870 0.999 
Acy 5 × 103 13.699 0.965 2 × 104 11.364 1.000 
Ant 9 × 103 13.889 0.997 1 × 104 12.195 0.999 
Ph 7 × 103 12.821 0.996 9 × 103 11.628 0.998 
Fl 9 × 103 13.699 0.999 2 × 104 10.989 0.999 
Flu 8 × 103 12.048 0.998 1 × 104 11.765 0.999 
BaA 7 × 103 12.821 0.998 1 × 104 11.111 1.000 
Chr 7 × 103 12.346 0.998 7 × 103 11.628 0.997 
Pyr 1 × 104 11.765 0.999 4 × 103 12.346 0.979 
BaP 7 × 103 12.821 0.998 5 × 103 12.500 0.999 
BkF 8 × 103 12.346 0.998 5 × 103 12.346 0.978 
BbF 1 × 103 13.333 0.993 1 × 104 10.869 0.999 
D[ah]A 6 × 103 15.152  9 × 103 11.628 0.999 
B[ghi]P 1 × 104 11.905 0.999 4 × 103 10.309 0.995 
InP 8 × 103 13.333 0.999 6 × 103 13.158 0.999 
wPET-AC
wPS-AC
PAHsk2 (mg g−1 min−1)qe2 (mg g−1)R2k2 (mg g−1 min−1)qe2 (mg g−1)R2
Nap 5 × 103 12.987 0.973 1 × 104 10.870 0.999 
Ace 5 × 103 13.158 0.999 7 × 104 10.870 0.999 
Acy 5 × 103 13.699 0.965 2 × 104 11.364 1.000 
Ant 9 × 103 13.889 0.997 1 × 104 12.195 0.999 
Ph 7 × 103 12.821 0.996 9 × 103 11.628 0.998 
Fl 9 × 103 13.699 0.999 2 × 104 10.989 0.999 
Flu 8 × 103 12.048 0.998 1 × 104 11.765 0.999 
BaA 7 × 103 12.821 0.998 1 × 104 11.111 1.000 
Chr 7 × 103 12.346 0.998 7 × 103 11.628 0.997 
Pyr 1 × 104 11.765 0.999 4 × 103 12.346 0.979 
BaP 7 × 103 12.821 0.998 5 × 103 12.500 0.999 
BkF 8 × 103 12.346 0.998 5 × 103 12.346 0.978 
BbF 1 × 103 13.333 0.993 1 × 104 10.869 0.999 
D[ah]A 6 × 103 15.152  9 × 103 11.628 0.999 
B[ghi]P 1 × 104 11.905 0.999 4 × 103 10.309 0.995 
InP 8 × 103 13.333 0.999 6 × 103 13.158 0.999 
Table 4

Intra-particle diffusion parameters for adsorption of PAHs on wPET-AC and wPS-AC

wPET-AC
wPS-AC
PAHsC (mg g−1)Kid (min−1)R2C (mg g−1)Kid (min−1)R2
Nap 0.038 0.012 0.992 0.260 0.093 0.992 
Ace 0.365 0.124 0.897 0.334 0.139 0.969 
Acy 0.349 0.152 0.884 0.343 0.148 0.876 
Ant 0.169 0.108 0.828 0.23 0.188 0.872 
Ph 0.152 0.085 0.960 0.178 0.053 0.839 
Fl 0.202 0.052 0.976 0.199 0.055 0.947 
Flu 0.195 0.076 0.868 0.189 0.084 0.833 
BaA 0.042 0.016 0.974 0.038 0.204 0.879 
Chr 0.040 0.015 0.930 0.042 0.013 0.949 
Pyr 0.055 0.006 0.959 0.054 0.006 0.984 
BaP 0.027 0.011 0.974 0.028 0.009 0.973 
BkF 0.039 0.010 0.998 0.037 0.012 0.982 
BbF 0.055 0.008 0.999 0.052 0.012 0.873 
D[ah]A 0.039 0.024 0.798 0.052 0.012 0.873 
B[ghi]P 0.030 0.015 0.979 0.012 0.030 0.994 
InP 0.028 0.015 0.895 0.031 0.020 0.952 
wPET-AC
wPS-AC
PAHsC (mg g−1)Kid (min−1)R2C (mg g−1)Kid (min−1)R2
Nap 0.038 0.012 0.992 0.260 0.093 0.992 
Ace 0.365 0.124 0.897 0.334 0.139 0.969 
Acy 0.349 0.152 0.884 0.343 0.148 0.876 
Ant 0.169 0.108 0.828 0.23 0.188 0.872 
Ph 0.152 0.085 0.960 0.178 0.053 0.839 
Fl 0.202 0.052 0.976 0.199 0.055 0.947 
Flu 0.195 0.076 0.868 0.189 0.084 0.833 
BaA 0.042 0.016 0.974 0.038 0.204 0.879 
Chr 0.040 0.015 0.930 0.042 0.013 0.949 
Pyr 0.055 0.006 0.959 0.054 0.006 0.984 
BaP 0.027 0.011 0.974 0.028 0.009 0.973 
BkF 0.039 0.010 0.998 0.037 0.012 0.982 
BbF 0.055 0.008 0.999 0.052 0.012 0.873 
D[ah]A 0.039 0.024 0.798 0.052 0.012 0.873 
B[ghi]P 0.030 0.015 0.979 0.012 0.030 0.994 
InP 0.028 0.015 0.895 0.031 0.020 0.952 

Isotherm study

Adsorption mechanisms pathways, adsorbents' surface properties and capacities of adsorbents which are basic parameters to design an effective adsorption system, could be determined with the help of adsorption isotherms. Adsorption isotherms explain how the contaminants, for example, PAHs interact with the adsorbent materials. Freundlich and Langmuir isotherm models have been used to understand and determine the PAH adsorption mechanisms and adsorption capacity of the adsorbents. The equilibrium adsorption data were evaluated using the Freundlich and Langmuir isotherm models. The calculated Langmuir constants (K1 and qm), as well as the coefficients of correlation (R2) for isotherms, are detailed in Table 5 for wPET-AC and wPS-AC. The results reveal that the adsorption data fitted better with Langmuir adsorption isotherms in the case of both wPET-AC and wPS-AC.

Table 5

Isotherm model parameters for PAHs adsorption over wPET-AC and wPS-AC

Langmuir isotherm parameters
Langmuir isotherm parameters
wPET-AC
wPS-AC
PAHsqm (mg/g)K1R2qm (mg/g)K1R2
Nap 5.376 1.431 0.927 4.739 1.381 0.955 
Ace 4.405 1.517 0.918 3.817 1.295 0.963 
Acy 6.289 1.360 0.959 5.525 1.372 0.977 
Ant 7.353 1.210 0.952 6.289 1.283 0.987 
Ph 6.329 1.315 0.938 5.464 1.340 0.977 
Fl 7.576 1.204 0.891 5.747 1.323 0.988 
Flu 5.714 1.330 0.974 5.682 1.327 0.976 
BaA 6.061 1.371 0.922 5.051 1.374 0.977 
Chr 6.410 1.297 0.944 5.435 1.329 0.985 
Pyr 6.622 1.241 0.966 5.988 1.284 0.991 
BaP 6.494 1.298 0.959 5.714 1.331 0.987 
BkF 6.410 1.278 0.974 5.988 1.297 0.987 
BbF 5.556 1.471 0.910 4.505 1.426 0.993 
D[ah]A 4.950 1.606 0.887 4.049 1.442 0.958 
B[ghi]P 6.369 1.262 0.984 6.250 1.266 0.988 
InP 6.711 1.276 0.957 6.024 1.316 0.982 
Langmuir isotherm parameters
Langmuir isotherm parameters
wPET-AC
wPS-AC
PAHsqm (mg/g)K1R2qm (mg/g)K1R2
Nap 5.376 1.431 0.927 4.739 1.381 0.955 
Ace 4.405 1.517 0.918 3.817 1.295 0.963 
Acy 6.289 1.360 0.959 5.525 1.372 0.977 
Ant 7.353 1.210 0.952 6.289 1.283 0.987 
Ph 6.329 1.315 0.938 5.464 1.340 0.977 
Fl 7.576 1.204 0.891 5.747 1.323 0.988 
Flu 5.714 1.330 0.974 5.682 1.327 0.976 
BaA 6.061 1.371 0.922 5.051 1.374 0.977 
Chr 6.410 1.297 0.944 5.435 1.329 0.985 
Pyr 6.622 1.241 0.966 5.988 1.284 0.991 
BaP 6.494 1.298 0.959 5.714 1.331 0.987 
BkF 6.410 1.278 0.974 5.988 1.297 0.987 
BbF 5.556 1.471 0.910 4.505 1.426 0.993 
D[ah]A 4.950 1.606 0.887 4.049 1.442 0.958 
B[ghi]P 6.369 1.262 0.984 6.250 1.266 0.988 
InP 6.711 1.276 0.957 6.024 1.316 0.982 

Adsorption thermodynamics

The investigations of thermodynamic parameters for adsorption studies are another essential factor. Thermodynamic parameters are used to decide that a process is physical or chemical, non-spontaneous or spontaneous, and exothermic or endothermic in nature. The thermodynamic parameters including ΔG°, ΔH° and ΔS° were calculated in batch mode experiments at various temperatures.

For the thermodynamic parameters calculation, Equations (1)–(3) were applied (Konicki et al. 2012):
(1)
(2)
(3)

Results for thermodynamic parameters are detailed in Tables 6 and 7 for wPET-AC and wPS-AC, respectively. The results revealed that an increase in temperature enhances the adsorption of PAHs onto wPET-AC and wPS-AC. The adsorption phenomenon boosts up at higher temperatures due to the creation of more adsorption sites and intraparticle diffusion. The adsorption process is non-spontaneous and endothermic in nature, which was confirmed by the positive values of ΔG°, ΔS° and ΔH°.

Table 6

Thermodynamic factors for the adsorption of PAHs onto wPET-AC

wPET-AC
Nap
Ace
Temperature (°C)ΔG° (kJ mol−1)ΔH° (kJ mol−1)ΔS° (kJ mol−1)ΔG° (kJ mol−1)ΔH° (kJ mol−1)ΔS° (kJ mol−1)
20 7.4333 16.9107 0.0313 7.7732 20.918 0.0437 
30 7.0332 7.1001 
40 6.9831 6.7222 
50 6.3993 6.4590 
 Acy Ant 
20 7.1539 15.9961 0.0287 6.9060 14.7324 0.0258 
30 7.1915 6.7371 
40 6.8631 6.3359 
50 6.2807 6.1808 
 Ph Fl 
20 7.0825 16.3666 0.0242 6.9640 14.6410 0.0256 
30 6.7241 6.5037 
40 6.5866 6.3292 
50 6.3268 6.1813 
 Flu BaA 
20 7.0276 12.8202 0.0193 7.2546 14.7158 0.0248 
30 6.6506 6.8386 
40 6.5687 6.6625 
50 6.4195 6.4954 
 Chr Pyr 
20 7.0223 13.3523 0.0293 6.8024 10.3925 0.0122 
30 6.7460 6.4536 
40 6.5344 6.4345 
50 6.3985 6.4175 
 BaP BkF 
20 7.0168 14.5079 0.0248 6.9111 12.6788 0.190 
30 6.8078 6.7665 
40 6.4006 6.4661 
50 6.3326 6.3790 
 BbF D[ah]A 
20 7.3831 17.6264 0.0336 7.7726 23.1129 0.0511 
30 6.4196 6.8912 
40 6.4894 6.5632 
50 6.2715 6.1993 
 B[ghi]P InP 
20 6.8204 11.1740 0.0145 6.9904 14.9486 0.0264 
30 6.5898 6.6549 
40 6.3803 6.3127 
50 6.4167 6.2341 
wPET-AC
Nap
Ace
Temperature (°C)ΔG° (kJ mol−1)ΔH° (kJ mol−1)ΔS° (kJ mol−1)ΔG° (kJ mol−1)ΔH° (kJ mol−1)ΔS° (kJ mol−1)
20 7.4333 16.9107 0.0313 7.7732 20.918 0.0437 
30 7.0332 7.1001 
40 6.9831 6.7222 
50 6.3993 6.4590 
 Acy Ant 
20 7.1539 15.9961 0.0287 6.9060 14.7324 0.0258 
30 7.1915 6.7371 
40 6.8631 6.3359 
50 6.2807 6.1808 
 Ph Fl 
20 7.0825 16.3666 0.0242 6.9640 14.6410 0.0256 
30 6.7241 6.5037 
40 6.5866 6.3292 
50 6.3268 6.1813 
 Flu BaA 
20 7.0276 12.8202 0.0193 7.2546 14.7158 0.0248 
30 6.6506 6.8386 
40 6.5687 6.6625 
50 6.4195 6.4954 
 Chr Pyr 
20 7.0223 13.3523 0.0293 6.8024 10.3925 0.0122 
30 6.7460 6.4536 
40 6.5344 6.4345 
50 6.3985 6.4175 
 BaP BkF 
20 7.0168 14.5079 0.0248 6.9111 12.6788 0.190 
30 6.8078 6.7665 
40 6.4006 6.4661 
50 6.3326 6.3790 
 BbF D[ah]A 
20 7.3831 17.6264 0.0336 7.7726 23.1129 0.0511 
30 6.4196 6.8912 
40 6.4894 6.5632 
50 6.2715 6.1993 
 B[ghi]P InP 
20 6.8204 11.1740 0.0145 6.9904 14.9486 0.0264 
30 6.5898 6.6549 
40 6.3803 6.3127 
50 6.4167 6.2341 
Table 7

Thermodynamic factors for the adsorption of PAHs onto wPS-AC

wPS-AC
Nap
Ace
Temperature (oC)ΔG° (kJ mol−1)ΔH° (kJ mol−1)ΔS° (kJ mol−1)ΔG° (kJ mol−1)ΔH° (kJ mol−1)ΔS° (kJ mol−1)
20 7.4333 15.7716 0.0276 7.7732 19.1471 0.0380 
30 7.0332 7.1001 
40 6.9831 6.7222 
50 6.5279 6.6577 
 Acy Ant 
20 7.1539 14.8071 0.0248 6.9060 13.1943 0.0208 
30 7.1915 6.7371 
40 6.8631 6.3359 
50 6.4141 6.3539 
 Ph Fl 
20 7.0825 13.0280 0.0198 6.9640 11.3808 0.0166 
30 6.7241 6.5037 
40 6.5866 6.3292 
50 6.4769 6.4966 
 Flu BaA 
20 7.0276 12.7454 0.0191 7.2546 14.7158 0.0248 
30 6.6506 6.8386 
40 6.5687 6.6625 
50 6.4278 6.4954 
 Chr Pyr 
20 7.0223 11.8558 0.0162 6.8024 9.5362 0.0094 
30 6.7460 6.4536 
40 6.5344 6.4345 
50 6.5675 6.5132 
 BaP BkF 
20 7.0168 13.1610 0.0204 6.9111 12.0389 0.170 
30 6.8078 6.7665 
40 6.4006 6.4661 
50 6.4845 6.4506 
 BbF D[ah]A 
20 7.3831 14.6493 0.0250 7.7726 20.5855 0.0430 
30 6.4196 6.8912 
40 6.4894 6.5632 
50 6.5720 6.4839 
 B[ghi]P InP 
20 6.8204 10.9495 0.0138 6.9904 13.7846 0.0227 
30 6.5898 6.6549 
40 6.3803 6.3127 
50 6.4417 6.3648 
wPS-AC
Nap
Ace
Temperature (oC)ΔG° (kJ mol−1)ΔH° (kJ mol−1)ΔS° (kJ mol−1)ΔG° (kJ mol−1)ΔH° (kJ mol−1)ΔS° (kJ mol−1)
20 7.4333 15.7716 0.0276 7.7732 19.1471 0.0380 
30 7.0332 7.1001 
40 6.9831 6.7222 
50 6.5279 6.6577 
 Acy Ant 
20 7.1539 14.8071 0.0248 6.9060 13.1943 0.0208 
30 7.1915 6.7371 
40 6.8631 6.3359 
50 6.4141 6.3539 
 Ph Fl 
20 7.0825 13.0280 0.0198 6.9640 11.3808 0.0166 
30 6.7241 6.5037 
40 6.5866 6.3292 
50 6.4769 6.4966 
 Flu BaA 
20 7.0276 12.7454 0.0191 7.2546 14.7158 0.0248 
30 6.6506 6.8386 
40 6.5687 6.6625 
50 6.4278 6.4954 
 Chr Pyr 
20 7.0223 11.8558 0.0162 6.8024 9.5362 0.0094 
30 6.7460 6.4536 
40 6.5344 6.4345 
50 6.5675 6.5132 
 BaP BkF 
20 7.0168 13.1610 0.0204 6.9111 12.0389 0.170 
30 6.8078 6.7665 
40 6.4006 6.4661 
50 6.4845 6.4506 
 BbF D[ah]A 
20 7.3831 14.6493 0.0250 7.7726 20.5855 0.0430 
30 6.4196 6.8912 
40 6.4894 6.5632 
50 6.5720 6.4839 
 B[ghi]P InP 
20 6.8204 10.9495 0.0138 6.9904 13.7846 0.0227 
30 6.5898 6.6549 
40 6.3803 6.3127 
50 6.4417 6.3648 

Mechanisms of PAHs adsorption

Generally, AC adsorbs pollutants from aqueous solutions, depending on the solution conditions, the adsorbent surface characteristics, and the nature of the adsorbate. Various mechanisms by which the carbon surface functional groups may affect the adsorption process from aqueous solution have been proposed: electron-donor-electron-acceptor interactions (Mattson et al. 1969), hydrogen bonding (Coughlin & Ezra 1968; Mahajan et al. 1980; Yuan et al. 2010), and π–π dispersive interactions (Coughlin & Ezra 1968). Lyklema (1995) reported that the PAHs adsorption from aqueous solution is basically an exchange process. Therefore, the adsorption process of organic compounds on the surface of AC has been described as a complex interaction between electrostatic and dispersion interactions, especially in the case of weak organic electrolytes. However, since PAHs are not considered to be electrolytic, the intermediate hydrogen bonding, as well as van der Waals dispersive interactions, have become the main mechanisms proposed in the literature (Yuan et al. 2010). The mechanism of adsorption of PAH by AC is given in the reaction scheme in Figure 9.

Figure 9

Reaction scheme: preparation and mechanism of adsorption of AC.

Figure 9

Reaction scheme: preparation and mechanism of adsorption of AC.

Close modal

Comparison of adsorption potential of different adsorbents

The comparative adsorption efficiency of the various adsorption media reported in the literature and the AC, i.e., wPET-AC and wPS-AC for PAHs adsorption are detailed in Table 8. The efficiency of wPET-AC was probably because of the larger surface area (65.35 m2 g−1 for wPET-AC compared to 50.14 m2 g−1 for wPS-AC) as well as hydrophobicity which advances higher adsorption capacity than the wPS-AC utilized in our research. The decision for adsorbent for water treatment applications relies upon many factors, for example, reusability, flexibility, availability of material, nontoxicity, efficiency, etc. wPET-AC shows high potential for water treatment applications with a removal efficiency >85% for all PAHs at the same time.

Table 8

Details of adsorption media and their PAHs removal efficiency

MaterialsPAHs removal efficiency (%)Reference
Combined permanganate (Mn(VII))/bisulfite advanced oxidation process Pyr, Chr, BaA, BbF, BkF, BaP, and D[ah]A (85–100%) Fang et al. (2021)  
Clinoptilolite NaP (100) Puszkarewicz & Kaleta (2020)  
Graphene wool Ph (99.9), Pyr (99.1) Adeola & Forbes (2019)  
Stearoyl grafted cellulose Ph (97.61), Pyr (96.94), Kim et al. (2018)  
Grafted cellulose 16 PAHs (94 and 98) Euvrard et al. (2014)  
Activated carbon bed NaP (98.1), Ace (98.2), Acy (98.4), Ant (98.9), Ph (100), Fl (94.7) Smol et al. (2014)  
Activated carbon coconut shells and anthracite based Ph (95), Ant (97.5), Flu (96.5), Pyr (98), BaP (88), Cry (94.5), BbF (86), BkF (88), BaA (94.5), InP (73), B[ghi]P (69) Amstaetter et al. (2012)  
Light weight expanded clay aggregate Ph (92.61), Flu (93.91), Pyr (94.15) Nkansah et al. (2012)  
Activated carbon Nap (97), Fl (98) Awoyemi (2011)  
Coke-derived porous carbon NaP, Fl, Ph, Pyr, Flu (>99 for all PAHs) Yuan et al. (2010)  
Wood ashes Fl, Pyr, Cry, BaA, BbF, BkF, BaP, D[ah]A, InP, B[ghi]P (>99) Pérez-Gregorio et al. (2010)  
Inorgano–organo-bentonite Ph (99) Ma & Zhu (2006)  
Wood char Pyr, Ph, BaA (≥60) Wang et al. (2006)  
wPS-AC Nap (83.61), Ace (79.4), Acy (87.5), Ant (89.64), Ph (85.33), Fl (84.66), Flu (87.02), BaA (84.7), Chr (82.3), Pyr (84.1), BaP (85.07), BkF (86.23), BbF (82.15), D[ah]A (85.09),B [ghi]P (86.54), InP (89.25) This article 
wPET-AC Nap (88.16), Ace (85.9), Acy (92.74), Ant (96.22), Ph (90.48), Fl (96.31), Flu (87.02), BaA (91.10), Chr (88.38), Pyr (87.46), BaP (90.07), BkF (88.23), BbF (92.15), D[ah]A (95.09), B[ghi]P (87.54), InP (94.25) This article 
MaterialsPAHs removal efficiency (%)Reference
Combined permanganate (Mn(VII))/bisulfite advanced oxidation process Pyr, Chr, BaA, BbF, BkF, BaP, and D[ah]A (85–100%) Fang et al. (2021)  
Clinoptilolite NaP (100) Puszkarewicz & Kaleta (2020)  
Graphene wool Ph (99.9), Pyr (99.1) Adeola & Forbes (2019)  
Stearoyl grafted cellulose Ph (97.61), Pyr (96.94), Kim et al. (2018)  
Grafted cellulose 16 PAHs (94 and 98) Euvrard et al. (2014)  
Activated carbon bed NaP (98.1), Ace (98.2), Acy (98.4), Ant (98.9), Ph (100), Fl (94.7) Smol et al. (2014)  
Activated carbon coconut shells and anthracite based Ph (95), Ant (97.5), Flu (96.5), Pyr (98), BaP (88), Cry (94.5), BbF (86), BkF (88), BaA (94.5), InP (73), B[ghi]P (69) Amstaetter et al. (2012)  
Light weight expanded clay aggregate Ph (92.61), Flu (93.91), Pyr (94.15) Nkansah et al. (2012)  
Activated carbon Nap (97), Fl (98) Awoyemi (2011)  
Coke-derived porous carbon NaP, Fl, Ph, Pyr, Flu (>99 for all PAHs) Yuan et al. (2010)  
Wood ashes Fl, Pyr, Cry, BaA, BbF, BkF, BaP, D[ah]A, InP, B[ghi]P (>99) Pérez-Gregorio et al. (2010)  
Inorgano–organo-bentonite Ph (99) Ma & Zhu (2006)  
Wood char Pyr, Ph, BaA (≥60) Wang et al. (2006)  
wPS-AC Nap (83.61), Ace (79.4), Acy (87.5), Ant (89.64), Ph (85.33), Fl (84.66), Flu (87.02), BaA (84.7), Chr (82.3), Pyr (84.1), BaP (85.07), BkF (86.23), BbF (82.15), D[ah]A (85.09),B [ghi]P (86.54), InP (89.25) This article 
wPET-AC Nap (88.16), Ace (85.9), Acy (92.74), Ant (96.22), Ph (90.48), Fl (96.31), Flu (87.02), BaA (91.10), Chr (88.38), Pyr (87.46), BaP (90.07), BkF (88.23), BbF (92.15), D[ah]A (95.09), B[ghi]P (87.54), InP (94.25) This article 

Desorption and reusability

In the adsorption processes, desorption/regeneration of adsorbents is one of the essential aspects, as it controls the economy of water treatment technology (Ali 2012). For the recovery and effective regeneration of adsorbents, acids (such as H2SO4, HCOOH, HCl, CH3COOH, and HNO3), buffer solutions (such as bicarbonate and phosphate), salts (such as NH4NO3, CaCl2.2H2O, KNO3, (NH4)2SO4, C6H5Na3O7.2H2O, KNO3, NaCl and KCl), deionized water as well as chelating agents have been used in various studies (Lata et al. 2015). In this study, HCl solution was used for the desorption experiment. This acid gave satisfactory desorption results. The regeneration studies were carried out in batches using 100 mL of standard solutions indicating that wPET-AC and wPS-AC had the potential for reusability (Figure 10).

Figure 10

Regenerating efficiency of wPET-AC and wPS-AC.

Figure 10

Regenerating efficiency of wPET-AC and wPS-AC.

Close modal

In the present work, the efficiency of wPET-AC and wPS-AC for adsorption of 16 PAHs from aqueous medium was examined. The effects of operating variables such as pH, contact time, temperature, adsorbent dose, and initial PAHs concentration were investigated. An increase in adsorbent dose, initial PAHs concentration, and contact time improved the adsorption efficiency. The adsorption process after a contact time of 2 hr reached equilibrium. Kinetics studies showed that the adsorption system followed a pseudo-second-order equation. The study revealed that the Langmuir equation represented the best fit for the experimental data. Thermodynamic results indicated that the adsorption of PAHs onto wPET-AC and wPS-AC is endothermic in nature. These results indicated that wPET-AC can be used as a suitable and efficient adsorbent for the removal of PAHs from wastewater. This study will benefit the entire community through minimizing the environmental pollution caused by solid plastic waste and as well as removing the toxic PAHs from wastewater. Thus, the study serves to develop a route for better reuse of waste plastic materials, thereby minimizing the environmental degradation due to waste plastics.

We are very thankful to the staff of the Centralized Resource Laboratories and Materials Research Laboratory, University of Peshawar for providing instrument facilities.

All relevant data are included in the paper or its Supplementary Information.

Abbas
S. T.
Mustafa
M.
Faize
A.
Rah
A. Z.
2013
Adsorption of Pb+2 and Zn+2 ions from oil wells onto activated carbon produced from Rice Husk in batch adsorption process
.
Journal of Chemical and Pharmaceutical Research
5
,
240
250
.
Abdel-Shafy
H. I.
Mansour
M. S.
2016
A review on polycyclic aromatic hydrocarbons: source, environmental impact, effect on human health and remediation
.
Egyptian Journal of Petroleum
25
(
1
),
107
123
.
Adeola
A. O.
Forbes
P. B.
2019
Optimization of the sorption of selected polycyclic aromatic hydrocarbons by regenerable grapheme wool
.
Water Science and Technology
80
(
10
),
1931
1943
.
Ali
I.
2012
New generation adsorbents for water treatment
.
Chemical Reviews
112
(
10
),
5073
5091
.
Álvarez-Torrellas
S.
Ribeiro
R. S.
Gomes
H. T.
Ovejero
G.
García
J.
2016
Removal of antibiotic compounds by adsorption using glycerol-based carbon materials
.
Chemical Engineering Journal
296
,
277
288
.
Amstaetter
K.
Eek
E.
Cornelissen
G.
2012
Sorption of PAHs and PCBs to activated carbon: coal versus biomass-based quality
.
Chemosphere
87
(
5
),
573
578
.
Ania
C. O.
Cabal
B.
Pevida
C.
Arenillas
A.
Parra
J. B.
Rubiera
F.
Pis
J. J.
2007
Effects of activated carbon properties on the adsorption of naphthalene from aqueous solutions
.
Applied Surface Science
253
(
13
),
5741
5746
.
Arora
C.
Sahua
D.
Bhartib
D.
Tamrakara
V.
Sonia
S.
Sharma
S.
2019a
Adsorption of hazardous dye crystal violet from industrial waste using low-cost adsorbent Chenopodium album
.
Desalination and Water Treatment
167
,
324
332
.
Arora
C.
Soni
S.
Sahu
S.
Mittal
J.
Kumar
P.
Bajpai
P. K.
2019b
Iron based metal organic framework for efficient removal of methylene blue dye from industrial waste
.
Journal of Molecular Liquids
284
,
343
352
.
Arora
C.
Kumar
P.
Soni
S.
Mittal
J.
Mittal
A.
Singh
B.
2020
Efficient removal of malachite green dye from aqueous solution using Curcuma caesia based activated carbon
.
Desalination and Water Treatment
195
,
341
352
.
Asha
P.
Sanju
S.
Jyoti
M.
Alok
M.
Charu
A.
2021
Sequestration of crystal violet from aqueous solution using ash of black turmeric rhizome
.
Desalination and Water Treatment
220
,
342
352
.
Awe
A. A.
Opeolu
B. O.
Fatoki
O. S.
Ayanda
O. S.
Jackson
V. A.
Snyman
R.
2020
Preparation and characterisation of activated carbon from Vitis vinifera leaf litter and its adsorption performance for aqueous phenanthrene
.
Applied Biological Chemistry
63
(
1
),
1
17
.
Awoyemi
A.
2011
Understanding the Adsorption of Polycyclic Aromatic Hydrocarbons From Aqueous Phase Onto Activated Carbon
. Doctoral Dissertation.
Balati
A.
Shahbazi
A.
Amini
M. M.
Hashemi
S. H.
2015
Adsorption of polycyclic aromatic hydrocarbons from wastewater by using silica-based organic–inorganic nanohybrid material
.
Journal of Water Reuse and Desalination
5
(
1
),
50
63
.
Bhatnagar
A.
Anastopoulos
I.
2017
Adsorptive removal of bisphenol A (BPA) from aqueous solution: a review
.
Chemosphere
168
,
885
902
.
Björklund
K.
Li
L.
2015
Evaluation of low-cost materials for sorption of hydrophobic organic pollutants in stormwater
.
Journal of Environmental Management
159
,
106
114
.
Blanchard
M.
Teil
M. J.
Ollivon
D.
Legenti
L.
Chevreuil
M.
2004
Polycyclic aromatic hydrocarbons and polychlorobiphenyls in wastewaters and sewage sludges from the Paris area (France)
.
Environmental Research
95
(
2
),
184
197
.
Busetti
F.
Heitz
A.
Cuomo
M.
Badoer
S.
Traverso
P.
2006
Determination of sixteen polycyclic aromatic hydrocarbons in aqueous and solid samples from an Italian wastewater treatment plant
.
Journal of Chromatography A
1102
(
1–2
),
104
115
.
Canzano
S.
Capasso
S.
Natale
M. D.
Erto
A.
Iovino
P.
Musmarra
D.
2014
Remediation of groundwater polluted by aromatic compounds by means of adsorption
.
Sustainability
6
(
8
),
4807
4822
.
Chang
C. F.
Chang
C. Y.
Chen
K. H.
Tsai
W. T.
Shie
J. L.
Chen
Y. H.
2004a
Adsorption of naphthalene on zeolite from aqueous solution
.
Journal of Colloid and Interface Science
277
(
1
),
29
34
.
Chang
C. F.
Chang
C. Y.
Höll
W.
Ulmer
M.
Chen
Y. H.
Groß
H. J.
2004b
Adsorption kinetics of polyethylene glycol from aqueous solution onto activated carbon
.
Water Research
38
(
10
),
2559
2570
.
Charlesworth
M.
Service
M.
Gibson
C. E.
2002
PAHs contamination of Irish sediments
.
Marine Pollution Bulletin
44
,
1421
1424
.
Comninellis
C.
Kapalka
A.
Malato
S.
Parsons
S. A.
Poulios
I.
Mantzavinos
D.
2008
Advanced oxidation processes for water treatment: advances and trends for R&D
.
Journal of Chemical Technology & Biotechnology: International Research in Process, Environmental & Clean Technology
83
(
6
),
769
776
.
Coughlin
R. W.
Ezra
F. S.
1968
Role of surface acidity in the adsorption of organic pollutants on the surface of carbon
.
Environmental Science & Technology
2
(
4
),
291
297
.
Crisafully
R.
Milhome
M. A. L.
Cavalcante
R. M.
Silveira
E. R.
De Keukeleire
D.
Nascimento
R. F.
2008
Removal of some polycyclic aromatic hydrocarbons from petrochemical wastewater using low-cost adsorbents of natural origin
.
Bioresource Technology
99
(
10
),
4515
4519
.
Dabek
L.
Ozimina
E.
2009
Oxidation of organic contaminants adsorbed on activated carbons
.
Ochrona Środowiska I Zasobów Naturalnych
41
,
427
436
.
Deng
J. H.
Zhang
X. R.
Zeng
G. M.
Gong
J. L.
Niu
Q. Y.
Liang
J.
2013
Simultaneous removal of Cd(II) and ionic dyes from aqueous solution using magnetic graphene oxide nanocomposite as an adsorbent
.
Chemical Engineering Journal
226
,
189
200
.
Ding
S. L.
Sun
Y. Z.
Yang
C. N.
Xu
B. H.
2009
Removal of copper from aqueous solutions by bentonites and the factors affecting it
.
Mining Science and Technology (China)
19
(
4
),
489
492
.
Eisler
R.
1987
Polycyclic Aromatic Hydrocarbon Hazards to Fish, Wildlife, and Invertebrates: A Synoptic Review (No. 11)
.
Fish and Wildlife Service, US Department of the Interior, Washington, DC
.
Euvrard
E.
Druart
C.
Poupeney
A.
Crini
N.
Vismara
E.
Lanza
T.
Torri
G.
Gavoille
S.
Crini
G.
2014
Grafted cellulose for PAHs removal present in industrial discharge waters
. In:
EGU General Assembly Conference Abstracts
, p.
1742
.
Fang
J.
Zhao
R.
Rao
B.
Rakowska
M.
Athanasiou
D.
Millerick
K.
Wei
S.
Lei
X.
Lou
H. H.
Reible
D. D.
2021
Removal of polycyclic aromatic hydrocarbons from water using Mn(III)-based advanced oxidation process
.
Journal of Environmental Engineering
147
(
3
),
04021002
.
Gong
Z.
Alef
K.
Wilke
B. M.
Li
P.
2007
Activated carbon adsorption of PAHs from vegetable oil used in soil remediation
.
Journal of Hazardous Materials
143
(
1–2
),
372
378
.
Gupta
H.
2015
Removal of phenanthrene from water using activated carbon developed from orange rind
.
International Journal of Scientific Research in Environmental Sciences
3
(
7
),
248
.
Gupta
H.
Gupta
B.
2016
Adsorption of polycyclic aromatic hydrocarbons on banana peel activated carbon
.
Desalination and Water Treatment
57
(
20
),
9498
9509
.
Hale
S. E.
Elmquist
M.
Brändli
R.
Hartnik
T.
Jakob
L.
Henriksen
T.
Werner
D.
Cornelissen
G.
2012
Activated carbon amendment to sequester PAHs in contaminated soil: a lysimeter field trial
.
Chemosphere
87
(
2
),
177
184
.
Hall
S.
Tang
R.
Baeyens
J.
Dewil
R.
2009
Removing polycyclic aromatic hydrocarbons from water by adsorption on silicagel
.
Polycyclic Aromatic Compounds
29
(
3
),
160
183
.
Herman
V.
Takacs
H.
Duclairoir
F.
Renault
O.
Tortai
J. H.
Viala
B.
2015
Core double–shell cobalt/graphene/polystyrene magnetic nanocomposites synthesized by in situ sonochemical polymerization
.
RSC Advances
5
(
63
),
51371
51381
.
Hidayat
D.
Supriyanto
R.
Permana
D. F.
2019
Adsorption of polycyclic aromatic hydrocarbons using low-cost activated carbon derived from rice husk
. In:
Journal of Physics: Conference Series
.
IOP Publishing
, Vol.
1338
(
1
), p.
012005
.
Ho
Y. S.
McKay
G.
1998
Sorption of dye from aqueous solution by peat
.
Chemical Engineering Journal
70
(
2
),
115
124
.
Hu
Y.
He
Y.
Wang
X.
Wei
C.
2014
Efficient adsorption of phenanthrene by simply synthesized hydrophobic MCM-41 molecular sieves
.
Applied Surface Science
311
,
825
830
.
Ikram
R.
Jan
B. M.
Ahmad
W.
2020
Advances in synthesis of graphene derivatives using industrial wastes precursors; prospects and challenges
.
Journal of Materials Research and Technology
9
(
6
),
15924
15951
.
Ilyas
M.
Ahmad
W.
Khan
H.
Yousaf
S.
Khan
K.
Nazir
S.
2018
Plastic waste as a significant threat to environment–a systematic literature review
.
Reviews on Environmental Health
33
(
4
),
383
406
.
Ilyas
M.
Ahmad
W.
Khan
H.
Ahmad
I.
2019
Application of composite adsorbents prepared from waste PS and PET for removal of Cr and Cu ions from wastewater
.
Desalination and Water Treatment
171
,
144
157
.
Ilyas
M.
Ahmad
W.
Khan
H.
Yousaf
S.
2021
Potentially poisonous elements removal from vehicle-wash wastewater and aqueous solutions using composite adsorbents
.
Desalination and Water Treatment
224
,
331
342
.
Lata
S.
Singh
P. K.
Samadder
S. R.
2015
Regeneration of adsorbents and recovery of heavy metals: a review
.
International Journal of Environmental Science and Technology
12
(
4
),
1461
1478
.
Lawrinenko
M.
2014
Anion Exchange Capacity of Biochar
.
Graduate Theses
,
Iowa State University
.
Li
Y.
Chen
B.
Zhu
L.
2010
Enhanced sorption of polycyclic aromatic hydrocarbons from aqueous solution by modified pine bark
.
Bioresource Technology
101
(
19
),
7307
7313
.
Li
C. H.
Wong
Y. S.
Wang
H. Y.
Tam
N. F. Y.
2015
Anaerobic biodegradation of PAHs in mangrove sediment with amendment of NaHCO3
.
Journal of Environmental Sciences
30
,
148
156
.
Liao
Q.
Sun
J.
Gao
L.
2008
The adsorption of resorcinol from water using multi-walled carbon nanotubes
.
Colloids and Surfaces A: Physicochemical and Engineering Aspects
312
(
2–3
),
160
165
.
Liu
H.
Shuang
L.
Jinhong
G.
Huu
H. N.
Wenshan
G.
Zizhang
G.
Jing
W.
Yiran
L.
2014
Enhancement of Cr(VI) removal by modifying activated carbon developed from Zizania caduciflora with tartaric acid during phosphoric acid activation
.
Chemical Engineering Journal
246
,
168
174
.
Lyklema
J.
1995
Fundamentals of Interface and Colloid Science, Vol. II. Solid–Liquid Interfaces
.
Academic Press
,
New York
.
Ma
J.
Zhu
L.
2006
Simultaneous sorption of phosphate and phenanthrene to inorgano–organo-bentonite from water
.
Journal of Hazardous Materials
136
(
3
),
982
988
.
Mahajan
O. P.
Moreno-Castilla
C.
Walker
P. L.
Jr.
1980
Surface-treated activated carbon for removal of phenol from water
.
Separation Science and Technology
15
(
10
),
1733
1752
.
Manoli
E.
Samara
C.
1999
Occurrence and mass balance of polycyclic aromatic hydrocarbons in the Thessaloniki sewage treatment plant
.
Journal of Environmental Quality
28
(
1
),
176
187
.
Mansouri
E.
Yousefi
V.
Ebrahimi
V.
Eyvazi
S.
Hejazi
M. S.
Mahdavi
M.
Asghar
M.
Tarhriz
V.
2020
Overview of ultraviolet-based methods used in polycyclic aromatic hydrocarbons analysis and measurement
.
Separation Science Plus
3
(
4
),
112
120
.
Mattson
J. A.
Mark
H. B.
Jr
Malbin
M. D.
Weber
W. J.
Jr.
Crittenden
J. C.
1969
Surface chemistry of active carbon: specific adsorption of phenols
.
Journal of Colloid and Interface Science
31
(
1
),
116
130
.
Mironyuk
I.
Tatarchuk
T.
Naushad
M.
Vasylyeva
H.
Mykytyn
I.
2019
Highly efficient adsorption of strontium ions by carbonated mesoporous TiO2
.
Journal of Molecular Liquids
285
,
742
753
.
Moradi-Rad
R.
Omidi
L.
Kakooei
H.
Golbabaei
F.
Hassani
H.
Abedin-Loo
R.
Azam
K.
2014
Adsorption of polycyclic aromatic hydrocarbons on activated carbons: kinetic and isotherm curve modeling
.
International Journal of Occupational Hygiene
6
(
1
),
43
49
.
Müller
S.
Totsche
K. U.
Kögel-Knabner
I.
2007
Sorption of polycyclic aromatic hydrocarbons to mineral surfaces
.
European Journal of Soil Science
58
(
4
),
918
931
.
Njoku
V. O.
Islam
M. A.
Asif
M.
Hameed
B. H.
2014
Utilization of sky fruit husk agricultural waste to produce high quality activated carbon for the herbicide bentazon adsorption
.
Chemical Engineering Journal
251
,
183
191
.
Nkansah
M. A.
Christy
A. A.
Barth
T.
Francis
G. W.
2012
The use of lightweight expanded clay aggregate (LECA) as sorbent for PAHs removal from water
.
Journal of Hazardous Materials
217
,
360
365
.
O'Brine
T.
Thompson
R. C.
2010
Degradation of plastic carrier bags in the marine environment
.
Marine Pollution Bulletin
60
(
12
),
2279
2283
.
Oller
I.
Malato
S.
Sánchez-Pérez
J.
2011
Combination of advanced oxidation processes and biological treatments for wastewater decontamination – a review
.
Science of the Total Environment
409
(
20
),
4141
4166
.
Paune
F.
Caixach
J.
Espadaler
I.
Om
J.
Rivera
J.
1998
Assessment on the removal of organic chemicals from raw and drinking water at a Llobregat river water works plant using GAC
.
Water Research
32
(
11
),
3313
3324
.
Peng
R. H.
Fu
X. Y.
Zhao
W.
Tian
Y. S.
Zhu
B.
Han
H. J.
Xu
J.
Yao
Q. H.
2014
Phytoremediation of phenanthrene by transgenic plants transformed with a naphthalene dioxygenase system from Pseudomonas
.
Environmental Science & Technology
48
(
21
),
12824
12832
.
Pérez-Gregorio
M. R.
García-Falcón
M. S.
Martínez-Carballo
E.
Simal-Gándara
J.
2010
Removal of polycyclic aromatic hydrocarbons from organic solvents by ashes wastes
.
Journal of Hazardous Materials
178
(
1–3
),
273
281
.
Ping
L.
Luo
Y.
Wu
L.
Qian
W.
Song
J.
Christie
P.
2006
Phenanthrene adsorption by soils treated with humic substances under different pH and temperature conditions
.
Environmental Geochemistry and Health
28
(
1–2
),
189
195
.
Puszkarewicz
A.
Kaleta
J.
2020
The efficiency of the removal of naphthalene from aqueous solutions by different adsorbents
.
International Journal of Environmental Research and Public Health
17
(
16
),
5969
.
Putra
E. K.
Pranowo
R.
Sunarso
J.
Indraswati
N.
Ismadji
S.
2009
Performance of activated carbon and bentonite for adsorption of amoxicillin from wastewater: mechanisms, isotherms and kinetics
.
Water Research
43
(
9
),
2419
2430
.
Qamar
Z.
Khan
S.
Khan
A.
Aamir
M.
Nawab
J.
Waqas
M.
2017
Appraisement, source apportionment and health risk of polycyclic aromatic hydrocarbons (PAHs) in vehicle-wash wastewater, Pakistan
.
Science of the Total Environment
605
,
106
113
.
Qiu
T.
Zeng
Y.
Ye
C.
Tian
H.
2012
Adsorption thermodynamics and kinetics of p-xylene on activated carbon
.
Journal of Chemical & Engineering Data
57
(
5
),
1551
1556
.
Ravindra
K.
Sokhi
R.
Van Grieken
R.
2008
Atmospheric polycyclic aromatic hydrocarbons: source attribution, emission factors and regulation
.
Atmospheric Environment
42
(
13
),
2895
2921
.
Saucier
C.
Karthickeyan
P.
Ranjithkumar
V.
Lima
E. C.
Dos Reis
G. S.
de Brum
I. A.
2017
Efficient removal of amoxicillin and paracetamol from aqueous solutions using magnetic activated carbon
.
Environmental Science and Pollution Research
24
(
6
),
5918
5932
.
Shih
K. L.
Lederberg
J.
1976
Chloramine mutagenesis in Bacillus subtilis
.
Science
192
(
4244
),
1141
1143
.
Smol
M.
Włodarczyk-Makuła
M.
Włóka
D.
2014
Adsorption of polycyclic aromatic hydrocarbons (PAHs) from aqueous solutions on different sorbents. Civil and Environmental Engineering Reports.
Soni
S.
Bajpai
P. K.
Mittal
J.
Arora
C.
2020
Utilisation of cobalt doped iron based MOF for enhanced removal and recovery of methylene blue dye from waste water
.
Journal of Molecular Liquids
314
,
113642
.
Su
Y. H.
Zhu
Y. G.
Sheng
G.
Chiou
C. T.
2006
Linear adsorption of nonionic organic compounds from water onto hydrophilic minerals: silica and alumina
.
Environmental Science & Technology
40
(
22
),
6949
6954
.
Subramani
M.
Sepperumal
U.
2016
FTIR analysis of bacterial mediated chemical changes in polystyrene foam
.
Annals of Biological Research
7
(
5
),
55
61
.
Sweetman
A. J.
Jones
K. C.
2000
Declining PCB concentrations in the UK atmosphere: evidence and possible causes
.
Environmental Science & Technology
34
(
5
),
863
869
.
Tantawy
M. A.
Shatat
M. R.
El-Roudi
A. M.
Taher
M. A.
Abd-El-Hamed
M.
2014
Low temperature synthesis of belite cement based on silica fume and lime
.
International Scholarly Research Notices
2014
, 1–10.
Tatarchuk
T.
Paliychuk
N.
Bitra
R. B.
Shyichuk
A.
Mironyuk
I.
Ziółkowska
D.
2019
Adsorptive removal of toxic Methylene Blue and Acid Orange 7 dyes from aqueous medium using cobalt-zinc ferrite nanoadsorbents
.
Desalination and Water Treatment
150
,
374
385
.
Umpierres
C. S.
Thue
P. S.
Lima
E. C.
Reis
G. S. D.
de Brum
I. A.
Alencar
W. S. D.
Dias
S. L.
Dotto
G. L.
2018
Microwave-activated carbons from tucumã (Astrocaryum aculeatum) seed for efficient removal of 2-nitrophenol from aqueous solutions
.
Environmental Technology
39
(
9
),
1173
1187
.
USEPA
1994
Agency for Toxic Substances and Disease Registry. Toxicological profi le for polycyclic aromatic hydrocarbons (PAHs): Update
.
US Department of Health and Human Services, Public Health Services
,
Atlanta, GA
.
USEPA
2008
Polycyclic Aromatic Hydrocarbons (PAHS) Fact Sheet
.
National Center for Environmental Assessment, Office of Research and Development
,
Washington, DC
.
Valderrama
C.
Gamisans
X.
De las Heras
X.
Farran
A.
Cortina
J. L.
2008
Sorption kinetics of polycyclic aromatic hydrocarbons removal using granular activated carbon: intraparticle diffusion coefficients
.
Journal of Hazardous Materials
157
(
2–3
),
386
396
.
Vane
C. H.
Kim
A. W.
Beriro
D. J.
Cave
M. R.
Knights
K.
Moss-Hayes
V.
Nathanail
P. C.
2014
Polycyclic aromatic hydrocarbons (PAH) and polychlorinated biphenyls (PCB) in urban soils of Greater London, UK
.
Applied Geochemistry
51
,
303
314
.
Wang
X.
Sato
T.
Xing
B.
2006
Competitive sorption of pyrene on wood chars
.
Environmental Science & Technology
40
(
10
),
3267
3272
.
Wang
J.
Chen
Z.
Chen
B.
2014
Adsorption of polycyclic aromatic hydrocarbons by graphene and graphene oxide nanosheets
.
Environmental Science & Technology
48
(
9
),
4817
4825
.
Wang
J.
Wang
C.
Huang
Q.
Ding
F.
He
X.
2015
Adsorption of PAHs on the sediments from the Yellow River delta as a function of particle size and salinity
.
Soil and Sediment Contamination: An International Journal
24
(
2
),
103
115
.
Waqas
M.
Khan
S.
Chao
C.
Shamshad
I.
Qamar
Z.
Khan
K.
2014
Quantification of PAHs and health risk via ingestion of vegetable in Khyber Pakhtunkhwa Province, Pakistan
.
Science of the Total Environment
497
,
448
458
.
Weber
W. J.
Jr.
Morris
J. C.
1963
Kinetics of adsorption on carbon from solution
.
Journal of the Sanitary Engineering Division – ASCE
89
,
31
59
.
Witt
G.
1995
Polycyclic aromatic hydrocarbons in water and sediment of the Baltic Sea
.
Marine Pollution Bulletin
31
(
4–12
),
237
248
.
Wright
D. A.
Welbourn
P.
2002
Environmental Toxicology
, Vol.
11
.
Cambridge University Press
, Cambridge, UK.
Xifang
S. U. N.
Chun
L. I.
Zhansheng
W. U.
Xiaolin
X. U.
Ling
R. E. N.
Hongsheng
Z. H. A. O.
2007
Adsorption of protein from model wine solution by different bentonites
.
Chinese Journal of Chemical Engineering
15
(
5
),
632
638
.
Yao
Y.
Gao
B.
Fang
J.
Zhang
M.
Chen
H.
Zhou
Y.
Creamer
A. E.
Sun
Y.
Yang
L.
2014
Characterization and environmental applications of clay–biochar composites
.
Chemical Engineering Journal
242
,
136
143
.
Yuan
M.
Tong
S.
Zhao
S.
Jia
C. Q.
2010
Adsorption of polycyclic aromatic hydrocarbons from water using petroleum coke-derived porous carbon
.
Journal of Hazardous Materials
181
(
1–3
),
1115
1120
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY 4.0), which permits copying, adaptation and redistribution, provided the original work is properly cited (http://creativecommons.org/licenses/by/4.0/).