Journal of Hazardous Materials Advances: Ehsan Mirzaee, Majid Sartaj

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Journal of Hazardous Materials Advances 6 (2022) 100083

Contents lists available at ScienceDirect

Journal of Hazardous Materials Advances


journal homepage: www.elsevier.com/locate/hazadv

Activated carbon‐based magnetic composite as an adsorbent for removal of


polycyclic aromatic hydrocarbons from aqueous phase: Characterization,
adsorption kinetics and isotherm studies
Ehsan Mirzaee∗, Majid Sartaj
Department of Civil Engineering, University of Ottawa, 161 Louis-Pasteur Private, Ottawa, ON, K1N 6N5, Canada

a r t i c l e i n f o a b s t r a c t

Keywords: This study investigated the preparation, characterization, and capacity of a magnetic powder activated carbon
Magnetic powder activated carbon (MPAC) composite to remove polycyclic aromatic hydrocarbons (PAHs) from aqueous solutions. The iron oxide
Polycyclic aromatic hydrocarbons nanoparticles deposited on the PAC surface enabled the adsorbent to be simply and rapidly recovered from the
Adsorption
solutions using a magnetic field. The PAH compounds were adsorbed on the MPAC surface through 𝜋-𝜋 and H-𝜋
𝜋-𝜋 interactions
interactions based on the peak position changes observed in the FTIR spectra of MPAC after adsorption. The PAHs
Adsorption uptake
adsorption by MPAC was relatively fast, reaching equilibrium in 6 h with the removal efficiency ranging from
95.6 to 100.0%. The pseudo-second order model exhibited the best fit of the kinetics data, suggesting that all the
MPAC adsorption sites had an equal affinity for the PAHs compounds and the adsorption process was chemical.
The results of the kinetics experiments also indicated the slower adsorption rate of the higher molecular weight
PAHs due to the slower transfer of these analytes to the accessible adsorption sites of MPAC. The Langmuir model
best described the isotherm adsorption of both low molecular weight (LMW) and high molecular weight (HMW)
PAHs, with an R2 in the range of 0.73–0.96. This model also showed that the MPAC particles had a maximum
adsorption uptake ranging from 8.74 to 11.37 μg/mg for the LMW PAHs and 8.43 to 20.21 μg/mg for the HMW
PAHs, respectively.

1. Introduction Sixteen PAHs have been designated as high-priority pollutants by


various environmental agencies, such as the U.S. Environmental Pro-
Polycyclic aromatic hydrocarbons (PAHs) are a group of persistent tection Agency (USEPA) and the Canadian Council of Ministers of
organic contaminants which contain two to six aromatic rings with low the Environment (CCME, 2010; USEPA, 1986). These PAH compounds
water solubility and vapor pressure (Table 1). These compounds can are toxic, thermodynamically stable, and resistant to biodegradation
travel long distances and contaminate air, soil, and water resources (Abdel-Shafy and Mansour, 2016; Akinpelu et al., 2019; CCME, 2010;
(Abdel-Shafy and Mansour, 2016; Izawa et al., 2021; Wang et al., 2021; Mirzaee et al., 2017; USEPA, 2000), and their discharge into the envi-
Zhai et al., 2022). The main sources of PAHs discharge into the environ- ronment could have severe impacts on human health and living organ-
ment are anthropogenic activities including the processing and handling ism (Cheng et al., 2019; Diggs et al., 2011; Li et al., 2014; Okoli et al.,
of petroleum products and the incomplete combustion of organic mate- 2011). The USEPA high-priority PAHs are difficult to break down once
rial (e.g. coal, wood, and petrol) (Wang et al., 2021). In many commer- dispersed into water or soil due to their low water solubility, high melt-
cial and industrial sites in North America, the concentrations of PAHs ing point, and other distinctive features (Table 1) (Cheng et al., 2019;
exceeded the regional and/or federal standards (Meskar et al., 2018). Meng et al., 2019; Wang et al., 2017).
For example, in the United States, there are about 1300 abandoned con- In recent decades, various treatment techniques have been used to
taminated sites where the soil has been highly contaminated with PAHs remove PAHs from aqueous solutions, including electrochemical pro-
as a result of oil refinery activities (USEPA, 2021). Similar cases have cess (Ajab et al., 2020; Falciglia et al., 2016; Gitipour et al., 2014;
been documented in Canada, with more than 6800 PAH-contaminated Mirzaee et al., 2017), biodegradation (Hassan et al., 2018), chemical
sites mostly located in industrial areas (CCME, 2010). oxidation (Liao et al., 2018), solvent extraction, photolysis, and thermal
desorption (Ahmed et al., 2017; Hamann et al., 2016; Lam et al., 2017).
Adsorption using activated carbon (AC) is another treatment technique


Corresponding author.
E-mail address: emirz085@uottawa.ca (E. Mirzaee).

https://doi.org/10.1016/j.hazadv.2022.100083
Received 9 February 2022; Received in revised form 29 April 2022; Accepted 5 May 2022
2772-4166/© 2022 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/)
E. Mirzaee and M. Sartaj Journal of Hazardous Materials Advances 6 (2022) 100083

Table 1
Physical and chemical characteristics of PAH compounds (Manoli and Samara, 1999).

PAHs properties
PAH compound Structure Molecular weight (g/mol) Solubility 25 0 C (mg.dm−3 ) Log Kow Vapor pressure (Pa) 20 0 C

Naphthalene 128 32 2300 0.0492

Acenaphthene 154 3.4 21000 10−3 – 10−2

Acenaphthylene 152 3.93 12000 10−3 – 10−2

Fluorene 166 1.9 15000 10−3 – 10−2

Phenanthrene 178 1-1.3 29000 6.8 × 10−4

Anthracene 178 0.05-0.07 28000 2 × 10−4

Fluoranthene 202 0.26 340000 10−6 – 10−4

Pyrene 202 0.14 2 × 105 6.9 × 10−9

Benz[a]anthracene ∗ 228 0.01 4 × 105 5 × 10−9

Chrysene∗ 228 0.002 4 × 105 10−11 – 10−6

Benzo[b]fluoranthene ∗ 252 - 4 × 105 10−11 – 10−6

Benzo[b]fluoranthene ∗ 252 - 7 × 106 9.6 × 10−7

Benzo[a]pyrene ∗ 252 0.0038 106 5 × 10−9

Benzo[ghi]perylene 276 0.00026 107 ̴10−9

Indeno[1,2,3-cd]pyrene ∗ 276 - 5 × 107 ̴10−10

Dibenz[a,h]anthracene ∗ 278 0.0005 106 ̴10−10



Those indicated with an asterisk ( ) are considered carcinogenic.

that has extensively been employed for the removal of PAHs due to its including precipitation, co-precipitation, and impregnation. According
fast and easy implementation, high efficiency, low cost, and less or no to the results, the magnetic composite synthesized by the precipitation
harmful by-products (Cheng et al., 2019; Khan et al., 2015; Kumar et al., method and powder AC had the best performance for eliminating PAHs
2019; Solano et al., 2021). The recovery of AC in the adsorption process, from aqueous solution. To continue and advance the previous feasibil-
however, could be challenging if the contaminated aqueous phase con- ity study, the authors conducted a more detailed research to explore
tains other constituents and fine particles. To overcome this potential the affinity of a recoverable magnetic powder activated carbon (MPAC)
issue, the AC properties can be modified to make their recovery and re- composite, synthesized by the precipitation method, for both low molec-
cycling feasible (Ai et al., 2011; Bao et al., 2014; Juang et al., 2018). ular weight (LMW) and high molecular weight (HMW) PAHs in aqueous
Magnetization is one of the modification methods which is used to com- solutions. The objectives were to study adsorption isotherms and kinet-
bine AC and iron oxide to synthesize a recoverable magnetic AC ad- ics and to characterize MPAC using the BET, XRD, FE-SEM, EDS, and
sorbent for water/soil treatment purposes (Ai et al., 2011; Juang et al., FTIR tests. To the best of the authors’ knowledge, this is the first compre-
2018; Lompe et al., 2017). However, it has been reported that the type of hensive study on the application of MPAC for the adsorption of a broad
magnetization can affect the composite characteristics such as specific range of USEPA priority PAHs from an aqueous medium. The previous
surface area, pore volume, and adsorption capacity (Ma et al., 2017; studies either investigated other types of carbon-based materials such as
Oliveira et al., 2002). graphene oxide (Han et al., 2012) and carbon nanotubes (Zhang et al.,
Mirzaee and Sartaj (2021) assessed the feasibility of preparing mag- 2019) or granular AC (GAC) for removal of a limited number of PAH
netic AC using powder and granular AC and different synthesis methods, compounds (Eeshwarasinghe et al., 2018; Inbaraj et al., 2021).

2
E. Mirzaee and M. Sartaj Journal of Hazardous Materials Advances 6 (2022) 100083

2. Experimental the PAHs concentrations ranging from 0.5 to 10 mg/L for the batch
adsorption isotherm experiments. All the working solutions in the flasks
2.1. Materials were agitated at 50 rpm for 1 h to ensure a homogeneous distribution
of the PAH compounds in the solutions. Acetonitrile was used as the
A coconut shell powder AC (PAC) used as the base for organic solvent to prepare the working solutions because it improves the
the magnetic composite was obtained from CalgonCarbon Co. solubilization of PAHs in the background solutions with regard to the
(Pennsylvania, USA). The PAC particles were smaller than 45 applied PAHs concentration range (Crisafully et al., 2008; Vidal et al.,
μm in size. Two PAH standard solutions (10 and 500 mg/L 2011).
in acetonitrile) containing acenaphthene (ACE), chrysene (CHR),
anthracene (ANT), benzo[ghi]perylene (B[ghi]P), benzo[a]pyrene
2.5. Batch adsorption isotherms and kinetics experiments
(B[a]P), benz[a]anthracene (B[a]A), benzo[b]fluoranthene (B[b]F),
benzo[k]fluoranthene (B[k]F), dibenz[a,h]anthracene (D[ah]A), and
The study of adsorption kinetics is essential for understanding the
phenanthrene (PHE) were supplied by Sigma Aldrich Co. (St Louis, USA)
rate of reaction, reaction pathways, and the efficiency of the adsorption
and AccuStandard Co. (Connecticut, USA), respectively. Iron (II) sulfate
process (Costa et al., 2017; Lamichhane et al., 2016). To prepare the
heptahydrate (FeSO4 •7H2 O) was purchased from Sigma-Aldrich Co. (St.
background solutions, 200 mg/L NaN3 and 1 mmol KNO3 were mixed
Louis, USA). Potassium nitrate (KNO3 ), sodium azide (NaN3 ), acetoni-
with 30 mL deionized water in a series of 125 mL conical flasks to inhibit
trile (HPLC grade, >99%), and methylene chloride (HPLC grade, >99%)
any potential biological activities (Gan et al., 2017). The mixtures were
were obtained from VWR (Quebec, Canada). Deionized water used in the
then homogenized on an orbital shaker device at 50 rpm and 24 ± 1 °C
batch adsorption experiments was sourced from a Millipore-Q purifica-
for 0.5 h. The PAHs working solution was added to each flask to adjust
tion system (Millipore Co., Bedford, USA).
the PAHs concentration to 1 mg/L. The solutions inside the flasks were
then mixed for 1 h on the orbital shaker at 50 rpm and 24 ± 1 °C to
2.2. Preparation of PAC/Fe3 O4 composite
become homogeneous. Afterward, 30 mg of MPAC was added to each
flask, and the suspensions were agitated at 100 rpm for preselected time
The MPAC (PAC/Fe3 O4 ) composite was synthesized using a mod-
intervals ranging from 2 min to 6 h. Finally, the used MPAC particles
ified precipitation method. To prepare the composite, 2.78 g ferrous
were recovered from the suspensions by placing a magnetic bar at the
sulfate (FeSO4 .7H2 O) was added to a 250 mL conical flask containing
bottom of the flask and separating the aqueous solution.
100 mL deionized water. The solution was then mixed with 0.5 g of
For the adsorption isotherm experiments, samples with different ini-
PAC, followed by agitation for 30 min using a magnetic stirrer. To pre-
tial concentrations of PAHs ranging from 0.5 to 10 mg/L were prepared.
cipitate the hydrated iron oxide particles, 10 mL NaOH solution (10%)
After mixing with 30 mg of PAC or MPAC composite, the samples were
was added dropwise to the flask. The suspension was subsequently kept
agitated on the orbital shaker at 100 rpm and 24±1 °C for 6 h. The test
under stirring at 100 °C for 1 h to place the synthesized iron oxide par-
duration was determined based on the results of kinetic experiments.
ticles onto PAC. The prepared composite particles were dried overnight
All the tests were conducted in duplicate and the flasks were wrapped
and then repeatedly rinsed with deionized water until their pH became
with aluminum foil to protect them against the potential effects of light.
neutral. The MPAC particles <45 μm were finally separated by a mag-
In the end, the MPAC particles were separated from the samples using
netic bar and then placed in screw cap vials for the adsorption studies
the magnetic bar. In the case of PAC, the adsorbent particles were sep-
(Nethaji et al., 2013).
arated using the centrifugation method. Blank samples were prepared
with the PAHs mixture alone and the results indicated that the adsorp-
2.3. Characterization of prepared MPAC
tion of PAHs to the flask wall was negligible. The residual concentration
of PAHs in the samples was determined by the HPLC instrument, as ex-
The morphology of MPAC and the bare PAC particles was analyzed
plained in Section 2.7. In addition, samples with and without biological
by a field emission scanning electron microscopy (FE-SEM, JSM-7500F,
inhibitors (NaN3 and KNO3 ) were provided to assess their possible in-
JEOL, USA) equipped with an EDS detector (Oxford EDS system). The
teractions with PAHs. The PAH adsorption with MPAC and without the
energy-dispersive spectroscopy (EDS) technique was used to evaluate
use of NaN3 or KNO3 was also explored in the preliminary tests. No dis-
the distribution of magnetite nanoparticles (Fe3 O4 ) in the matrix of the
cernible difference was observed in the PAHs concentration due to the
prepared composite. The EDS analysis was carried out by creating a
presence of the biological inhibitors in the solutions.
spectrum of the elements found in the structure of the PAC and MPAC
The solid phase concentration of PAH compounds adsorbed onto PAC
particles. The nature of the under-evaluation particles was also exam-
or MPAC at equilibrium was calculated using the following equation:
ined using a Rigaku Ultima IV X-ray diffraction analyzer (XRD). All XRD
( )
patterns of MPAC and the bare PAC samples were determined in the C0 − Ce V
range 10° to 80° with a scan speed of 4°/min. Also, Fourier transform qe = (1)
M
infrared (FTIR) spectroscopy (Nicolet 6700 FTIR spectrometer, Thermo
Fischer, USA) was used to assess the performance of MPAC functional Where qe (μg/mg) is the PAHs adsorption uptake of PAC or MPAC,
groups in the PAH adsorption process. The FTIR spectra of MPAC be- C0 and Ce represent the initial and equilibrium concentration of the
fore and after adsorption were obtained in the wavenumber range 400– PAH analytes in the solutions (μg/mL), respectively. V is the volume
4000 cm−1 using 32 scans with a resolution of 4 cm−1 for each sample. of the solution (mL), and M shows the mass of PAC or MPAC used in
The Brunauer–Emmett–Teller (BET) surface area of PAC and MPAC were the adsorption process (mg) (Babakhani and Sartaj, 2020; Gürses et al.,
measured by conducting N2 adsorption-desorption using an Accelerated 2014).
Surface Area & Porosimetry System (ASAP) 2020 at 77 K (Micromeritics The Freundlich, Langmuir, and Temkin isotherm models were se-
Instruments Inc., GA, USA). The porosity distribution of the adsorbents lected to evaluate the PAH adsorption uptake of PAC or MPAC using
was determined by the Barrett-Joyner-Hanlenda (BJH) method. the experimental data. According to the Freundlich model, an adsor-
bent surface contains numerous adsorption sites with varied potential
2.4. Preparation of the mixed PAHs working solutions sorption capacity, leading to multilayer and heterogeneous adsorption
of contaminants onto the adsorbent (Jain et al., 2018; Vidal et al., 2011).
1 mL of a 500 mg/L PAHs standard solution was added to a 5 mL The Freundlich model can be displayed as the following equation:
volumetric flask and diluted with acetonitrile to a final concentration of 1
n
100 mg/L. The prepared solution was then serially diluted to provide qe = KF Ce f (2)

3
E. Mirzaee and M. Sartaj Journal of Hazardous Materials Advances 6 (2022) 100083

where KF is the Freundlich constant known as the adsorption capacity Sahoo and Prelot, 2020). The following equation is used to describe the
parameter and nf shows the adsorption intensity. PSO process:
It has been reported that 1/nf values in the range 0.1–1 are ev- 𝑡 1 1
idence of a favourable adsorption process, suggesting a stronger in- = + 𝑡 (7)
𝑞𝑡 𝐾2 𝑞𝑒2 𝑞𝑒
teraction between the adsorbent surface and the adsorbate molecules
(Babakhani and Sartaj, 2020; Eeshwarasinghe et al., 2018; Vidal et al., where k2 is the rate constant parameter of PSO at equilibrium
2011). (g/μg.min). The values of K2 and qe can be derived from the intercept
The Langmuir isotherm model describes an adsorbent with a finite and slope of the linear plot between t/qt and time (min) (Akinpelu et al.,
number of homogeneous binding sites that all show the same affinity 2019).
for the adsorbate molecules. This model also assumes that each bind- The IPD model considers the importance of the diffusion phe-
ing site only adsorbs one molecule, leading to the formation of a single nomenon during adsorption and describes the movement of the ana-
monolayer of the adsorbate on the adsorbent surface (Jain et al., 2018; lyte from the solution to the solid phase (adsorbent). According to this
Khataee et al., 2013). The following equation represents the Langmuir model, the PAHs adsorption onto MPAC involves two steps: the first
model. step represents the macropore diffusion of PAH compounds from the
KL Ce aqueous solution to the MPAC exterior surface (external diffusion), and
qe = qm . (3) the second step shows the micropore diffusion of PAHs into the MPAC
1 + KL Ce
pores. The following equation (Weber and Morris equation) was applied
where KL is the Langmuir affinity constant (mL/μg), and qm (μg/mg) is to explain the diffusion mechanism of PAHs in the adsorption process:
the adsorption capacity of adsorbent, through which a certain number
of the adsorbate molecules is sorbed by a unit mass of the adsorbent q𝑡 = 𝐾3 𝑡0.5 + A (8)
(μg/mg). Both KL and qm are considered as characteristics of the ad- where K3 is the IPD rate constant (μg/g.min0.5 ), and A displays the thick-
sorbent and adsorbate pair. Using the parameters from the Langmuir ness of the boundary layer (ug/mg). The K3 and A parameters can be
model, the separation factor, RL , which is a dimensionless constant, can calculated by plotting qt versus t0.5 , where the slope of the linear plot
be obtained from the following equation: reflects K3 and the intercept illustrates the A parameter or boundary
1 layer effects (Jain et al., 2018; SenthilKumar et al., 2011).
RL = (4)
1 + C m KL Coefficient of determination (R2 ) and residual root-mean-squared er-
ror (RMSE) were employed to evaluate the goodness of fitting of each
where Cm represents the maximum initial concentration of PAHs and
model with the experimental data (Yang et al., 2006). A higher R2 value
KL is the Langmuir constant. The separation factor is inversely related
(close to 1) and a lower RMSE value (close to 0) indicates a better fit of
to the binding strength (Black et al., 2014). An RL value in the range
the modelling (Wang et al., 2020; Wu et al., 2020).
of 0-1 indicates the favourable adsorption process. However, when RL
approaches 0, the adsorption process becomes irreversible, and when
2.7. PAHs measurement
RL equals 1.0, the adsorption is linear (Mirahsani et al., 2020).
The Temkin isotherm model considers indirect adsorbent/adsorbate
The PAHs concentrations in the samples were determined using
interactions and their effects on the adsorption process. Another phe-
High-Pressure Liquid Chromatography (HPLC, Hewlett–Packard 1100,
nomenon considered by this model is the linear reduction in adsorption
Agilent Technologies, USA) equipped with a reverse phase HPLC col-
heat of all the adsorbate molecules due to the increase in coverage of the
umn (ZORBAX Eclipse PAH, 95Å, 4.6 × 250mm × 5μm) and multi-
adsorbent surface (Ayawei et al., 2017; Karaca et al., 2013; Piccin et al.,
ple wavelength detector (MWD). For this purpose, the treated solu-
2011). The following equilibrium equation is used for the Temkin model
tions were poured into 50 mL centrifuge tubes and then extracted with
to plot PAHs adsorption isotherms of the MPAC composite:
methylene chloride (HPLC grade, >99%) using a vortex mixer (Cole-
RT Parmer CO, QC, Canada). The mixtures were then centrifuged at 1500
qe = (ln KT Ce ) (5)
B rpm for 2 min and the solvents containing the analytes were removed
where R is the universal gas constant (8.314 J/mol/K); T is the con- from the tubes by a pipette (Mahgoub, 2016; Zhang et al., 2010). This
stant parameter related to the heat of adsorption (J/mol); and B and KT process was repeated three times and all the extracts were combined
are the Temkin isotherm and equilibrium binding constants (mL/μg), before being concentrated by Kuderna-Danish (K-D) apparatus. The ex-
respectively. tracts were condensed to 1 mL and exchanged to the highly compatible
solvent for HPLC, acetonitrile, using K-D according to method 8310,
2.6. Kinetics modeling USEPA (USEPA, 1986). Finally, 25μL of the acetonitrile solution was
injected onto the HPLC column to determine the concentration of the
Three kinetic models, (pseudo-first-order (PFO), pseudo-second- PAH analytes. The chromatography was conducted with acetonitrile and
order (PSO), and Intra-particle diffusion (IPD) models, were employed deionized water (95:5, v/v) as the mobile phase using a flow rate of 1
to describe the adsorption kinetics of PAHs onto MPAC. The PFO model mL/min. The PAH analytes in the HPLC column were detected at 210.8,
evaluates the adsorption in the solid-liquid system using the Lagergren’s 224.4, 230.8, 254.4, and 270.9 nm by use of the MWD (Bao et al., 2014;
equation (Eq. (6)): Cao et al., 2016). However, the wavelengths 210.8, 224.4, and 230.8 nm
( ) were mostly used to interpret the studied PAHs concentrations using the
ln qe − qt = lnqe − K1 .t (6)
relevant calibration curves.
where qt and qe introduce the quantity of the PAH compounds (𝜇g/mg)
sorbed at time t and equilibrium, respectively, and k1 (min−1 ) is the rate 3. Results and discussion
constant parameter for PFO at equilibrium. K1 can be obtained from the
linear plot of ln (qe -qt ) versus time (min) (Jain et al., 2018). According 3.1. Characterization of unmodified and magnetic PAC
to this model, one PAH compound is sorbed onto one adsorption site of
the adsorbent surface. The XRD diffraction patterns of the MPAC and unmodified PAC sam-
The PSO model takes into account the chemisorption kinetics from ples are shown in Fig. 1. The two strong diffraction peaks that appeared
liquid solutions. This model assumes that the type of adsorption pro- at 2𝜃 = 23.4° and 2𝜃 = 43.8° in the XRD pattern of PAC reflect the amor-
cess is chemical, and the adsorption rate relies on adsorption ca- phous structure of the PAC particles. The characteristic peaks in the
pacity, not on the concentration of the adsorbate (Fu et al., 2021; XRD pattern of MPAC were found at 30.20 , 35.60 , 43.30 , 53.80 , 57.10 ,

4
E. Mirzaee and M. Sartaj Journal of Hazardous Materials Advances 6 (2022) 100083

Table 2
The pore characteristics of PAC and MPAC based on N2 adsorption-
desorption isotherms.

BET surface Micropore Total pore


Sample area (m2 /g) vol. (cm3 /g) vol. (cm3 /g)

PAC 920.6 0.32 0.41


MPAC 324.7 0.10 0.14

Do et al. (2011) and Li et al. (2017) reported complete removal of


methyl orange and methylene blue from water using MAC composites.
The BET test results of these studies revealed a reduction of 31% and
69% in the surface area of the AC particles after magnetization, respec-
tively. Despite the reduction in the surface area, both studies confirmed
that the MAC composites could retain their adsorption capacities for the
Figure 1. The XRD patterns of unmodified PAC and MPAC composite. contaminants during the treatment process (Do et al., 2011; Li et al.,
2017). Park et al. (2015) observed up to 10 times higher removal of
natural organic matter (NOM) by MPAC prepared using the impreg-
and 62.80 . All these peaks correspond to the cubic spinel structure of nation method, even though its surface area was estimated to be 50%
the magnetic nanoparticles (magnetite or maghemite) formed onto the lower than the surface area of the bare PAC. Kim et al. (2013), how-
MPAC surface (Ao et al., 2008; Song et al., 2015). The spinel reflection ever, reported a decrease in the adsorption of NOM by MPAC com-
at (311), (400), and (440) showing the crystal structure of the iron ox- pared to the bare PAC. The researchers attributed this reduction to
ide nanoparticles are in agreement with the iron oxide standard data the lower micropore volume and surface area of MPAC than the bare
(Joint Committee for Powder Diffraction Studies, Card No. 19-0629) PAC. Also, Zahoor (2014) proposed that MPAC composite adsorbed 20%
(Ghanbari et al., 2021). The broad diffraction peak centred at 2𝜃 = 23.40 less humic acid than the bare PAC and attributed this reduction to the
of the PAC XRD pattern represents the characteristic reflection of the AC lower available surface area of the composite (Lompe et al., 2017).
particles. This peak in the XRD pattern of MPAC, however, emerged as However, the reduction in the adsorption capacity of an MPAC com-
a very small characteristic peak, which was due to the formation of a posite could be due to the lower PAC mass fraction in the composite
layer of magnetite nanoparticles on the PAC surface after the magneti- matrix.
zation (Ao et al., 2008; Bao et al., 2014; Nethaji et al., 2013). The size On the other hand, Lompe et al. (2017) proposed that the attachment
of the synthesized magnetic nanoparticles on MPAC can be determined of iron oxide nanoparticles onto PAC would increase the porosity of
by use of the Scherrer equation as follows (Ao et al., 2008): the MPAC composite. For MPAC with an iron oxide mass fraction of
Kλ 38% and mean size of 17 nm, the researchers estimated the specific
D= (10)
𝛽cos𝜃 pore volume per g of the adsorbent to be ranged from 0.15–0.39 mL/g,
where D represents the size of crystallites (nm), K is the Scherrer con- showing a significant increase compared to the bare PAC. Based on this
stant (K=0.9), 𝜆 is the X-ray wavelength equal to 1.5418, 𝛽 is the full estimate, they concluded that the presence of iron oxide nanoparticles
width at half maximum of the diffraction peak considered for (311) and in the matrix of PAC could compensate for the loss of pore volume in the
(440) plane of the spinel reflections, and 𝜃 is the angle of X-ray diffrac- micropores and mesopores of the adsorbent after the synthesis process
tion. The Scherrer equation (Eq. (10)) yielded values of 14.53 and 11.38 and help the composite maintain its adsorption capacity for the analytes
nm for the two spinel reflections, implying that the produced iron ox- (Lompe et al., 2017).
ide particles maintained a nano-size crystal structure in the composite Figure 3 presents the SEM analysis of the PAC and MPAC samples.
matrix (Ao et al., 2008). The PAC particles (Fig. 3a) had a porous and bulky amorphous structure
Figure 2 presents N2 adsorption-desorption isotherms and pore size with a smooth surface. In the MPAC matrix (Fig. 3b), however, the sur-
distribution of the MPAC and unmodified PAC particles. The hystere- face was covered with the crystalline magnetic nanoparticles distributed
sis loops of both the samples show that their isotherms were of type uniformly onto PAC in the shape of spongy crust (Castro et al., 2009;
I according to IUPAC classification; therefore, they possessed a meso- Jain et al., 2018; Lompe et al., 2017).
porous structure (Ge et al., 2016; Juang et al., 2018; Thommes et al., The EDS analysis was used to gather information on the elemental
2015). In addition, the hysteresis loop of the MPAC isotherm is smaller distribution of magnetite nanoparticles on MPAC. As seen in Fig. 3c,
than the PAC, indicating that the formation of magnetite nanoparticles the MPAC spectrum contains two sharp peaks for Fe and O, in addition
on the PAC surface reduced its total pore volume (Juang et al., 2018). to the peaks of carbon. This implies that the magnetite nanoparticles
The pore size distribution shown in Fig. 2b confirms this conclusion. were appropriately produced and attached to the PAC matrix during
The average pore size of PAC decreased after the magnetization process the synthesis process. Figure 3c further shows that the carbon content
due to the occupation of the pores with the iron oxide nanoparticles. in the structure of the magnetic composite was significantly higher than
Consequently, the micropore volume of the prepared MPAC composite the iron content, implying the primary role of PAC as the adsorbent in
was lower than the unmodified PAC, which is in agreement with the the composite structure (Jain et al., 2018; Lompe et al., 2017; Park et al.,
results reported in the literature (Castro et al., 2009; Do et al., 2011; 2015).
Juang et al., 2018; Silva et al., 2017).
Table 2 shows that the magnetite nanoparticles deposited on the sur- 3.2. FTIR spectra analysis of magnetic ACs
face and inside the pores of PAC reduced the BET surface area, micro-
pore volume, and total pore volume from 920 to 324 m2 /g, 0.32 to 0.1 The FTIR analysis of MPAC before and after PAHs adsorption is
cm3 /g, and 0.41 to 0.14 cm3 /g, respectively, a reduction of nearly 65% shown in Fig. 4. The sharp peaks appeared at 484 and 542 cm−1 in
for each parameter. The literature confirms that the reduction of BET the spectrum of the unloaded MPAC shifted to 509 and 553 cm−1 in
surface area for the AC particles is usually more than 30% after mag- the spectrum of PAH-loaded MPAC, respectively. This can be due to
netization, depending on the ratio of precursors used in the synthesis the stretching vibration of Fe–O bond, indicating the existence of Fe3 O4
process (Castro et al., 2009; Do et al., 2011; Juang et al., 2018; Tu et al., nanoparticles on the surface of MPAC (Jain et al., 2018). The medium
2021). peak at 793 cm−1 in the unloaded MPAC shifted to 789 cm−1 after

5
E. Mirzaee and M. Sartaj Journal of Hazardous Materials Advances 6 (2022) 100083

Figure. 2. (a) N2 adsorption-desorption isotherms and (b) pore size distribution of PAC and MPAC.

the PAHs adsorption. This change is attributed to the deformation out 3.3. Batch adsorption results
of plane of the C–H bond in the PAHs aromatic rings (Costa et al.,
2017). The peak centred at 883 cm−1 in the spectrum of MPAC shifted 3.3.1. Adsorption kinetics of PAHs onto MPAC
to 879 cm−1 after the PAHs adsorption, which corresponds to the hy- The adsorption kinetics of LMW and HMW PAHs onto MPAC were
droxyl groups attached by the hydrogen bonds in the MPAC compos- measured and interpreted by the PFO, PSO, and IPD models. As shown
ite (Awoyemi, 2011; Dong et al., 2018; Jain et al., 2018; Wang et al., in Table 3, all the three models satisfactorily described the kinetics data
2014). Furthermore, the weak signal observed at 1120 cm−1 in the spec- with R2 values ranging from 0.83 to 0.95 for PFO, 0.90 to 0.99 for PSO,
trum of MPAC before the adsorption process shifted to 1131 cm−1 af- and 0.90 to 0.96 for IPD, respectively. The RMSE values obtained for
ter the PAHs uptake, which can be attributed to the stretching vibra- PSO (Table 3) ranged from 0.04 to 0.10 μg/mg for the PAH compounds,
tion of C=C bonds in the carboxylic groups of the magnetic composite which were lower than the values for PFO (0.07 to 0.13 μg/mg) and IPD
(Dong et al., 2018; Inbaraj et al., 2021). The broad bands located at (0.06 to 0.11 μg/mg). Considering the greater R2 and the lower RMSE
3103 cm−1 and 3398 cm−1 in the spectrum of the PAH-loaded MPAC values for PSO, this model was selected as the best fit for the kinetics
can be associated with the stretching vibration of the hydroxyl groups data.
(–OH) (Silva et al., 2017; Wang et al., 2014). All these peak shifts ob- The fitting of the PSO model to the experimental data is also il-
served in the spectrum of PAH-loaded MPAC confirm the formation of lustrated in Fig. 5 using a non-linear regression method. The kinetic
interactions between the hydroxyl and carboxyl groups of MPAC and the curves in Fig. 5a show the rapid adsorption of LMW PAHs within the
aromatic rings of PAH analytes during the adsorption process (Inbaraj first hour of contact time, proposing that this group of PAHs was pri-
et al., 2021). marily sorbed by the easily accessible adsorption sites on the external

6
E. Mirzaee and M. Sartaj Journal of Hazardous Materials Advances 6 (2022) 100083

Figure 3. The SEM images of (a) PAC (b) MPAC (c) Energy-dispersive spectroscopy (EDS) spectra of MPAC.

Table 3
The PFO, PSO, and IPD kinetics parameters for PAHs adsorption onto MPAC (PAH concentration of 1μg/mL and contact time ranging
from 2 to 360 min).

Kinetic models

PFO PSO IPD


2 2
PAH qe (cal) K1 R RMSE qe (cal) K2 R RMSE A K3 R2 RMSE

ACE 0.89 0.55 0.95 0.07 0.93 1.00 0.98 0.04 0.72 0.02 0.96 0.06
PHE 0.95 0.26 0.95 0.07 0.99 0.42 0.99 0.04 0.64 0.02 0.90 0.10
ANT 0.91 0.09 0.90 0.10 0.99 0.12 0.96 0.06 0.38 0.04 0.91 0.10
Ʃ3LMW∗ 2.71 0.24 0.92 0.27 2.87 0.13 0.97 0.16 1.75 0.08 0.92 0.26
B[a]A 0.97 0.04 0.91 0.11 1.04 0.06 0.94 0.08 0.27 0.05 0.91 0.11
CHR 0.99 0.03 0.93 0.09 1.08 0.04 0.95 0.08 0.20 0.05 0.92 0.10
B[b]F 0.97 0.04 0.88 0.12 1.03 0.06 0.93 0.09 0.27 0.05 0.92 0.09
B[k]F 0.97 0.03 0.91 0.10 1.05 0.04 0.94 0.08 0.21 0.05 0.94 0.08
B[a]P 0.93 0.03 0.90 0.10 1.00 0.05 0.94 0.08 0.25 0.04 0.93 0.09
B[ghi]P 0.88 0.05 0.83 0.13 0.93 0.09 0.90 0.10 0.19 0.05 0.90 0.09
D[a,h]A 0.93 0.04 0.85 0.13 0.98 0.07 0.91 0.10 0.19 0.05 0.91 0.09
Ʃ7HMW∗ 6.64 0.03 0.89 0.78 7.10 0.01 0.93 0.61 1.77 0.32 0.93 0.61
Ʃ10PAHs∗ 8.79 0.08 0.85 1.26 9.58 0.01 0.92 0.88 3.52 0.40 0.93 0.83

Ʃ3LMW and Ʃ7HMW represent the sum of individual LMW and HMW PAHs, respectively. The sum of all PAH compounds was shown
by Ʃ10PAHs.

7
E. Mirzaee and M. Sartaj Journal of Hazardous Materials Advances 6 (2022) 100083

Figure 4. FTIR spectra of MPAC before and after the PAHs adsorption (PAH concentration of 1μg/mL and contact time of 6 h).

surface and in the larger-sized mesopores of the MPAC. After the initial represented a poorer fit of this model to the kinetic data. According to
contact time, however, the PAHs adsorption rate was declined due to Eq. (8), the relationship between qt and t1/2 in IPD is linear and the in-
the slower transfer of PAHs to less accessible sites of the MPAC microp- tercept displays the boundary layer effects. If the intra-particle diffusion
ores (Eeshwarasinghe et al., 2018; Valderrama et al., 2008; Yuan et al., phenomenon occurs in the adsorption process, the plot of qt vs. t1/2 will
2010). The adsorption rate of HMW PAHs (Fig. 5b) was slower than the pass through the origin. In this work, however, the line did not pass the
LMW PAHs, taking them around three hours to reach equilibrium. A pos- origin, indicating that other processes might be involved in the adsorp-
sible explanation for this behaviour is that with an increase in the molec- tion of PAHs onto MPAC. Besides, the greater values of A exhibited the
ular weight and the number of aromatic rings, the PAHs show higher more significant effects of the boundary layer on the adsorption process
resistance against diffusion into the magnetic adsorbent. As shown in (Jain et al., 2018).
Table 3, the rate constants of both PSO and PFO models were reduced According to Fig. 5a, the highest removal percentage for ACE, PHE,
as the molecular weight of the PAH compounds increased, confirming a ANT in the first hour of the adsorption process were 94.2%, 98.8%,
slower adsorption rate of HMW PAHs than LMW PAHs in the first hours and 86.4%, respectively, showing that the LMW PAHs adsorption by
of the experiments. the MPAC composite rapidly reached equilibrium. The removal rate of
The fitting of the data to PSO suggests that the kinetics of the ad- the LMW PAHs did not have a remarkable change after 3 h, except for
sorption process are primarily controlled through sharing or exchange of ANT whose removal rate reached 100%. In the first 3 h of the HMW
electrons between the adsorbent (MPAC) surface and adsorbate (PAHs) PAHs sorption, the highest removal rate (100%) was achieved for B[a]A
molecules, leading to the formation of 𝜋-𝜋 and H- 𝜋 interactions. Ho and and B[b]F, whereas the adsorption of B[ghi]P from the solution reached
McKay (1998) showed that PSO was a better predictor of the adsorp- 92.77% at the same contact time. The removal of the rest of HMW PAHs
tion kinetics compared to PFO in terms of the PAHs uptake by AC and ranged from 94.57% to 99.87% after the 3 h contact time, suggesting
soil particles. They also proposed that the PFO model was only appro- a high affinity of the MPAC composte for adsorption of these insoluble
priate for interpreting the initial half an hour of the adsorption process and persistent PAH compounds in water. The high adsorption rates of
(Lamichhane et al., 2016). According to Cheng et al. (2019), the adsorp- the LMW and HMW PAHs after 3 h can be considered beneficial in large-
tion kinetics of naphthalene (NAP), ACE, and PHE onto the porous car- scale applications as it reduces the capital and operational costs of the
bon fitted most satisfactorily to the PSO model. These researchers found treatment process (Inbaraj et al., 2021).
a great agreement between the experimental and calculated adsorption Similar results were obtained by Awoyemi (2011) and
capacities using this model. Long et al. (2008), however, demonstrated Haro et al. (2011), where these researchers reported that the adsorption
that the predicted adsorption capacity for NAP was close to the exper- rate of NAP onto AC was considerably faster than fluorene (FLU). Both
imental values only for PFO, although both PFO and PSO models de- the studies attributed this difference in the kinetic behaviour of NAP
scribed the kinetics data very well (R2 >0.99). and FLU to the lower molar volume and molecular dimensions of the
The IPD model describes two diffusion rates of adsorption for the former compound than the latter one (Eeshwarasinghe et al., 2018).
PAH compounds through the mesopores and micropores of MPAC. Yuan et al. (2010) conducted a study on the adsorption of NAP, PHE,
The diffusion rates of PAHs are inversely proportional to their molec- FLU, PYR, and Fluoranthene (FLUO) using one type of coke-derived
ular weight, hence LMW PAHs diffuse faster than HMW PAHs. porous carbon. According to their results, NAP, which is known as
(Eeshwarasinghe et al., 2018). The R2 values obtained by the IPD for the smallest and most water-soluble PAH compound, was adsorbed
the PAH compounds were lower than the PSO model (Table 3), which faster than the other PAHs with the same concentration. Besides, the

8
E. Mirzaee and M. Sartaj Journal of Hazardous Materials Advances 6 (2022) 100083

Figure 5. Adsorption kinetics of (a) LMW PAHs and (b) HMW PAHs onto MPAC, PSO model (PAH concentration of 1μg/mL and contact time from 2 to 360 min).

adsorption rate obtained for this LMW compound was 94.0% only after the isotherm models in the order of Langmuir> Temkin> Freundlich for
30 min contact time, while this rate for FLUO and PYR was 74.0% LMW PAHs, and Langmuir> Freundlich> Temkin for the HMW PAHs.
and 78.4%, respectively. The porous carbon particles were also able to The relatively greater R2 values of the Langmuir model (0.73–0.96)
adsorb 87.8% of PHE and 88.2% of FLU, which were higher than the compared to the two other isotherm models indicated that Langmuir
rates obtained for FLUO and PYR, possibly due to the lower molecular described the experimental data in a greater congruence. The RMSE
weight and dimension of the two former compounds. values determined by the Langmuir, Freundlich, and Temkin models
for the PAH compounds ranged 0.51–1.73, 0.56–2.35, and 0.78–2.00
3.3.2. Adsorption isotherms of PAHs onto PAC and MPAC μg/mg, respectively (Table 4). These values for the total LMW and HMW
The Langmuir, Freundlich, and Temkin isotherm models were em- PAHs were obtained at 3.40 and 4.04 μg/mg for Langmuir, 5.40 and
ployed to simulate the adsorption isotherm equilibrium data and inter- 4.54 μg/mg for Freundlich, and 3.90 and 5.595 μg/mg for the Temkin
pret PAHs adsorption uptake capacity of the MPAC composite. Table 4 Model. The lesser values of RMSE along with the greater coefficients of
presents the isotherm parameters obtained by non-linear regression determination confirmed that the Langmuir model was the best fit to
analysis of these models and the experimental data. According to the R2 the experimental isotherm data, therefore verifying the monolayer ad-
values shown in the table, the adsorption equilibrium data fitted well to sorption of PAHs onto MPAC (Jain et al., 2018; Li et al., 2017). It means

9
E. Mirzaee and M. Sartaj Journal of Hazardous Materials Advances 6 (2022) 100083

Table 4
Parameters of the Langmuir, Freundlich and Temkin isotherm models obtained for the adsorption of PAH compounds onto MPAC and PAC (the PAHs
concentrations in the range 0.5-10 μg/mL and contact time of 6 h).

Isotherm models

Langmuir Freundlich Temkin

PAH KL qmax R2 RSME RL Kf 1/nf R2 RMSE KT B R2 RMSE

ACE 5.74 9.51 0.73 1.73 0.02 6.68 0.30 0.51 2.35 50.64 1.27 0.64 2.00
PHE 10.14 8.74 0.90 0.99 0.01 6.68 0.28 0.72 1.69 49.98 1.29 0.75 1.60
ANT 1.72 11.37 0.90 1.02 0.05 6.36 0.50 0.81 1.45 15.73 0.95 0.91 0.98
Ʃ3LMW 1.37 29.03 0.88 3.40 0.07 13.38 0.36 0.70 5.40 11.75 0.39 0.84 3.90
B[a]A 1.05 8.43 0.95 0.59 0.09 3.81 0.46 0.93 0.65 48.31 2.11 0.84 1.01
CHR 0.65 9.36 0.95 0.55 0.13 3.31 0.55 0.94 0.64 13.76 1.56 0.91 0.78
B[b]F 0.87 9.01 0.96 0.53 0.10 3.69 0.51 0.94 0.64 20.24 1.65 0.90 0.80
B[k]F 0.38 13.20 0.95 0.59 0.21 3.39 0.67 0.95 0.64 10.07 1.30 0.89 0.90
B[a]P 0.25 20.21 0.93 0.81 0.28 3.91 0.78 0.92 0.85 8.38 1.06 0.87 1.07
B[ghi]P 0.31 14.13 0.96 0.51 0.24 3.14 0.70 0.96 0.56 7.96 1.23 0.91 0.80
D[a,h]A 0.43 13.18 0.95 0.60 0.19 3.68 0.66 0.94 0.66 10.82 1.26 0.90 0.90
Ʃ7HMW 0.07 78.94 0.95 4.04 0.57 7.57 0.61 0.94 4.54 1.74 0.20 0.90 5.95
Ʃ10PAHs 0.09 104.98 0.94 6.87 0.54 12.30 0.54 0.90 8.89 1.44 0.13 0.91 8.64
PAC Langmuir Freundlich Temkin
Ʃ3LMW 0.61 39.83 0.90 2.78 0.14 12.48 0.43 0.87 3.13 30.35 0.60 0.71 4.79
Ʃ7HMW 0.58 85.68 0.99 1.16 0.15 27.80 0.52 0.99 2.39 22.84 0.22 0.86 7.73
Ʃ10PAHs 0.21 152.60 0.99 2.23 0.33 25.31 0.65 0.98 4.18 3.78 0.10 0.96 6.04

that the adsorption sites of MPAC were homogeneously distributed on ter modelling the experimental data with Langmuir, the qmax of GAC
its surface, and the adsorbed molecules do not interact with each other for these PAHs was determined: 0.14, 0.11, 0.15, 0.23, 0.11, and 0.09
during the adsorption process (El Khames Saad et al., 2014). According μg/mg, respectively, with the highest for ANT (Valderrama et al., 2008).
to the isotherm model, the main adsorption mechanism of PAHs con- The maximum PAH adsorption capacities obtained for both PAC and
sists of rapid attraction of these compounds by the adsorption sites of MPAC in this work were higher than those found for GAC in the previ-
MPAC, and consequently, formation of interactions between 𝜋 electrons ous two studies. In addition, a wide range of USEPA priority PAHs was
of the PAHs aromatic rings and carboxyl and hydroxyl functional groups investigated in this work, and the adsorption capacity of the prepared
of MPAC. The PAHs hydrophobicity in the aqueous solution, however, MPAC composite was assessed for each of them. Another benefit of this
can affect the duration of the adsorption process. The more hydrophobic work was the possibility of recovering the MPAC particles from aque-
PAHs are surrounded by water molecules, forming stronger hydropho- ous solution, even after utilizing them in several adsorption processes,
bic interactions and being stabilized in water. Therefore, for this group based on the preliminary tests results. The latter is crucial in terms of
of PAHs, the speed of attraction by the MPAC functional groups can preventing the formation of a new source of pollution by removing the
be slower compared to the less hydrophobic PAHs (e.g., ACE and PHE) PAH-loaded MPAC from solution after treatment. Also, in contrast to
(Costa et al., 2017). conventional separation methods such as centrifugation, the magnetic
In the Langmuir model, the qmax parameter indicates the maxi- recovey has no operational costs.
mum adsorption capacity of contaminant onto adsorbent. As shown The values of separation factor (RL ) obtained for the PAHs adsorption
in Table 4, the qmax values for the adsorbed PAHs ranged from onto MPAC are presented in Table 4. As can be seen from the Table, the
8.74 to 11.37 μg/mg for LMW PAHs and 8.43 to 20.21 μg/mg for RL values were in the range of 0.01 (PHE) to 0.28 (B[a]P) (0<RL <1),
HMW PAHs. There exists only one study in the scientific literature implying that the adsorption process was favourable for all the PAHs
investigating the adsorption capacity of MPAC for individual PAHs. (Eeshwarasinghe et al., 2019; Jain et al., 2018; Yakout et al., 2013). In
Inbaraj et al. (2021) prepared a magnetic adsorbent through a series the case of the Langmuir affinity constant (KL ), its greater values for
of carbonization, thermal activation, and magnetization procedures per- LMW PAHs were an indication of the higher adsorption rates of these
formed on powdered green tea waste biomass (Inbaraj et al., 2021). The compounds than the HMW PAHs, which confirmed the results of the ki-
prepared magnetic adsorbent was used in the adsorption isotherm ex- netics experiments (Anah and Astrini, 2018). According to Table 4, the
periments for removal of only four PAH compounds, i.e., B[b]F, B[a]P, values of 1/nf (Freundlich model) for most of the PAHs were between
CHR, and B[a]A, from aqueous solutions. After fitting the isotherm data 0.28 and 0.78, inferring that the adsorption process was favourable and
with the Langmuir model, the maximum adsorption capacity of MPAC there were strong interactions between the MPAC adsorption sites and
for the four PAH compounds was determined as 28.08, 22.75, 19.14, PAHs (Eeshwarasinghe et al., 2018; Jain et al., 2018; Kalaruban et al.,
and 15.86 μg/mg, respectively, which are relatively higher than the 2019; Rusmin et al., 2015; Yakout et al., 2013). In the Temkin model,
ones obtained in this study. Zhang et al. (2019) examined different the adsorption of PAHs is attributed to the formation of uniform bind-
types of magnetic carbon nanotubes to adsorb one PAH compound, PHE, ing energies between the PAH compounds and the adsorbent functional
from contaminated solutions. The qmax parameter estimated for PHE ad- groups. The values of the equilibrium binding energy constant (KT )
sorption of these adsorbents was in the range of 20.0 to 74.6 μg/mg (From Table 4) decrease with the increase in the number of PAH aro-
(Zhang et al., 2019). Despite the high adsorption capacity of the mag- matic rings, which suggests that the heat of adsorption would linearly
netic carbon nanotubes for PHE, there was no information about their decrease when the MPAC surface is covered with the heavier PAH com-
efficiencies for the adsorption of heavier PAH compounds, which are pounds (Younis et al., 2015).
considered to be more toxic and carcinogenic. The same deficiency was The results of the adsorption isotherm models for the bare PAC are
found for a study conducted by Eeshwarasinghe et al. (2019), where also presented in Table 4 and Fig. 6. All the three isotherm models fitted
the researchers employed GAC with a qmax of 2.63 μg/mg for ACE the adsorption data very well, with the most appropriate correlation and
and 7.36 μg/mg for PHE to remove these two LMW PAHs from water. minimum deviation attained for the Langmuir model (R2 ≥ 0.90, RMSE≤
Valderrama et al. (2008) used GAC for adsorption of a group of LMW 2.78). The maximum adsorption capacities of PAC for the total LMW and
and HMW PAHs, including NAP, ACE, FLU, ANT, PYR, and FLUO. Af- HMW PAHs were 39.83 and 85.68 ug/mg, showing a reduction of nearly

10
E. Mirzaee and M. Sartaj Journal of Hazardous Materials Advances 6 (2022) 100083

Figure 6. Adsorption isotherms of (a) LMW PAHs, (b) HMW PAHs, (c) Ʃ3LMW, Ʃ7HMW, and Ʃ10PAHs (PAH concentration from 0.5 to 10 μg/mL and contact time
of 6 h) (d) Comparison of the PAHs molecular weight and their removal efficiency obtained at the equilibrium of adsorption using MPAC composite.

27% and 9%, respectively, compared with the values gained for MPAC of a textile dye (Basic Yellow 13) from aqueous solutions. The re-
(29.03 and 78.94 ug/mg). This could be attributed to the decrease of searchers reported 94.4% of the contaminant removal from the solu-
the adsorbent surface area after magnetization (the MPAC surface area: tions within 10 min reaction time. After treatment, the recovery of
324.7 m2 /g, Table 2). the contaminant-loaded PAC was carried out using the centrifugation
The application of bare and magnetic PAC was investigated by sev- method.D’Cruz et al. (2020) synthesized MPAC for the removal of pro-
eral studies for removal of different types of organic and inorganic mazine from wastewater. Their findings revealed complete adsorption
contaminants. However, there are no detailed discussions in the lit- of the contaminant by the magnetic composite after 10 min, which was
erature on the adsorption of LMW and HMW PAHs by these adsor- an indication of the MPAC successful synthesis and its high affinity for
bents. Khataee et al. (2013) employed the bare PAC for the removal the contaminant. Tu et al. (2021) investigated the removal of dibenzo-

11
E. Mirzaee and M. Sartaj Journal of Hazardous Materials Advances 6 (2022) 100083

Figure 6. Continued

p-dioxin from aqueous solutions using PAC and MPAC. According to ter using the adsorbent after 2 h contact time. All of the above re-
their results, both the adsorbents were able to sorb more than 97% of sults confirmed the high adsorption capacity of the bare and mag-
the contaminant after 1 h. In another study, Hassani et al. (2018) used netic PAC for. The result obtained by these studies further concludes
a magnetic carbon composite for adsorption of organic dye in water . the high adsorption capacity of PAC and MPAC for various types of
They reported that 93.4% of the contaminant was removed from wa- contaminants.

12
E. Mirzaee and M. Sartaj Journal of Hazardous Materials Advances 6 (2022) 100083

Figure 6a and b demonstrate the isotherms data obtained for ad- equilibrium in 6h with a complete removal obtained for most of the PAH
sorption of LMW and HMW PAHs onto MPAC for a range of PAH con- compounds. The Langmuir model fitted the adsorption isotherm data
centrations from 0.5 to 10 μg/mL at room temperature. The Langmuir the best with R2 values greater than 0.9 for most PAHs. The maximum
model was used as the best model fit to the experimental data to describe adsorption capacity was in the range of 8.74–11.37 μg/mg for LMW
the adsorption behavior of the PAH compounds by MPAC. As shown in PAHs and 8.43–20.21 μg/mg for HMW PAHs according to the Lang-
Fig. 6a, the amount of the sorbed ACE and PHE sharply increased when muir model. The pseudo-second-order (PSO) model provided the best
the concentrations of these two compounds were raised from 0.5 to 2.5 fitting to the data, indicating that the adsorption of PAHs onto MPAC
μg/mL. At the same concentration range, the ANT adsorption uptake was dominated by the chemical adsorption, and the PAHs uptake rate
was slower than the two other LMW PAHs, possibly due to its higher was dependent on MPAC adsorption capacity, not on concentrations of
molar weight and dimension and its slower transfer to the MPAC ad- PAHs. The results of this study revealed that the MPAC composite had
sorption sites. From Fig. 6b, all the HMW PAHs showed a similar ad- a strong affinity for eliminating PAHs of various molecular weights and
sorption behaviour, with lower adsorption equilibrium than the LMW aromatic ring numbers from aqueous solutions.
PAHs at the initial concentrations from 0.5 to 2.5 ug/mL. However, as
it is evident from this figure and Table 4, the maximum adsorption ca- Declaration of Competing Interest
pacity of MPAC for the heavier PAHs, including B[k]F, B[a]P, B[ghi]P,
and D[a,h]A, was higher than the LMW PAHs. It demonstrates that the The authors declare that they have no known competing financial
adsorption of the HMW PAHs by MPAC was a slower process compared interests or personal relationships that could have appeared to influence
to the LMW PAHs; however, they were able to form stronger bonds with the work reported in this paper.
the MPAC surface (Valderrama et al., 2008).
Figure 6c depicts the adsorption capacity of the bare and magnetic Acknowledgments
PAC for the sum of 3 LMW PAHs, 7 HMW PAHs, and 10 PAHs fitted with
the Langmuir model. Despite the lower adsorption capacity of MPAC The authors would like to thank the staff at the Faculty of Engineer-
compared to the bare PAC, the magnetic adsorbent still showed a high ing, Dr. Ataollah Babakhani, and CAMAR Center of the University of Ot-
affinity for the PAHs. Besides, the difficulty in recovering the PAC parti- tawa for their assistance and help during the characterization and anal-
cles after use would limit their application for the treatment of aqueous ysis phases. This study was supported by the Natural Sciences and Engi-
solutions, particularly for larger-scale projects (Juang et al., 2018). neering Research Council of Canada (NSERC) [Grant number: 211162].
In this work, the prepared MPAC composite showed a high affinity
for all the PAHs during the adsorption process. As illustrated in Fig. 6d, References
after the equilibrium of adsorption was attained, almost all the LMW and
Abdel-Shafy, H.I., Mansour, M.S.M., 2016. A review on polycyclic aromatic hydrocarbons:
HMW PAHs were completely removed by MPAC (except for B[a]P and source, environmental impact, effect on human health and remediation. Egypt. J. Pet.
B[ghi], which their removal rate was obtained 95.68 and 95.60%, re- 25, 107–123. doi:10.1016/j.ejpe.2015.03.011.
spectively). It indicates that the magnetic adsorbent was able to adsorb Ahmed, M.B., Zhou, J.L., Ngo, H.H., Guo, W., Thomaidis, N.S., Xu, J., 2017. Progress
in the biological and chemical treatment technologies for emerging contaminant
all the PAH compounds with different molecular weights and numbers of removal from wastewater: a critical review. J. Hazard. Mater. 323, 274–298.
aromatic rings. The higher PAH uptake by MPAC compared to the ranges doi:10.1016/j.jhazmat.2016.04.045.
of PAH adsorption reported in the literature for GAC is consistent with Ai, L., Zhou, Y., Jiang, J., 2011. Removal of methylene blue from aqueous solution by
montmorillonite/CoFe2O4 composite with magnetic separation performance. Desali-
the pore adsorption mechanism. According to this mechanism, smaller nation 266, 72–77. doi:10.1016/j.desal.2010.08.004.
adsorbents like PAC have a larger external surface area and shorter dif- Ajab, H., Isa, M.H., Yaqub, A., 2020. Electrochemical oxidation using Ti/RuO2 anode for
fusion path length, enabling them to adsorb the PAH compounds at a COD and PAHs removal from aqueous solution. Sustain. Mater. Technol. 26, e00225.
doi:10.1016/j.susmat.2020.e00225.
faster and higher rate (Rakowska et al., 2013). Besides, owing to their
Akinpelu, A.A., Ali, M.E., Johan, M.R., Saidur, R., Qurban, M.A., Saleh, T.A., 2019. Poly-
magnetic properties, the MPAC composite used in this work had higher cyclic aromatic hydrocarbons extraction and removal from wastewater by carbon nan-
recovery rates, leading to the removal of all the PAH-loaded MPAC par- otubes: a review of the current technologies, challenges and prospects. Process Saf.
Environ. Prot. 122, 68–82. doi:10.1016/j.psep.2018.11.006.
ticles from the solutions after adsorption.
Anah, L., Astrini, N., 2018. Isotherm adsorption studies of Ni(II) ion removal from aqueous
PAH desorption from the magnetic composite was explored using solutions by modified carboxymethyl cellulose hydrogel. IOP Conf. Ser. Earth Environ.
solid phase extraction method (USEPA, 3550b), and the preliminary Sci. 160. doi:10.1088/1755-1315/160/1/012017.
tests results revealed incomplete removal of PAHs from the MPAC sur- Ao, Y., Xu, J., Fu, D., Yuan, C., 2008. A simple route for the preparation of anatase titania-
coated magnetic porous carbons with enhanced photocatalytic activity. Carbon N. Y.
face. Thermal desorption was another method that was assessed in the 46, 596–603. doi:10.1016/j.carbon.2008.01.009.
preliminary tests for the recovery of PAHs from MPAC. It was observed Awoyemi, A., 2011. Master’s Thesis. Dep. Chem. Eng. Appl. Chem. Univ. Toronto,
that the use of high temperatures in the process disintegrated the mag- pp. 1–146.
Ayawei, N., Ebelegi, A.N., Wankasi, D., 2017. Modelling and interpretation of adsorption
netic composite, indicating that this method was not appropriate for the isotherms. J. Chem. 2017. doi:10.1155/2017/3039817.
PAHs recovey (Mirzaee and Sartaj, 2021). Babakhani, A., Sartaj, M., 2020. Removal of Cadmium (II) from aqueous solution
using tripolyphosphate cross-linked chitosan. J. Environ. Chem. Eng. 8, 103842.
doi:10.1016/j.jece.2020.103842.
4. Conclusions Bao, X., Qiang, Z., Chang, J.H., Ben, W., Qu, J., 2014. Synthesis of carbon-
coated magnetic nanocomposite (Fe3O4 at C) and its application for sulfon-
A magnetic powder activated carbon (MPAC) composite was syn- amide antibiotics removal from water. J. Environ. Sci. (China) 26, 962–969.
doi:10.1016/S1001-0742(13)60485-4.
thesized using the precipitation method to remove PAHs from aqueous Black, R., Sartaj, M., Mohammadian, A., Qiblawey, H.A.M., 2014. Biosorption of Pb
solutions. The FE-SEM and EDS results confirmed the presence of the and Cu using fixed and suspended bacteria. J. Environ. Chem. Eng. 2, 1663–1671.
magnetite (Fe3 O4 ) nanoparticles on the activated carbon base of the doi:10.1016/j.jece.2014.05.023.
Cao, Y., Yang, B., Song, Z., Wang, H., He, F., Han, X., 2016. Wheat straw biochar amend-
prepared composite. The MPAC composite also exhibited a crystalline
ments on the removal of polycyclic aromatic hydrocarbons (PAHs) in contaminated
structure with a wide pore size distribution typical of the MAC family, soil. Ecotoxicol. Environ. Saf. 130, 248–255. doi:10.1016/j.ecoenv.2016.04.033.
as confirmed by the SEM image. According to the FTIR test results, the Castro, C.S., Guerreiro, M.C., Gonçalves, M., Oliveira, L.C.A., Anastácio, A.S., 2009.
Activated carbon/iron oxide composites for the removal of atrazine from aqueous
electron-donor-acceptor interaction (𝜋-𝜋 interactions) was the main ad-
medium. J. Hazard. Mater. 164, 609–614. doi:10.1016/j.jhazmat.2008.08.066.
sorption mechanism formed between the analytes and MPAC. Although CCME, 2010. Canadian soil quality guidelines for the protection of environmental and
there was a decrease of 65% in the BET surface area of the AC parti- human health: Polycyclic aromatic hydrocarbons. Can. Environ. Qual. Guidel. 19.
cles after magnetization, the synthesized MPAC was effective in remov- Cheng, H., Bian, Y., Wang, F., Jiang, X., Ji, R., Gu, C., Yang, X., Song, Y., 2019. Green
conversion of crop residues into porous carbons and their application to efficiently
ing PAHs from the aqueous phase with efficiency in the range of 95.6 remove polycyclic aromatic hydrocarbons from water: Sorption kinetics, isotherms
to 100%. The adsorption of LMW and HMW PAHs onto MPAC reached and mechanism. Bioresour. Technol. 284, 1–8. doi:10.1016/j.biortech.2019.03.104.

13
E. Mirzaee and M. Sartaj Journal of Hazardous Materials Advances 6 (2022) 100083

Costa, J.A.S., de Jesus, R.A., da Silva, C.M.P., Romão, L.P.C., 2017. Efficient adsorption Jain, M., Yadav, M., Kohout, T., Lahtinen, M., Garg, V.K., Sillanpää, M., 2018. Devel-
of a mixture of polycyclic aromatic hydrocarbons (PAHs) by Si–MCM–41 mesoporous opment of iron oxide/activated carbon nanoparticle composite for the removal of
molecular sieve. Powder Technol. 308, 434–441. doi:10.1016/j.powtec.2016.12.035. Cr(VI), Cu(II) and Cd(II) ions from aqueous solution. Water Resour. Ind. 20, 54–74.
Crisafully, R., Milhome, M.A.L., Cavalcante, R.M., Silveira, E.R., De Keukeleire, D., Nasci- doi:10.1016/j.wri.2018.10.001.
mento, R.F., 2008. Removal of some polycyclic aromatic hydrocarbons from petro- Juang, R.S., Yei, Y.C., Liao, C.S., Lin, K.S., Lu, H.C., Wang, S.F., Sun, A.C., 2018.
chemical wastewater using low-cost adsorbents of natural origin. Bioresour. Technol. Synthesis of magnetic Fe3O4/activated carbon nanocomposites with high sur-
99, 4515–4519. doi:10.1016/j.biortech.2007.08.041. face area as recoverable adsorbents. J. Taiwan Inst. Chem. Eng. 90, 51–60.
D’Cruz, B., Madkour, M., Amin, M.O., Al-Hetlani, E., 2020. Efficient and recoverable mag- doi:10.1016/j.jtice.2017.12.005.
netic AC-Fe3O4 nanocomposite for rapid removal of promazine from wastewater. Kalaruban, M., Loganathan, P., Nguyen, T.V., Nur, T., Hasan Johir, M.A., Nguyen, T.H.,
Mater. Chem. Phys. 240. doi:10.1016/j.matchemphys.2019.122109. Trinh, M.V., Vigneswaran, S., 2019. Iron-impregnated granular activated carbon for
Diggs, D.L., Huderson, A.C., Harris, K.L., Myers, J.N., Banks, L.D., Rekhadevi, P.V., arsenic removal: application to practical column filters. J. Environ. Manage. 239, 235–
Niaz, M.S., Ramesh, A., 2011. Polycyclic aromatic hydrocarbons and digestive tract 243. doi:10.1016/j.jenvman.2019.03.053.
cancers: a perspective. J. Environ. Sci. Heal. - Part C Environ. Carcinog. Ecotoxicol. Karaca, S., Gürses, A., Açişli, Ö., Hassani, A., Kiranşan, M., Yikilmaz, K., 2013.
Rev. 29, 324–357. doi:10.1080/10590501.2011.629974. Modeling of adsorption isotherms and kinetics of Remazol Red RB adsorption
Do, M.H., Phan, N.H., Nguyen, T.D., Pham, T.T.S., Nguyen, V.K., Vu, T.T.T., from aqueous solution by modified clay. Desalin. Water Treat. 51, 2726–2739.
Nguyen, T.K.P., 2011. Activated carbon/Fe3O4 nanoparticle composite: fabrication, doi:10.1080/19443994.2012.749368.
methyl orange removal and regeneration by hydrogen peroxide. Chemosphere 85, Khan, T.A., Chaudhry, S.A., Ali, I., 2015. Equilibrium uptake, isotherm and kinetic studies
1269–1276. doi:10.1016/j.chemosphere.2011.07.023. of Cd(II) adsorption onto iron oxide activated red mud from aqueous solution. J. Mol.
Dong, C.Di, Chen, C.W., Kao, C.M., Chien, C.C., Hung, C.M., 2018. Wood-biochar- Liq. 202, 165–175. doi:10.1016/j.molliq.2014.12.021.
supported magnetite nanoparticles for remediation of PAH-contaminated estuary sed- Khataee, A., Alidokht, L., Hassani, A., Karaca, S., 2013. Response surface analysis of re-
iment. Catalysts 8, 1–13. doi:10.3390/catal8020073. moval of a textile dye by a Turkish coal powder. Adv. Environ. Res. 2, 291–308.
Eeshwarasinghe, D., Loganathan, P., Kalaruban, M., Sounthararajah, D.P., Kandasamy, J., doi:10.12989/aer.2013.2.4.291.
Vigneswaran, S., 2018. Removing polycyclic aromatic hydrocarbons from water using Kim, S., Kim, J., Seo, G., 2013. Iron oxide nanoparticle-impregnated powder-activated
granular activated carbon: kinetic and equilibrium adsorption studies. Environ. Sci. carbon (IPAC) for NOM removal in MF membrane water treatment system. Desalin.
Pollut. Res. 25, 13511–13524. doi:10.1007/s11356-018-1518-0. Water Treat. 51, 6392–6400.
Eeshwarasinghe, D., Loganathan, P., Vigneswaran, S., 2019. Simultaneous removal of Kumar, J.A., Amarnath, D.J., Sathish, S., Jabasingh, S.A., Saravanan, A., Hemavathy, R.V.,
polycyclic aromatic hydrocarbons and heavy metals from water using granular acti- Anand, K.V., Yaashikaa, P.R., 2019. Enhanced PAHs removal using pyrolysis-assisted
vated carbon. Chemosphere 223, 616–627. doi:10.1016/j.chemosphere.2019.02.033. potassium hydroxide induced palm shell activated carbon: batch and column investi-
El Khames Saad, M., Khiari, R., Elaloui, E., Moussaoui, Y., 2014. Adsorption of an- gation. J. Mol. Liq. 279, 77–87. doi:10.1016/j.molliq.2019.01.121.
thracene using activated carbon and Posidonia oceanica. Arab. J. Chem. 7, 109–113. Lam, S.S., Liew, R.K., Wong, Y.M., Yek, P.N.Y., Ma, N.L., Lee, C.L., Chase, H.A., 2017.
doi:10.1016/j.arabjc.2013.11.002. Microwave-assisted pyrolysis with chemical activation, an innovative method to con-
Falciglia, P.P., De Guidi, G., Catalfo, A., Vagliasindi, F.G.A., 2016. Remediation of soils vert orange peel into activated carbon with improved properties as dye adsorbent. J.
contaminated with PAHs and nitro-PAHs using microwave irradiation. Chem. Eng. J. Clean. Prod. 162, 1376–1387. doi:10.1016/j.jclepro.2017.06.131.
296, 162–172. doi:10.1016/j.cej.2016.03.099. Lamichhane, S., Bal Krishna, K.C., Sarukkalige, R., 2016. Polycyclic aromatic hy-
Fu, B., Ferronato, C., Fine, L., Meunier, F., Luis Valverde, J., Ferro Fernandez, V.R., drocarbons (PAHs) removal by sorption: a review. Chemosphere 148, 336–353.
Giroir-Fendler, A., Chovelon, J.M., 2021. Removal of Pemetrexed from aque- doi:10.1016/j.chemosphere.2016.01.036.
ous phase using activated carbons in static mode. Chem. Eng. J. 405, 127016. Li, C., Lu, J., Li, S., Tong, Y., Ye, B., 2017. Synthesis of magnetic microspheres with sodium
doi:10.1016/j.cej.2020.127016. alginate and activated carbon for removal of methylene blue. Materials (Basel) 10.
Gan, X., TENG, Y., REN, W., MA, J., CHRISTIE, P., LUO, Y., 2017. Optimiza- doi:10.3390/ma10010084.
tion of ex-situ washing removal of polycyclic aromatic hydrocarbons from a Li, H., Qu, R., Li, C., Guo, W., Han, X., He, F., Ma, Y., Xing, B., 2014. Selective re-
contaminated soil using nano-sulfonated graphene. Pedosphere 27, 527–536. moval of polycyclic aromatic hydrocarbons (PAHs) from soil washing effluents us-
doi:10.1016/S1002-0160(17)60348-5. ing biochars produced at different pyrolytic temperatures. Bioresour. Technol. 163,
Ge, X., Wu, Zhansheng, Wu, Zhilin, Yan, Y., Cravotto, G., Ye, B.C., 2016. Enhanced PAHs 193–198. doi:10.1016/j.biortech.2014.04.042.
adsorption using iron-modified coal-based activated carbon via microwave radiation. Liao, X., Wu, Z., Li, Y., Luo, J., Su, C., 2018. Enhanced degradation of polycyclic aromatic
J. Taiwan Inst. Chem. Eng. 64, 235–243. doi:10.1016/j.jtice.2016.03.050. hydrocarbons by indigenous microbes combined with chemical oxidation. Chemo-
Ghanbari, F., Hassani, A., Wacławek, S., Wang, Z., Matyszczak, G., Lin, K.Y.A., sphere 213, 551–558. doi:10.1016/j.chemosphere.2018.09.092.
Dolatabadi, M., 2021. Insights into paracetamol degradation in aqueous solu- Lompe, K.M., Menard, D., Barbeau, B., 2017. The influence of iron oxide nanoparticles
tions by ultrasound-assisted heterogeneous electro-fenton process: key operat- upon the adsorption of organic matter on magnetic powdered activated carbon. Water
ing parameters, mineralization and toxicity assessment. Sep. Purif. Technol. 266. Res. 123, 30–39. doi:10.1016/j.watres.2017.06.045.
doi:10.1016/j.seppur.2021.118533. Long, C., Lu, J., Li, A., Hu, D., Liu, F., Zhang, Q., 2008. Adsorption of naphthalene onto
Gitipour, S., Firouzbakht, S., Mirzaee, E., Alimohammadi, M., 2014. Assessment of the carbon adsorbent from waste ion exchange resin: Equilibrium and kinetic charac-
soil screening levels due to ingestion and dermal absorption of chrysene and teristics. J. Hazard. Mater. 150 (3), 656–661. doi:10.1016/j.jhazmat.2007.05.015.
benzo[k]fluoranthene and appropriate remediation method for Dorson Abad. Envi- Ma, J., Sun, S., Chen, K., 2017. Facile and scalable synthesis of magnetite/carbon adsor-
ron. Monit. Assess. 186. doi:10.1007/s10661-014-3637-5. bents by recycling discarded fruit peels and their potential usage in water treatment.
Gürses, A., Hassani, A., Kranşan, M., Açşl, Ö., Karaca, S., 2014. Removal of methylene Bioresour. Technol. 233, 110–115. doi:10.1016/j.biortech.2017.02.075.
blue from aqueous solution using by untreated lignite as potential low-cost adsorbent: Mahgoub, H.A., 2016. Extraction techniques for determination of polycyclic
kinetic, thermodynamic and equilibrium approach. J. Water Process Eng. 2, 10–21. aromatic hydrocarbons in water samples. Int. J. Sci. Res. 5, 268–272.
doi:10.1016/j.jwpe.2014.03.002. doi:10.21275/v5i1.nov152648.
Hamann, E., Stuyfzand, P.J., Greskowiak, J., Timmer, H., Massmann, G., 2016. The fate Manoli, E., Samara, C., 1999. Polycyclic aromatic hydrocarbons in natural waters:
of organic micropollutants during long-term/long-distance river bank filtration. Sci. sources, occurrence and analysis. TrAC - Trends Anal. Chem. 18, 417–428.
Total Environ. 545–546, 629–640. doi:10.1016/j.scitotenv.2015.12.057. doi:10.1016/S0165-9936(99)00111-9.
Han, Q., Wang, Z., Xia, J., Chen, S., Zhang, X., Ding, M., 2012. Facile and tunable fabrica- Meng, Y., Liu, X., Lu, S., Zhang, T., Jin, B., Wang, Q., Tang, Z., Liu, Y., Guo, X.,
tion of Fe3O4/graphene oxide nanocomposites and their application in the magnetic Zhou, J., Xi, B., 2019. A review on occurrence and risk of polycyclic aromatic
solid-phase extraction of polycyclic aromatic hydrocarbons from environmental water hydrocarbons (PAHs) in lakes of China. Sci. Total Environ. 651, 2497–2506.
samples. Talanta 101, 388–395. doi:10.1016/j.talanta.2012.09.046. doi:10.1016/j.scitotenv.2018.10.162.
Haro, M., Cabal, B., Parra, J.B., Ania, C.O., 2011. On the adsorption kinetics and equilib- Meskar, M., Sartaj, M., Sedano, J.A.I., 2018. Optimization of operational param-
rium of polyaromatic hydrocarbons from aqueous solution. Adsorpt. Sci. Technol. 29, eters of supercritical fluid extraction for PHCs removal from a contaminated
467–478. doi:10.1260/0263-6174.29.5.467. sand using response surface methodology. J. Environ. Chem. Eng. 6, 3083–3094.
Hassan, S.S.M., Abdel-Shafy, H.I., Mansour, M.S.M., 2018. Removal of pyrene doi:10.1016/j.jece.2018.04.048.
and benzo(a)pyrene micropollutant from water via adsorption by green syn- Mirahsani, A., Giorgi, J.B., Sartaj, M., 2020. Ammonia removal from aqueous solution
thesized iron oxide nanoparticles. Adv. Nat. Sci. Nanosci. Nanotechnol. 9. by sodium functionalized graphene oxide: isotherm, kinetics, and thermodynamics.
doi:10.1088/2043-6254/aaa6f0. Desalin. Water Treat. 178, 143–154. doi:10.5004/dwt.2020.24961.
Hassani, A., Eghbali, P., Ekicibil, A., Metin, Ö., 2018. Monodisperse cobalt ferrite nanopar- Mirzaee, E., Gitipour, S., Mousavi, M., Amini, S., 2017. Optimization of to-
ticles assembled on mesoporous graphitic carbon nitride (CoFe2O4/mpg-C3N4): a tal petroleum hydrocarbons removal from Mahshahr contaminated soil using
magnetically recoverable nanocomposite for the photocatalytic degradation of or- magnetite nanoparticle catalyzed Fenton-like oxidation. Environ. Earth Sci. 76.
ganic dyes. J. Magn. Magn. Mater. 456, 400–412. doi:10.1016/j.jmmm.2018.02.067. doi:10.1007/s12665-017-6484-1.
Ho, Y.S., McKay, G., 1998. Sorption of dye from aqueous solution by peat. Chem. Eng. J. Mirzaee, E., Sartaj, M., 2021. Synthesis and evaluation of recoverable activated car-
70, 115–124. doi:10.1016/S1385-8947(98)00076-X. bon/Fe3O4 composites for removal of polycyclic aromatic hydrocarbons from aque-
Inbaraj, B.S., Sridhar, K., Chen, B.H., 2021. Removal of polycyclic aromatic hydrocarbons ous solution. Environ. Technol. Innov. 25, 102174. doi:10.1016/j.eti.2021.102174.
from water by magnetic activated carbon nanocomposite from green tea waste. J. Nethaji, S., Sivasamy, A., Mandal, A.B., 2013. Preparation and characterization of corn
Hazard. Mater. 415, 125701. doi:10.1016/j.jhazmat.2021.125701. cob activated carbon coated with nano-sized magnetite particles for the removal of
Izawa, M., Sakai, M., Mori, J.F., Kanaly, R.A., 2021. Cometabolic benzo[a]pyrene bio- Cr(VI). Bioresour. Technol. 134, 94–100. doi:10.1016/j.biortech.2013.02.012.
transformation by Sphingobium barthaii KK22 proceeds through the kata-annelated Okoli, C.G., Ogbuagu, D.H., Gilbert, C.L., Madu, S., Njoku-Tony, R.F., 2011. Proxi-
ring and 1-pyrenecarboxylic acid to downstream products. J. Hazard. Mater. Adv. 4, mal input of polynuclear aromatic hydrocarbons (PAHs) in groundwater sources
100018. doi:10.1016/j.hazadv.2021.100018. of Okrika Mainland, Nigeria. J. Environ. Prot. (Irvine,. Calif) 02, 848–854.
doi:10.4236/jep.2011.26096.

14
E. Mirzaee and M. Sartaj Journal of Hazardous Materials Advances 6 (2022) 100083

Oliveira, L.C.A., Rios, R.V.R.A., Fabris, J.D., Garg, V., Sapag, K., Lago, R.M., 2002. Acti- Vidal, C.B., Barros, A.L., Moura, C.P., de Lima, A.C.A., Dias, F.S., Vasconcellos, L.C.G.,
vated carbon/iron oxide magnetic composites for the adsorption of contaminants in Fechine, P.B.A., Nascimento, R.F., 2011. Adsorption of polycyclic aromatic hydrocar-
water. Carbon N. Y. 40, 2177–2183. doi:10.1016/S0008-6223(02)00076-3. bons from aqueous solutions by modified periodic mesoporous organosilica. J. Colloid
Park, H.S., Koduru, J.R., Choo, K.H., Lee, B., 2015. Activated carbons impregnated with Interface Sci. 357, 466–473. doi:10.1016/j.jcis.2011.02.013.
iron oxide nanoparticles for enhanced removal of bisphenol A and natural organic Wang, J., Chen, Z., Chen, B., 2014. Adsorption of polycyclic aromatic hydrocarbons by
matter. J. Hazard. Mater. 286, 315–324. doi:10.1016/j.jhazmat.2014.11.012. graphene and graphene oxide nanosheets. Environ. Sci. Technol. 48, 4817–4825.
Piccin, J.S., Dotto, G.L., Vieira, M.L.G., Pinto, L.A.A., 2011. Kinetics and mechanism of the doi:10.1021/es405227u.
food dye FD&C Red 40 adsorption onto chitosan. J. Chem. Eng. Data 56, 3759–3765. Wang, J., Wang, J., Zhao, Z., Chen, J., Lu, H., Liu, G., Zhou, J., Guan, X., 2017. PAHs
doi:10.1021/je200388s. accelerate the propagation of antibiotic resistance genes in coastal water microbial
Rakowska, M.I., Kupryianchyk, D., Grotenhuis, T., Rijnaarts, H.H.M., Koelmans, A.A., community. Environ. Pollut. 231, 1145–1152. doi:10.1016/j.envpol.2017.07.067.
2013. Extraction of sediment-associated polycyclic aromatic hydrocarbons with gran- Wang, J., Zhang, Q., Yang, H., Qiao, C., 2020. Adsorptive desulfurization of organic sul-
ular activated carbon. Environ. Toxicol. Chem. 32, 304–311. doi:10.1002/etc.2066. fur from model fuels by active carbon supported Mn (II): equilibrium, kinetics, and
Rusmin, R., Sarkar, B., Liu, Y., McClure, S., Naidu, R., 2015. Structural evolution of thermodynamics. Int. J. Chem. Eng. 2020. doi:10.1155/2020/2813946.
chitosan-palygorskite composites and removal of aqueous lead by composite beads. Wang, Q., Li, H., Feng, K., 2021. Effect of honeycomb, granular, and powder activated
Appl. Surf. Sci. 353, 363–375. doi:10.1016/j.apsusc.2015.06.124. carbon additives on continuous lactic acid fermentation of complex food waste with
Sahoo, T.R., Prelot, B., 2020. Adsorption Processes for the Removal of Contaminants from mixed inoculation. J. Biosci. Bioeng. doi:10.1016/j.jbiosc.2021.02.009, xxx.
Wastewater, Nanomaterials for the Detection and Removal of Wastewater Pollutants. Wu, Z., Sun, Z., Liu, P., Li, Q., Yang, R., Yang, X., 2020. Competitive adsorption of naphtha-
Elsevier Inc doi:10.1016/b978-0-12-818489-9.00007-4. lene and phenanthrene on walnut shell based activated carbon and the verification:
SenthilKumar, P., Ramalingam, S., Sathyaselvabala, V., Kirupha, S.D., Sivanesan, S., 2011. via theoretical calculation. RSC Adv. 10, 10703–10714. doi:10.1039/c9ra09447d.
Removal of copper(II) ions from aqueous solution by adsorption using cashew nut Yakout, S.M., Daifullah, A.A.M., El-Reefy, S.A., 2013. Adsorption of naphthalene,
shell. Desalination 266, 63–71. doi:10.1016/j.desal.2010.08.003. phenanthrene and pyrene from aqueous solution using low-cost activated car-
Silva, L.A.da, Borges, S.M.S., Paulino, P.N., Fraga, M.A., Oliva, S.T.de, Marchetti, S.G., bon derived from agricultural wastes. Adsorpt. Sci. Technol. 31, 293–302.
Rangel, M.do C., 2017. Methylene blue oxidation over iron oxide supported doi:10.1260/0263-6174.31.4.293.
on activated carbon derived from peanut hulls. Catal. Today 289, 237–248. Yang, K., Zhu, L., Xing, B., 2006. Adsorption of polycyclic aromatic hydrocarbons by car-
doi:10.1016/j.cattod.2016.11.036. bon nanomaterials. Environ. Sci. Technol. 40, 1855–1861. doi:10.1021/es052208w.
Solano, R.A., De León, L.D., De Ávila, G., Herrera, A.P., 2021. Polycyclic aromatic hy- Younis, S.A., El-Gendy, N.S., El-Azab, W.I., Moustafa, Y.M., 2015. Kinetic, isotherm,
drocarbons (PAHs) adsorption from aqueous solution using chitosan beads modified and thermodynamic studies of polycyclic aromatic hydrocarbons biosorption from
with thiourea, TiO2 and Fe3O4 nanoparticles. Environ. Technol. Innov. 21, 101378. petroleum refinery wastewater using spent waste biomass. Desalin. Water Treat. 56,
doi:10.1016/j.eti.2021.101378. 3013–3023. doi:10.1080/19443994.2014.964331.
Song, H.K., Sonkaria, S., Khare, V., Dong, K., Lee, H.T., Ahn, S.H., Kim, H.K., Yuan, M., Tong, S., Zhao, S., Jia, C.Q., 2010. Adsorption of polycyclic aromatic hydrocar-
Kang, H.J., Lee, S.H., Jung, S.P., Adams, J.M., 2015. Pond sediment mag- bons from water using petroleum coke-derived porous carbon. J. Hazard. Mater. 181,
netite grains show a distinctive microbial community. Microb. Ecol. 70, 168–174. 1115–1120. doi:10.1016/j.jhazmat.2010.05.130.
doi:10.1007/s00248-014-0562-7. Zahoor, M., 2014. Removal of humic acid from water through adsorption–ultrafiltration
Thommes, M., Kaneko, K., Neimark, A.V., Olivier, J.P., Rodriguez-Reinoso, F., Rou- hybrid processes. Desalin. Water Treat. 52, 7983–7992.
querol, J., Sing, K.S.W., 2015. Physisorption of gases, with special reference to the Zhai, J., Burke, I.T., Stewart, D.I., 2022. Potential reuse options for biomass combustion
evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure ash as affected by the persistent organic pollutants (POPs) content. J. Hazard. Mater.
Appl. Chem. 87, 1051–1069. doi:10.1515/pac-2014-1117. Adv. 5, 100038. doi:10.1016/j.hazadv.2021.100038.
Tu, Y.J., Premachandra, G.S., Boyd, S.A., Sallach, J.B., Li, H., Teppen, B.J., Johnston, C.T., Zhang, J., Li, R., Ding, G., Wang, Y., Wang, C., 2019. Sorptive removal of phenan-
2021. Synthesis and evaluation of Fe3O4-impregnated activated carbon for dioxin threne from water by magnetic carbon nanomaterials. J. Mol. Liq. 293, 111540.
removal. Chemosphere 263. doi:10.1016/j.chemosphere.2020.128263. doi:10.1016/j.molliq.2019.111540.
USEPA, 2021. National Priorities List (NPL) Sites - by State Superfund [WWW Document]. Zhang, S., Niu, H., Hu, Z., Cai, Y., Shi, Y., 2010. Preparation of carbon coated Fe3O4
URL. nanoparticles and their application for solid-phase extraction of polycyclic aromatic
USEPA, 2000. Deposition of Air Pollutants to the Great Waters, Third Report to Congress. hydrocarbons from environmental water samples. J. Chromatogr. A 1217, 4757–4764.
Office of Air Quality Planning and Standards. doi:10.1016/j.chroma.2010.05.035.
USEPA, 1986. Polycyclic Aromatic Hydrocarbons.
Valderrama, C., Gamisans, X., de las Heras, X., Farrán, A., Cortina, J.L., 2008. Sorp-
tion kinetics of polycyclic aromatic hydrocarbons removal using granular acti-
vated carbon: intraparticle diffusion coefficients. J. Hazard. Mater. 157, 386–396.
doi:10.1016/j.jhazmat.2007.12.119.

15

You might also like