Readings (1)
Readings (1)
Readings (1)
A R T I C L E I N F O A B S T R A C T
Keywords: PFASs have been attracting worldwide attention in recent years. In this study, a novel surface-impregnated
Biochar activated carbon (AC) has been optimally synthesized utilizing waste Karanja shells. Surface functional group
PFAS modification has been performed using Cetyltrimethylammonium bromide (CTAB). PFOA-selectivity analysis of
Emerging contaminants
the adsorbent against different background species present at typical river water concentration was performed.
Adsorbent synthesis
To make the process sustainable, the development of regenerant specifications as well as recovery of the
regenerant and PFOA from the spent regenerant solution were investigated. A biomass/acid ratio of 1:2.5,
temperature 450 ◦ C, and 1 % CTAB solution was found to be optimum synthesis parameters for the CTAB-
modified AC (CTAB_K-AC) for PFOA removal. By CTAB-impregnation, the PFOA uptake capacity of the adsor
bent increased up to threefold. The maximum PFOA adsorption capacities exhibited by Karanja AC (K_AC) and
CTAB_K-AC were 157.1 mg/g & 455.8 mg/g, respectively. The adsorbent exhibited high PFOA selectivity against
all background co-contaminants i.e., NOM (humic acid), clay (bentonite), anions, and cations, and removed ~90
% PFOA. The regenerant (50 % ethanol with 1 % NH4OH aqueous solution) showed 80–90 % PFOA recovery.
The distillation process at ~80 ◦ C temperature recovered ~90 % PFOA and 80–90 % of the regenerant reagents
from the spent regenerant solution. The packed-bed column run of the adsorbent showed high PFOA removal
capacity from the river water matrix and the breakthrough data was well fitted with Thomas and Yoon Nelson
Model with R2 values of 0.97 for both the models. The results showed that novel CTAB_K-AC can be an effective
adsorbent for PFOA adsorptive remediation from aqueous media with high treatment selectivity towards the
PFOA in a materially closed-loop sustainable process.
* Corresponding author.
E-mail address: tnawaz@iitb.ac.in (T. Nawaz).
https://doi.org/10.1016/j.jwpe.2023.103965
Received 30 April 2023; Received in revised form 13 June 2023; Accepted 20 June 2023
Available online 29 June 2023
2214-7144/© 2023 Elsevier Ltd. All rights reserved.
M.A.N. Shaikh et al. Journal of Water Process Engineering 54 (2023) 103965
PFAS compounds by adsorption using commercially available adsor preparing the AC due to their high lignocellulosic carbon content
bents such as synthetic resins, metal oxides, polymers and activated (around 71 %) and effectiveness in attaching the desirable functional
carbon (AC) has been reported [56–61]. The application of AC adsorbent group-containing surfactant compounds such as Cetyl
is a relevant approach for PFAS removal from aqueous media since AC's trimethylammonium bromide (CTAB) (functional group: quaternary
properties are tunable and PFAS-selective functional groups can be ammonium), tert-Butylamine (TBA) (functional group: -NH2) by hy
incorporated within its framework to facilitate high and targeted PFAS drophobic interactions between the surfactants carbon chains with the
removal. carbonaceous matter of the AC [64–66]. Further, to incorporate PFAS
From the literature review, it is evident that many studies have re selective functional groups like quaternary ammonium, amine, etc. in
ported the application of AC as adsorbent for PFAS removal from the carbonaceous moiety of the Karanja AC, we have carried out mod
aqueous media. Deng et al. [1] synthesized granular activated carbon ifications using additives such as CTAB and TBA. The literature search
(GAC) using bamboo for the removal of PFOA and PFOS from synthetic reveals that the present study is the first work that has utilized CTAB-
water, which showed high uptake capacity i.e., 2.32 mmol/g for PFOS impregnated AC from Karanja waste shells for PFAS-specific removal
and 1.15 mmol/g for PFOA, but the PFAS concentration range consid demonstrating high uptake capacity and selectivity.
ered in the study was between 20 and 250 mg/L, which was quite high as Another key drawback of the available studies is the lack of sufficient
compared to their reported concentration under realistic conditions. The investigation into the impact of the background water matrix on the
study reported 98 % regeneration of PFAS-saturated GACs using 50 % PFAS-removal performance of the adsorbents. This is important since
ethanol solution at 45 ◦ C. Inyang and Dickenson [62] conducted a pilot- PFAS is found in trace concentration and the background species are in
scale study for the removal of perfluoroalkyl acids (PFAAs) from lake much higher concentration in the aqueous media and therefore the
water using hardwood biochar and the performance was compared with background species can potentially impact AC's PFAS removal perfor
GAC and anthracite. The results of the study showed that the presence of mance via one or combination of the following three pathways (i) by
DOM in the influent water matrix not only blocked the pores and active screening PFAS from the AC, (ii) the background species themselves can
sites of the adsorbent but also caused a reversal of surface charge on the interact with PFAS and compete with AC for PFAS in the solution and
adsorbent which negatively affected its PFAS removal efficiency. Guo (iii) the background species can compete with PFAS in the solution and
et al. [63] utilized corn straw for the biochar synthesis and reported an can get adsorbed on AC and block the adsorbent sites for PFAS removal.
uptake capacity of 169.6 mg PFOS/g. Similarly, Fagbayigbo et al. [2] Any one or a possible combination of the above three phenomena would
synthesized AC from Vitis vinifera (grape) leaf for adsorptive removal of affect PFAS adsorption on the sorbent. Therefore, a systematic study is
PFOA and PFOS and recorded an uptake capacity of 78.9 mg/g and needed to identify the species that can potentially affect PFAS removal
75.13 mg/g for PFOA and PFOS respectively. In another study, coconut by the AC, and process chemistry needs to be developed to nullify their
shell was utilized for the synthesis of structured and porous biochar impact such that PFAS removal is feasible. Our study is among the few
through alkali activation for the removal of perfluoroalkyl carboxylic studies that has comprehensively explored the role of background spe
acids (PFCAs). Although the biochar showed a high adsorption capacity cies in affecting PFAS removal using AC. We have studied the back
of 1269 mg/g for PFOA, the range of initial PFAS concentrations was up ground species by quantifying their impact on PFAS adsorption using
to 300 mg/L which was a thousand times higher than the realistic values multiple thermodynamic parameters such as selectivity coefficient,
[60]. Steigerwald & Ray, [3] demonstrated the application of ground partitioning fraction ratio, and partitioning coefficient and explored the
spent coffee biochar for PFOS removal with an uptake capacity of 43.4 role of solution pH and species concentration in mitigating their
mg PFOS /g. The effect of divalent cations (Ca2+ & Mg2+ ions) and DOM interference.
was also studied and the results showed that the adsorbent's PFOS Also, only a handful of studies have focused on developing a closed-
removal efficiency was enhanced by 24 % in the presence of divalent loop material utilization process in the context of adsorptive PFAS
cations, but suffered a 60 % reduction due to the DOM. However, no removal where all the raw materials are recovered after their application
regeneration study was conducted for the adsorbent. In another study, in their nearly original form, suitable for further reuse. In this regard, the
Wan et al. [4] employed a surface-modified biochar for PFOA and PFOS regeneration of the PFAS-loaded AC is critical. Therefore, more studies
removal and achieved >85 % efficiency in the presence of a variety of are needed that would investigate regeneration chemistry, determining
typical background species. Moreover, the regeneration through 50 % the regenerant's composition and strength and subsequently recovering
methanol solution exhibited ~90 % efficiency and increased the PFAS PFAS and the regenerant from the spent regenerant solution. In the
concentration up to 25-fold but a critical aspect of recovery of the present study, we have developed a closed-loop material utilization
regenerant and PFAS from the spent regenerant solution was not process with no waste generation where the sorbent has been regener
explored in this study. Table S1 (SI) shows a summary of the major ated to its original form. Moreover, to further make the process sus
studies conducted for adsorptive removal of PFOA through AC. tainable, we have focused not on merely removing PFAS from the
One major aspect that has not been sufficiently explored in the aqueous media but also on recovering them. Therefore, we explored the
literature is utilization of multiple adsorption mechanisms in a single development of regenerant specification and recovery of the regenerant
sorbent that can be leveraged to remove a molecule like PFAS via and PFOA from the spent regenerant solution, required for long-term
multiple removal pathways. There is a need to develop adsorbents reusability of the sorbent and the regenerant and sustainability of the
incorporated with diverse physico-chemical properties that can facili process. To the best of our knowledge, it will be the first comprehensive
tate PFAS removal via a combination of hydrophobic, electrostatic, study for the development of a closed-loop sustainable process with high
hydrogen bonding, and ion-exchange interactions. In this respect, an treatment selectivity and uptake capacity towards the PFOA by an
adsorbent such as AC can be synthesized from precursor materials with adsorbent including the regeneration of adsorbent and recovery of the
high lignocellulosic content that would provide a high carbon yield and regenerant and PFOA.
content upon preparation. Subsequently, chemical additives with In this study, we have achieved: (1) synthesis of novel adsorbent i.e.,
desired functional groups can be impregnated within the AC framework CTAB-impregnated AC of waste Karanja shells, and optimization of the
that can boost electrostatic, hydrogen bonding, or ion-exchange inter adsorbent synthesis parameters, (2) material characterization of the
action with the targeted PFAS adsorbate and improve its selective adsorbent (3) PFOA removal performance analysis of the adsorbent (4)
adsorption. This approach is practical since the water matrix can be of selectivity analysis of the adsorbent for PFOA against typical anions,
diverse composition in nature, therefore an adsorbent with multiple cations, NOM (humic acid) and bentonite clay (5) investigation of
removal mechanisms will be a robust approach to address the problem regeneration potential of sorbents using organic (methanol and ethanol)
of PFAS under different solution chemistry. Guided by this philosophy, and inorganic (NaCl, NaOH, NH4OH, NH4Cl, and KCl) regenerant, (6)
we have tried Karanja waste shells (Pongamia pinnata) as a pre-cursor for recovery of the regenerant and PFOA from the spent regenerant
2
M.A.N. Shaikh et al. Journal of Water Process Engineering 54 (2023) 103965
solution, and (7) design and operate a continuous packed-bed column CTAB concentration were optimized based on PFOA removal perfor
for PFOA removal from the influent containing PFOA for river water mance. Five different impregnation ratios 1:1 to 1:3 was taken for the
matrix. selection of the best biomass/acid ratio. Four different temperatures
300, 450, 600, and 750 ◦ C were taken for the carbonization of acid-
2. Methodology activated biomass, and the obtained biochar was checked for their
PFOA removal capacity by performing batch tests. CTAB (containing
2.1. Material ammonium group in its chain) and TBA (containing amine group in its
chain) solution (0.5, 1.0, 1.5, and 2.0 %) were used for functional group
High purity grade (>99 % assay) Sodium Chloride (NaCl), tert- modification of the activated biochar. After optimization of all the pa
Butylamine TBA (CH₃)₃CNH₂, Hydrochloric acid (99 %), Boric acid rameters, the final Karanja AC was named K_AC and CTAB modified
granular, Sulfuric acid (99 %), Ammonia solution (NH4OH), Ammonium Karanja AC was named CTAB_K-AC.
chloride (NH4Cl), Sodium hydroxide (NaOH), Potassium chloride (KCl),
Calcium Chloride (CaCl2), Bentonite, Magnesium Chloride (MgCl2),
Sodium nitrate (NaNO3), Ortho phosphoric acid (H3PO4), Sodium 2.4. Characterization of the adsorbents
phosphate (Na3PO4), Methanol, Ethanol, Humic acid, Cetyl
trimethylammonium bromide CTAB (C19H42BrN), Sodium bicarbonate 2.4.1. Material characterization
(NaHCO3), and Sodium sulphate (Na2SO4) purchased from Advent The surface morphology of the sorbent was examined using a field
Chembio Pvt. Ltd. were used in the study. Perfluorooctanoic acid emission gun scanning electron microscope (FEG-SEM Quanta 200).
(PFOA) and Acetonitrile (HPLC Grade) were purchased from Sigma Pore volume and surface area of adsorbents were analyzed by Brunner-
Aldrich. Emmet-Teller (BET) method through Quantachrome® ASiQwin™. The
functional groups and nature of chemical bonds present on the surface of
2.2. Synthesis of the adsorbents the adsorbents were observed by Fourier-transformed infrared spec
troscopy (3000-Hyperion, Bruker FTIR) for both before and after CTAB
Waste Karanja shells obtained from Indian Institute of Technology modification along with the PFOA adsorbed surfaces. For identification
Bombay (IITB) campus were utilized for synthesizing biochar adsor of crystallinity, X-ray powder diffraction (X' Pert Pro, PANalytical)
bents. Collected samples were washed in deionized water, air dried for analysis was performed. Elemental analysis was performed by Energy-
48 h, and transferred to the hot air oven (REMI) at 80 ◦ C. A schematic of dispersive X-ray spectroscopy (EDS). To analyze biomass thermal sta
the adsorbent synthesis process is shown in Fig. 1. The dried biomass bility, thermogravimetric analysis (TGA Shimadzu; TGA/DTG-60) was
samples were crushed and sieved in the size range of 75–500 μm. used by observing the change of weight with temperature.
Different activating agents (FeCl3, ZnCl2, HCl, NaOH, H3PO4, H2SO4)
were tried for activation of the biochar. However, on the basis of PFOA 2.4.2. Performance characterization
removal performance of the activated biochar, ortho-phosphoric acid
(15 M) was selected for further application due to its highest PFOA 2.4.2.1. Point of zero charge. The point of zero charge (pHzpc) experi
removal performance. Then five different biomass/acid (mg-Biomass/ ment was performed to measure the surface charge of the adsorbent as
mL of 15 M H3PO4) ratios (1:1 to 1:3) were taken and heated in a muffle previously described by Khan & Sarwar [5]. 0.01 N NaCl solution was
furnace under oxygen-limited conditions at temperatures of 300 and prepared with deionized water and pH was adjusted from 2 to 11 by
750 ◦ C (10 ◦ C/min), with the final temperature being maintained for 1.5 using 0.1 N HCl and 0.1 N NaOH. A fixed amount of the adsorbent was
h. The obtained biochar was rinsed in 0.1 N NaOH aqueous solution and added into each beaker with 100 mL of 0.01 N NaCl solution. The
deionized water until the pH of the wash water was neutral. Finally, the beakers were sealed and stirred for 24 h and the change in pH was
biochar samples were oven-dried at 105 ◦ C for 24 h and stored. A part of observed. The pH value, where zero pH change was observed, is pHzpc.
the biochar samples was loaded with CTAB and TBA to incorporate
tertiary ammonium and amine functional groups respectively in the 2.4.2.2. Batch equilibrium adsorption and rate kinetics studies. Batch
adsorbents. adsorption experiments were performed in an orbital shaker; 25 mg of
the adsorbent was added into the HDPE bottles containing 100 mL of
2.3. Optimization of the synthesis parameters PFOA solution ranging from concentration 1 to 250 mg/L for 24 h. In the
case of the rate kinetic experiments, the pH of the initial solution was
The synthesis parameters like impregnation ratio of the activating fixed at ~5, and the initial concentration of PFOA was 10.0 mg/L. The
agent (biomass/orthophosphoric acid), carbonization temperature, and samples were collected at different time intervals up to 24 h. Analysis of
PFOA was carried out by an ion-chromatograph (Metrohm 925 ECO IC-
04287, 2021) using a C-18 column (Altima 3–250/4.0) as per the
method developed by Metrohm [6]. All the experiments were performed
in triplicates. Percentage removal (Eq. (1)) and adsorption capacity (Eq.
(2)) was determined by using the following equations:
(Ci –Ce )
Removal percentage = *100% (1)
Ce
3
M.A.N. Shaikh et al. Journal of Water Process Engineering 54 (2023) 103965
adsorbent, binary solutions of PFOA and each background species were mode. The PFOA released from the adsorbent was analyzed by
separately prepared in triplicate and the adsorption of PFOA and the measuring the final concentration of PFOA in the spent regenerant. For
background species was observed at equilibrium. Selectivity parameters the selection of the best regenerant specs, organic solvents like ethanol
such as selectivity coefficient and partitioning coefficient were used to and methanol were studied in combination with inorganic regenerants
evaluate the preference of the adsorbent for PFOA over the background like NaCl, NH4Cl, KCl, NaOH, and NH4OH at different concentrations.
species. The background anions considered were carbonate, phosphate, The concentration and combination of the species giving the best per
silicate, sulphate, chloride, and nitrate and the cations considered were centage regeneration of the adsorbent were selected. The percent
calcium and magnesium. The effect of clay and NOM in the solution on regeneration (Eq. (8)) was calculated by using the following equation:
the selective adsorption behaviour of PFOA was checked in the presence
Mass of PFOA leached from the adsorbent
of bentonite and humic acid respectively. For all the selectivity experi Percent regeneration = × 100%
Mass of PFOA loaded on the adsorbent
ments, the adsorbent dose was kept below the optimum adsorption dose,
(8)
such that the limitation of active adsorption sites was maintained to
capture the competition between PFOA and the background species.
2.4.5. Recovery study
When a sorbent is introduced in the solution containing PFOA and the
The PFOA was recovered from the spent regenerant solution (con
background species (denoted as A in Eq. (3) below), depending on
taining 50 % ethanol with 1 % NH4OH aqueous solution) through a
adsorbent selectivity towards them, two simultaneous equilibrium re
distillation process as shown in Fig. 2. A round heating mantle was used
actions will be established.
to provide slow consistent heating for the spent regenerant at ~80 ◦ C. In
PFOA + Adsorbent− − − − − − − − − − − − − − − − Ad − PFOA (K1 ) order to convert NH+ 4 ions to aqueous NH3 (for ease of volatilization and
(3) separation from solution), pH of the spent regenerant was increased up
to 11. The vapours formed were cooled down with a condenser and
A− + Adsorbent − − − − − − − − − − − − − − − − − − Ad − A (K2 ) collected in another flask, which was placed above a magnetic stirrer
(4) with 10 mL of 10 % HCl solution to make the pH of the solution below 5
(for ease of the conversion of aqueous NH3 to dissolved NH+ 4 ions).
where K1 and K2 are the equilibrium constants for the reactions (3) and The recovered PFOA from the left-behind regenerant solution was
(4) respectively. If the adsorbent is selective towards PFOA in solution as collected. The recovered regenerant specification was checked for
compared to the competitive species, then K1 > K2. Therefore, the ammonia, PFOA and total recovered volume and stored for subsequent
selectivity coefficient (Ks) is defined as (Eq. (5)): cycle of regeneration.
K1
Ks = (5) 2.4.6. Packed bed column study
K2
To evaluate the sorbent performance under realistic conditions, a
If Ks > 1, then the adsorbent's selectivity towards PFOA will be fixed-bed mode of operation is preferred over batch mode, therefore a
higher than for the background competitive species. continuous flow packed-bed column study was performed to analyze the
Selectivity can also be studied by understanding how a species par break-through profile of PFOA. The Ganga River water matrix (Table 1)
titions its total mass into solid adsorbent and liquid solution phases. was used with 500 μg/L PFOA concentration. The column setup was
Therefore, partitioning fractional ratio (PFR) has been analyzed (Eq. (6)) used as shown in Fig. 3. For the study, 1.5 g CTAB_K-AC (Particle size:
which describes the fraction of total PFOA existing at equilibrium in the 150–250 μm) was charged as slurry in a column with a bed height of 6
solid adsorbent phase and the liquid solution phase. PFR values >1 cm. The river water pH was adjusted at around ~6.5, since while pre
indicate that for PFOA the solid phase is favoured over the solution paring the synthetic realistic matrix, some precipitation, perhaps of
phase. CaCO3, was noticed. Therefore, to keep all the dissolved river species
Qe *m within the solution, the pH was adjusted to ~6.5 which did not show any
PFR = (6) sign of precipitation. The PFOA-containing influent solution was passed
Ce *V
through the column at 2 mL/min flow rate, maintained by using a
where Qe = adsorption capacity of the adsorbent (mg/g) Ce = final peristaltic pump and empty bed contact time (EBCT) was 3 min. The
concentration of PFOA (mg/L), V = volume of the solution (L) and m = effluent samples were collected bed-volume wise and analyzed for PFOA
mass of the dry adsorbent (g). concentration. The breakthrough profile data was analyzed using
If PFR1 is the partitioning fractional ratio of PFOA in binary solution various mathematical models such as Thomas model and Yoon-Nelson
(along with background species) and PFR2 is the partitioning fractional Model to assess and optimize kinetic constant and predict the experi
ratio of PFOA without the presence of any background species (only mental data. The Thomas model (Eq. (9)) was used to describe the
PFOA containing solution as control). Then, partitioning coefficient (Pc) adsorption kinetics and equilibrium uptake capacity of adsorbent (qth )
can be defined as (Eq. (7)): and rate constant (Kth ) [7]. The Yoon-Nelson Model (Eq. (10)) was used
to determine the rate of adsorption for each adsorbate molecule [8].
PFR1
Pc = (7) Thomas Model:
PFR2
C0
The Pc >1 implies that the presence of background species favour Ce = ( ) (9)
ably affects PFOA removal by the adsorbent and vice-versa.
Kth qo M
1 + exp Q
− Kth C0 t
4
M.A.N. Shaikh et al. Journal of Water Process Engineering 54 (2023) 103965
Component (mg/L) Ganga Water Matrix (mg/L) Actual Taken (mg/L) K′C0 Q τ*C0 *Q
Kyn = ,q =
Ca2+
31.2 32 C50 M yn 1000*M
Mg2+ 21.1 24
Cl− 26.9 27 2.4.7. Analytical technique
NO−3 0.48 5
PO43− 15 15
For PFOA analysis, ion chromatography (Metrohm 925 ECO IC-
SiO2−3 10–30 20 04287, 2021) was used as per the method developed by Metrohm Inc.
HCO3− 164.25 165 [6]. The mobile phase for PFOA analysis was acetonitrile (30 %) and
2−
SO4 23.8 25 boric acid solution (10 mM). The flowrate for sample injection and
PFOA 0.5
–
analysis was 1 mL/ min. For rinsing of the tubing and the IC suppressor,
pH 8.3 6.5
a dilute sulfuric acid solution (50 mmol/L) and deionized water was
used. Avantor C-18 (Alltima 3–250/4.0) column used for separation of
5
M.A.N. Shaikh et al. Journal of Water Process Engineering 54 (2023) 103965
PFOA. The standard calibration curves were obtained for higher PFOA 3. Results and discussion
(0.5–25 mg/L) and lower PFOA (5–500 μg/L) concentration ranges
using standards prepared by high grade PFOA solutions. The concen 3.1. Optimization of the synthesis parameters of the adsorbent
trations were prepared by serial dilution of the standard stock solution of
50 mg/L and 1 mg/L for higher and lower concentration ranges 3.1.1. Effect of biomass/acid ratio and carbonization temperature
respectively. The correlation coefficient (r2) of obtained calibration Fig. 4 (A) shows the result of the effect of varying biomass/acid ratios
curves were 0.9997 & 0.9998 and the relative standard deviation (RSD) on % PFOA removal by the AC. It is evident that the biomass/acid ratio
were 2.921 % & 2.041 % for higher and lower concentration calibration significantly affects the performance of the adsorbent.
curves, respectively (Figs. S1 and S2, SI). The retention time of PFOA With the increase in impregnation ratio from 1:1 to 1:2.5, the PFOA
was around 5.5 min as shown in overlays of chromatograms (Fig. S3, SI). removal percentage initially increased from 4 to 46 % and then slightly
The limit of detection (LOD) for PFOA was calculated by using residual decreased at the higher ratio of 1:3 (Fig. 4 (A)). During the activation
standard deviation (Sy/x) and slope (m) of the calibration curve as re process, H3PO4 spreads inside the biomass, speeding up the swelling,
ported by Subramanian et al., [6]. The obtained values of Sy/x and the and enhancing the bond breakup reactions resulting in the dissolution of
slope of the lower PFOA concentration calibration were 0.4809 and lignin-carbohydrate complexes in the raw feedstock which also facili
0.9977. The LOD value obtained through Eq. (11) was 1.446 μg/L. tates homogenous heating [10]. Microporous carbon is produced at a
lower impregnation ratio which shows lower removal capacity but at
Sy/x
Limit of detection (LOD) = *3.3 (11) higher impregnation ratios micropores transform into mesopores/mac
m
ropores and produce highly porous carbon which exhibits higher uptake
Similarly, the cation and anion analysis were performed through the capacity [11]. FEG-SEM images of AC obtained at different biomass/
IC by using their respective mobile phase and separation column. Before acid ratios also exhibited an increment of porosity and surface irregu
analysis, all the samples were passed through 0.22 μm filter. The cali larity (Fig. 4 (A-I to A-IV)). A similar effect of H3PO4 activation has been
bration curves of anions and cations were made by using a standard previously reported [12]. However, further increasing the ratio from
solution supplied by Merck Sigma-Aldrich. For the quantification of 1:2.5 to 1:3 (Fig. 4 A-IV and A-V), the PFOA removal capacity did not
humic acid and bentonite clay, UV visible spectroscopy in the range of increase due to the shrinking and covering of pores by phosphate and
220 nm to 700 nm was used with a quartz cuvette of path length 1 mm. polyphosphate bonds [12].
The characteristic peaks for bentonite were obtained at 254 nm and for Fig. 4 (B) shows the effect of carbonization temperature on PFOA
humic acid at 232 nm, respectively. removal performance by the AC by varying temperatures between
Fig. 4. The effect of biomass/acid ratio (A) and carbonization temperature (B) on PFOA removal capacity of adsorbent along with respective FEG-SEM images of
adsorbent at different conditions.
6
M.A.N. Shaikh et al. Journal of Water Process Engineering 54 (2023) 103965
300 ◦ C to 750 ◦ C. As the carbonization temperature increased from interaction. However, for TBA loading, PFOA removal capacity of the
300 ◦ C to 600 ◦ C, the removal percentage of PFOA increased from 20 % adsorbent decreased as compared to K_AC. It might be due to the pres
to 47 % (Fig. 4 (B)). As the temperature increases, the porosity of the ence of the primary amine group whose lone pair of electrons are not
adsorbent increases. At temperatures higher than 250 ◦ C, cross-linking readily available for protonation due to π-π interaction between the lone
and polycondensation of H3PO4 molecules along with the elimination pair on the N atom and aromatic ring present in the AC, which prevents
of H2O creates a network structure of H4P2O7, which further enhances the protonation of N atom and thus effectively makes the amine group
the oxidation of carbonaceous material resulting in a more porous site unavailable for electrostatic interaction with the target PFOA
structure [13]. Due to the higher thermal conductivity of H3PO4, a lower molecules.
temperature and shorter time are required for carbonization [14]. FEG-
SEM images obtained at different activation temperatures are shown in
3.2. Adsorbent's material and performance characterization
Fig. 4 (B–I to B-III). The yield of K_AC decreased with the increasing
temperature and reached about 10 % at 750 ◦ C due to higher volatili
3.2.1. BET analysis
zation of the organic fraction of biomass (Fig. S4, SI). TGA analysis of
A nitrogen adsorption isotherm test was performed for the analysis of
Karanja powder was performed to check the effect of temperature on the
the pore size and surface area of the adsorbents. The data obtained from
volatilization of the biomass powder. The results showed that as the
the isotherm between the relative pressure (P/Po) from 0.01 to 1.0 was
temperature increases the volatilization of the biomass increases
plotted by linear isotherm and BET equation was used for the calculation
(Fig. S5, SI). The TGA results were consistent with the carbonization
of surface area. As shown in Fig. S6 (SI), the synthesized adsorbent K_AC
temperature considered in the study and at 450 ◦ C, around 50 % carbon
was observed to exhibit a high surface area of 974.65 m2/g and pore
yield was obtained. As the carbonization temperature is increased above
volume of 0.9 cc/g. In the case of CTAB_K-AC, surface area (438.5 m2/g)
600 ◦ C, thermal shrinkage of H3PO4 starts, resulting in the deposition in
was reduced due to the pore covering by the CTAB molecules on the
the form of chemically inactive polymetaphosphate salts which tend to
surface. Since the length of CTAB molecules (1.5–2 nm) [18] is smaller
reduce the uptake capacity of AC [13].
than the average pore diameter of K_AC (3.7 nm), the CTAB molecules
diffused inside the pores of K_AC, which also comparatively reduced the
3.1.2. Effect of surface modification by CTAB and TBA
pore volume of the CTAB_K_AC (0.6 cc/g). The mesoporous portion was
The results of functional group modification of K_AC showed that
higher in both the adsorbents, however, due to the surface modification
CTAB modification enhanced the PFOA removal capacity whereas TBA
through CTAB, the long alkyl chain caused an increment in the pore size
modification reduced PFOA removal capacity (Fig. 5). Both the surfac
of the CTAB_K-AC. Both the adsorbents exhibited a type H-IV hysteresis
tants were selected due to the presence of hydrophobic chain and
loop for nitrogen adsorption-desorption, which confirms the presence of
ammonium/amine functional groups which has been reported for their
a mesoporous texture (Fig. S6, SI).
selectivity towards PFAS. PFOA removal percent of K_AC was 45 %,
which increase up to 56 %, 79 %, 82 %, and 83 % through the loading of
3.2.2. FTIR analysis
0.5, 1.0, 1.5, and 2.0 % CTAB solution respectively. Since CTAB mole
FTIR spectra obtained by FTIR analysis of Karanja shell powder (KP)
cule contains a long hydrophobic chain and a quaternary ammonium
and the AC before and after CTAB-impregnation is shown in Fig. 6. Due
functional group which increase hydrophobic and electrostatic in
to the presence of amide and amine groups, and alcoholic group, KP
teractions of PFOA with CTAB_K-AC respectively [15]. The cationic
spectra show broad stretching between 3600 and 3200 cm− 1 and
quaternary ammonium head of CTAB molecules points outward from
1200–900 cm− 1, respectively [19]. The presence of acidic functional
the adsorbent's surface towards the solution and facilitates electrostatic
groups (alcoholic and carboxylic) due to the activation of KP with
interaction with the anionic end of PFOA whereas, the alkyl tail of CTAB
phosphoric acid is shown by the broad peak that K_AC spectra displayed
directs inside the pores of the adsorbent [16]. From the results (Fig. 5), 1
between 3600 cm− 1 and 3100 cm− 1, and stretches at 1600–1550 cm− 1
% CTAB solution was selected as the optimum CTAB loading. Patra et al.
was due to C–C vibrations in aromatic rings. The surface modification
[17] also used CTAB for surface modification of Pongamia Pinnata shells
with CTAB was exhibited by the symmetric and asymmetric stretching
derived AC for the removal of Congo Red and Direct Blue-6 dyes which
vibrations of –CH3 and –CH2 group at 2930 and 2850 cm− 1 [20,21]. The
have similar structures as PFOA and established enhanced removal of
exhausted adsorbent showed the presence of strong C–F stretches at
dyes due to the enhancement of electrostatic and hydrophobic
1050–1150 cm− 1, and alkane (–CH2 and -CH3) stretching between 2800
7
M.A.N. Shaikh et al. Journal of Water Process Engineering 54 (2023) 103965
Fig. 6. FTIR spectra obtained by FTIR analysis of Karanja shell powder (KP), CTAB_K-AC, and K_AC before and after PFOA adsorption (K_AC-F).
Fig. 7. The rate kinetics model plots of PFOA adsorption with (a) K_AC and (b) CTAB_K-AC along with used kinetics models and obtained values.
and 2950 cm− 1, which also confirmed the PFOA adsorption on the observed at ~4.5 and it was ~5.3 for CTAB_ K-AC (Fig. S7, SI). Due to
adsorbent surface [17,19]. the presence of quaternary ammonium group on CTAB_K-AC, the pH
value at which it exhibited zero charge increased to 5.3. Moreover, the
3.2.3. Point of zero charge (pHzpc) presence of quaternary ammonium group (which remains positively
The pHzpc is the value of pH at which overall net charge at the charged for all solution pH) will ensure the PFAS removal performance
surface of the adsorbent is zero [22]. The pHzpc value for K_AC was of CTAB_K-AC even under alkaline solution pH conditions via
8
M.A.N. Shaikh et al. Journal of Water Process Engineering 54 (2023) 103965
electrostatic interaction [23,24]. This is important since due to the low The faster equilibrium period observed in our study (4 h) can be
pKa value (~3), PFOA exists predominantly in anionic state in typical explained due to the higher specific surface area (974 m2/g) and larger
environmental conditions, which can facilitate their removal by the pore size (5.54 nm) of the adsorbents, as compared to the other studies
adsorbent prepared in the study. The pHzpc graph obtained from ex (534 m2/g and 12 h equilibrium time for Zhang et al. [28] and 0.5–0.75
periments is shown in Fig. S7 (SI). nm pore size and 12–24 h equilibrium time for Wu et al. [29]).
Fig. 8. Isotherm model plots of PFOA adsorption with (a) K_AC and (b) CTAB_K-AC along with used isotherm models and obtained values.
9
M.A.N. Shaikh et al. Journal of Water Process Engineering 54 (2023) 103965
what is observed during kinetic studies (Qe = 33 mg/g). Therefore, the will be available for the capturing of the anionic PFOA molecules
equilibrium isotherm result is consistent with the rate kinetic study and through electrostatic interaction even under alkaline pH conditions.
validates the isotherm and rate kinetic models. The results with K_AC
were consistent with prior research, but CTAB_K-AC exhibited quite
higher uptake as compared to the previous works. For example, 178.4 3.3. Selectivity Study
mg/g PFOA adsorption uptake capacity using AC is reported by D. Zhang
et al. [30]. A recent study conducted by Z. Wang et al. [31] also used 3.3.1. Effect of background bentonite clay on PFOA adsorption
poly(dimethyldiallylammonium) chloride (PDDA), a cationic polymer Fig. 10 shows selectivity coefficient, partitioning coefficient, and %
containing quaternary ammonium group for surface coating of biochar PFOA removal in the presence of varying initial bentonite clay con
and employed for PFAS remediation. A 10 to 3000-fold increase in PFAS centrations from 0 to 50 mg/L. It was observed that with increasing
sorption was observed after coating of biochar with PDDA. bentonite concentration from 0 to 15 mg/L in 0.01 meq./L PFOA solu
tion, K_AC showed an increasing trend in adsorption for PFOA as
3.2.6. Effect of adsorbent dosage and initial pH depicted in Fig. 10 (c). The PFOA adsorption efficiency increased up to
The constant normality tests were performed for PFOA with both the 99 %, however further increase of bentonite concentration from 15 to
adsorbents to check the effect of the adsorbent dose and the obtained 50 mg/L decreased the PFOA removal efficiency from 99 to 58.27 %.
results are shown in Fig. 9 (a). The results of the study clearly exhibited The results suggest that up to 15 mg/L, bentonite is aiding in the
increment of PFOA removal percentage as the adsorbent dose increased. removal of PFOA by K_AC, and beyond 15 mg/L, bentonite impedes the
For K_AC, the PFOA removal percentage increased from 45 % to 94 % as PFOA removal. This can be attributed to the bridging effect by bentonite
the dose of the adsorbent increases from 5 mg to 25 mg, because of the up to 15 mg/L concentration, as the adsorbed bentonite on K_AC creates
higher availability of the site for adsorption. However, further increase additional sites for hydrophobic interaction with PFOA leading to its
in the adsorbent dose did not improve the removal percentage due to removal. Montmorillonite clay was utilized by M. Wang et al. [33] to
limitation of the sorbate molecules. Almost similar trend was observed remediate PFOA and PFOS, and the authors speculated that the main
for CTAB_K-AC, PFOA removal percent increased from 57 % to 96 % as mechanisms causing PFAS molecules to be trapped were hydrophobic
the adsorbent dose increased from 5 to 20 mg. interaction and hydrogen bonding. This is also reflected in the
The effect of initial pH was analyzed through batch tests and the increasing partitioning of PFOA towards the adsorbent phase (PFR value
results obtained are shown in Fig. 9 (b). The study revealed that the was 44 in the absence of bentonite and increased to 907 in the presence
solution pH plays an important role in adsorption process. In the case of of 15 mg/L bentonite) which also resulted in higher Pc values (Table 2).
K_AC, the removal of PFOA slightly decreased from 90 to 83 % as the pH However, beyond 15 mg/L, bentonite is either overcrowding the K_AC
value increased from 2 to 5 and sharply decreased after a further in active sites and thus prevents PFOA from getting adsorbed or competing
crease of pH from acidic to basic condition. At pH value higher than the with the adsorbent for PFOA and keeping the PFOA in the solution via
pHzpc of the adsorbent, the surface of the adsorbents exhibits negative hydrophobic interaction. Even with a higher concentration of bentonite
surface charge which tends to repel anionic PFOA molecules existing (50 mg/L), higher partitioning towards the solid adsorbent phase was
under alkaline pH conditions. In a study conducted by Fagbayigbo et al. observed as compared to the liquid solution phase. However, Pc values
[2], a similar trend of pH effect was reported when grapes leaf derived came down to 0.15 (Fig. 10 (b)). Further, the selectivity of K_AC towards
AC was used for the PFOA removal. PFOA removal capacity of the PFOA also reduced as bentonite concentration increased from 5 to 50
adsorbent reduced as the conditions changed from acidic (pH 4) to mg/L and the Ks value reduced from 110 to 1.5, due to the adsorption of
alkaline (pH 9). However, for CTAB_K-AC, the pH effect was not sig bentonite on the surface of the adsorbent. It is inferred that increased
nificant and removal capacity remains almost same in the pH range 2–8, adsorption of PFOA might be accounted by the bridging effect provided
since the quaternary ammonium groups present in the chain of CTAB by the adsorbed bentonite on the AC. To confirm bentonite adsorption
remain permanently positively charged, independent of the pH of the on AC, an EDS analysis was performed. The analysis shows the presence
solutions [24,32]. Therefore, the positively charged functional group of silica along with fluoride on the adsorbent. These results also confirm
our observation that bentonite directly binds with the adsorbent
Fig. 9. Effect of (a) adsorbent dose and (b) initial pH on the PFOA removal percentage.
10
M.A.N. Shaikh et al. Journal of Water Process Engineering 54 (2023) 103965
Fig. 10. Effect of bentonite concentration on PFOA adsorption efficiency (0.01 meq./L PFOA, adsorbent dose: 20 mg, solution volume: 100 mL at pH 5).
11
M.A.N. Shaikh et al. Journal of Water Process Engineering 54 (2023) 103965
Fig. 11. Effect of humic acid concentration on PFOA adsorption efficiency (0.01 meq./L PFOA, adsorbent dose: 20 mg, solution volume: 100 mL at pH 5).
humic acid favoured the adsorption (Pc > 1) along with good selectivity
Table 3
towards the PFOA (Ks > 50) by CTAB_K-AC (Table 3), suggesting strong
Selectivity parameters for PFOA with Humic acid variation.
electrostatic interaction of quaternary ammonium groups on CTAB_K-
PFR Partitioning Selectivity coefficient AC overcame the impact of humic acid on PFOA adsorption. Further
coefficient (Pc) (Ks)
studies at lower concentrations of PFOA (1.0, 0.5, and 0.05 mg/L) along
Humic K_AC CTAB_K- K_AC CTAB_K- K_AC CTAB_K- with 5 mg/L humic acid, were conducted to verify the performance of
Acid (mg/ AC AC AC
adsorbent under a realistic scenario. As depicted in Fig. S10 (SI), 90 %
L)
and 58 % PFOA removal was observed at 1.0 and 0.5 mg/L PFOA con
5 862 1940 19.594 4.779 392.913 188.946 centration; however, a considerable PFOA adsorption inhibition was
10 471 822 10.325 2.047 389.875 51.403
observed at 0.05 mg/L PFOA, which resulted in lower (30 %) PFOA
15 305 160 8.838 0.400 176.714 17.878
25 34.5 235.96 0.754 0.537 25.703 74.776 removal. At a concentration ratio of 100 for humic acid/PFOA (mg/mg),
50 9.9 71.4 0.218 0.162 12.614 46.978 PFOA removal was reduced to 30 %, but still, PFOA favoured the solid
adsorbent phase over the liquid phase (PFR = 2.2) (Table S4, SI). A
recent study also reported the negligible effect of NOM (up to 20 mg/L)
availability of less effective surface area of adsorbent for PFOA on PFOA removal through AC [38].
adsorption (J. [35]). A study conducted by Kothawala et al. [36] also
found that DOM preferentially filled pores and adsorption sties of GAC, 3.3.3. Effect of anions on PFOA adsorption
inhibiting their availability for PFAS molecules. But, even with a higher Fig. 12 shows selectivity coefficient, partitioning coefficient, and %
concentration of humic acid (50 mg/L), more partitioning of PFOA to PFOA removal in the presence of different anions (1 meq./L) with 0.01
wards the solid phase was observed as compared to the liquid phase as meq./L PFOA solution. It can be seen that Cl− , NO−3 , and SO2− 4 slightly
also evident from PFR values of 9.9. However, Pc values came down to inhibited the performance of K_AC, and the adsorbent exhibited 95, 91,
0.2, which indicates an unfavourable impact of humic acid on PFOA and 83.9 % PFOA removal (Fig. 12(c)). However, PO3− 2−
4 , SiO3 , and CO3
−
adsorption at 50 mg/L HA with 0.01 meq./L PFOA concentration significantly reduced the PFOA uptake, and the removal reduced from
(Fig. 11(a)). Moreover, the selectivity of K_AC towards PFOA also 93 % to 28.5, 15.5, and 14 % for PO3− 2−
4 , SiO3 , and CO3 , respectively
−
reduced as humic acid concentration increased and the values of Ks (Table 4). It can be seen that PFR values for K_AC are >2 for all anions
reduced from 392 to 12 as depicted in Fig. 11(b). However, the Ks values expect SiO−3 and CO2− 3 . Hence, we can say that the adsorbent phase is
remained >1 indicating that the adsorbent still remains selective to still preferred by PFOA over the solution phase. However, this affinity
wards PFOA in the presence of higher humic acid concentration (5–50 has been somewhat affected in the presence of the anions as reflected by
mg/L). In a similar study of PFOA adsorption on the modified Mont Pc < 1. Moreover, Ks values were quite higher for all anionic species,
morillonite surface, an increase in humic acid concentration had meaning the adsorbent was more selective towards PFOA over the other
retarding effect on adsorbent performance [37]. On the contrary, for anionic species. It was suspected that the low removal efficiency could
CTAB_K-AC, humic acid did not cause any significant effect on the PFOA also be due to the high pH values observed in the solutions of PO3− 4 ,
adsorption efficiency (reduced from 98 % to 93 %) and the obtained SiO−3 , and CO2−
3 ions. Therefore, for these ions, their effect was further
values of PFR remained high (430–71). The presence of 5 and 10 mg/L
12
M.A.N. Shaikh et al. Journal of Water Process Engineering 54 (2023) 103965
Fig. 12. Effect of different anionic background species (1 meq./L) on PFOA adsorption efficiency.
13
M.A.N. Shaikh et al. Journal of Water Process Engineering 54 (2023) 103965
Fig. 13. Effect of different cationic background species (1 meq. /L) on PFOA adsorption efficiency.
Table 5 weakening the electrostatic interaction, NaCl, NaOH, NH4Cl, KCl, and
Effect of the background water matrix (Ganga River) on adsorbent performance NH4OH solutions (5 % w/v) were tested, among these NH4Cl and
along with 0.5 mg/L PFOA: NH4OH showed around 32 % and 25 % regeneration capacity for K_AC
PFOA removal PFR Partitioning and CTAB_K-AC respectively (Fig. 15). Further, NH4OH solution was
percentage coefficient (Pc) tried at 1, 2, 5, and 10 % (w/v), and it exhibits almost the same regen
CTAB_K-AC with realistic 88.09 200.8 5.4 eration capacity (~ 35 %) from 1 to 5 %, however, regeneration capacity
background deceased (18 %) at higher concentrations (10 %) of NH4OH (Fig. 14).
K_AC with realistic 87.85 178 4.9 Similarly, to weaken the hydrophobic interaction between PFOA and
background
the adsorbent, the aqueous solutions of ethanol and methanol were
checked at different concentrations. As ethanol concentration increased
charge neutralization as a possible mechanism [67]. from 5 to 50 %, the regeneration capacity of the regenerant increased
The aqueous concentration of various background anions, cations, from 23 to 60 % for K_AC and 24 to 43 % for CTAB_K-AC. When a similar
organic matter, and clay was taken from the Ganga River water matrix to concentration of methanol was used in the solution, the regeneration
study their combined effect on PFOA adsorption by the adsorbents. The capacity was slightly lower (50 % for K_AC and 39 % for CTAB_K-AC) as
pH of the prepared solution was adjusted to ~6.5. In the presence of the compared to ethanol. Results of the further regeneration experiment
background water matrix, the adsorbent's PFAS removal efficiency revealed that both inorganic and organic regenerant together (NH4OH
decreased from 98 % to 88 %. The result shows that the in presence of a and ethanol) can effectively desorb PFOA from the adsorbent. As the
realistic background water matrix, PFR values were 178 and 200 for concentration of ethanol increased from 10 % to 50 % along with 1 %
K_AC and CTAB_K-AC respectively, which show higher PFAS partition NH4OH, the recovery of PFOA increased from 53 to 92 % for K_AC and
ing towards the adsorbent as compared to the solution phase. These 46 to 85 % for CTAB_K-AC (Fig. 15). Further increase of ethanol content
results revealed that the synthesized adsorbents exhibited good PFOA along with 1 % NH4OH, did not show any enhancement in regeneration
removal performance under realistic scenarios and can remove PFOA in capacity. In a study conducted by Chularueangaksorn et al. [8], the
the presence of higher background concentrations of typical water ma combination of both organic and inorganic regenerant (5 % NaCl with
trix species. The selectivity study of individual species showed that both 70 % methanol) showed 93 % PFOS recovery from the anionic resin
humic acid and bentonite interact with the AC. Similar behaviour was column. A recent study conducted by Wan et al. [4], showed 95 %
observed for the realistic water background matrix as the humic acid desorption of PFOA from modified fly ash through 50 % methanol so
and bentonite were adsorbed by the adsorbent. The 10 % decrease in lution at pH 12. Deng et al. [1] also reported around 95 % PFOA re
PFOA removal efficiency of the adsorbent can be attributed to silicate covery from bamboo-derived AC by using a 50 % ethanol solution at
species, humic acid, and bentonite as shown in Table 5. 45 ◦ C.
Since PFOA is removed via hydrophobic as well as electrostatic in Fig. 16 shows the PFOA breakthrough profile of the packed-bed
teractions, therefore both organic as well as inorganic species were column run. It can be seen from the plot that the breakthrough hap
tested for the PFOA-loaded adsorbent's regeneration. To desorb PFOA by pens with in 2300–2500 bed volumes (breakthrough point-C/Co = 0.1).
14
M.A.N. Shaikh et al. Journal of Water Process Engineering 54 (2023) 103965
Fig. 14. Effect of variation of methanol, ethanol, and NH4OH in regenerant solution on PFOA recovery from adsorbent (PFOA loaded sorbent: 125 mg, Regenerant
volume: 100 mL, time: 8 h, and mixing speed: 160 rpm).
The result of the study revealed that in the presence of the back reduced the probability of both the external (fluid-film) and internal
ground water matrix of the Ganga River, the adsorbent (CTAB_K-AC) can (solid) mass transfer resistances [41]. Therefore, the assumptions of
treat PFOA with good removal efficiency. This is especially helpful for Thomas model were valid for our study, resulting in an excellent fit of
the design of large-scale treatment processes because the adsorbent the model for the experimental data. However, the value of Qmax ob
removed PFOA preferentially when it was present at a trace level against tained through the Thomas model was 5.71 mg/g, which was marginally
a high background of competing species at a typical river water con lower as compared to the Qe obtained through Langmuir isotherm model
centration. The column data were fitted by Thomas and Yoon-Nelson (6.85 mg/g) mainly due to two reasons (a) the presence of the back
model using non-linear regression. The Thomas model assumes ground water matrix during the column run and (b) the size of the
adsorption-desorption kinetics of Langmuir, together with the absence adsorbent was bigger (125–500 μm) than used in the batch equilibrium
of axial dispersion, and the external (fluid-film) and internal (solid) mass and kinetic study (125–250 μm). We noticed that the presence of
transfer resistances are both directly neglected in its derivation [41]. bentonite and humic acid caused pore blockage of the AC bed. There
This indicates that the rate of adsorption is governed by the surface fore, these species were excluded during the column run. Yoon -Nelson
reaction between the adsorbate and the adsorbent's leftover potential. model was also tested to estimate the 50 % break-through time of the
However, Yoon-Nelson model predicts the rate constant (Kyn ) and 50 % column which can be a parameter in the scaling up of the column. Since
adsorbate breakthrough time (τ), which can serve as a baseline for Thomas model and Yoon-Nelson Model are mathematically similar [43],
assessing the capacities of various adsorbent with fixed-bed column the experimental data also fitted well (R2 value 0.97) with the Yoon-
study. The better is the adsorbent, higher will be the (τ). The Yoon- Nelson Model (Fig. 16(b)). The value of 50 % breakthrough (τ) ob
Nelson model does not need data on the system's properties, like the tained through this model was 9375 min, which was closer to experi
kind of adsorbent and the adsorption bed's physical characteristics [42]. mental τ (9660 min) and column sorption capacity was 6.01 mg/g, quite
Fig. 16 (a) shows that the breakthrough experimental data of the comparable to that obtained through the Thomas model.
packed-bed column fitted well with the Thomas model with R2 value of
0.97. Since Thomas model considers Langmuir model in its mathemat 3.5.1. Column regeneration
ical framework, this also validates our batch study findings on isotherm Fig. 17 shows the PFOA concentration profile of column effluent
modeling where Langmuir model has been observed to best describe the when it was regenerated using 50 % ethanol and 1 % NH4OH aqueous
batch experimental data. Since the column diameter was about 1 cm in solution. Around 200-bed volumes of regenerant solution exhibited 89
our study, the axial dispersion can be considered negligible. Further % PFOA recovery from the exhausted column.
more, the size of the adsorbent was small (125–500 μm), which also
15
M.A.N. Shaikh et al. Journal of Water Process Engineering 54 (2023) 103965
Fig. 15. Effect of regenerant specs on PFOA recovery from adsorbent (PFOA loaded sorbent: 125 mg, Regenerant volume: 100 mL, time: 8 h, and mixing speed:
160 rpm).
2
R = 0.97 2
R = 0.97
th
K = 1.82 ml/ mg min '
K = 0.0003 g/ ml
Q0 = 5.71 mg/g
Q0 = 6.01 mg/g
Fig. 16. The plot of breakthrough data from packed-bed column run.
16
M.A.N. Shaikh et al. Journal of Water Process Engineering 54 (2023) 103965
Fig. 17. Effluent PFOA concentration profile in the regeneration of the fixed-bed column
3.6. Recovery study towards PFOA (CTAB_K-AC being more PFOA selective than K_AC) even
with trace levels of PFOA and exhibited high PFOA selectivity against
Using the recovery setup (Fig. 2), PFOA was successfully recovered different background co-contaminants such as natural organic matter
from the regenerant solution. Post-recovery mass balance shows that 90 (humic acid), clay (bentonite), anions (Cl− , NO−3 , SO2− 3−
4 , PO4 , SiO3 , and
−
% PFOA by weight, 80–90 % regenerant by volume, and ~ 85 % CO2−3 ) and cations (Ca2+
and Mg 2+
). In the presence of all the back
ammonium hydroxide by weight was recovered from the spent regen ground species, the adsorbent selectively removed >90 % of PFOA
erant solution. EDS analysis of the recovered PFOA confirmed the present at 0.5 mg/L in the solution. The regeneration study also showed
elemental composition (Fig. S12, SI). It can be observed that the the presence of electrostatic and hydrophobic interaction in the uptake
recovered PFOA has almost similar % elemental composition as of PFOA by the sorbent. Therefore, different organic and inorganic
compared to the PFOA standard. Though it can be seen that the carbon regenerant species were taken and the most desirable specification of 50
% is higher in the case of pure PFOA as compared to recovered PFOA. In % ethanol and 1 % NH4OH aqueous solution was considered which
the XRD analysis (Fig. S13, SI) of recovered PFOA, loss of crystallinity showed 80–90 % PFOA recovery. PFOA was successfully recovered from
was observed as the intensity of the peaks decreases in recovered PFOA the regenerant solution through distillation process. Around 80–90 %
as compared to control. PFOA and ~ 85 % regenerant was recovered from the spent regeneration
solution by distillation. The performance of CTAB_K-AC was checked in
4. Conclusions continuous mode process with river water matrix through packed-bed
column run. The adsorbent showed good PFOA removal capacity with
The aim of the study was to synthesize a PFAS-selective AC with treating up to 2000 bed volumes per gram of the adsorbent. The
multiple PFAS removal pathways and demonstrate a materially closed- breakthrough data of continuous flow packed-bed column was well
loop process for PFAS removal from river water matrix. For this pur fitted with Thomas and Yoon Nelson Model with R2 values 0.97 for both
pose, CTAB proved to be an effective surface modifying surfactant for AC the models. Conclusively, it was a comprehensive study through which
for PFOA removal application. The study achieved an optimal synthesis we synthesized CTAB-impregnated AC and developed of a closed-loop
of CTAB-impregnated AC using Karanja shell for PFOA removal from sustainable process with high selectivity and uptake capacity towards
aqueous media. The performance of K_AC enhanced with the increment the PFOA where all utilized materials, i.e., adsorbent, regenerant and
in biomass/acid ratio and temperature upto a certain extent. The opti PFOA were efficiently recovered.
mized synthesis parameter for PFOA removal was found to be 1:2.5
(acid/biomass ratio), 450 ◦ C (temperature), and loading with 1 % CTAB
solution. The adsorbent showed high porosity, surface area, and uptake Declaration of competing interest
capacity. The K_AC exhibited 974.4 m2/g specific surface area, whereas
CTAB modification reduced the surface area of CTAB_K-AC to 438.5 m2/ The authors declare that they have no known competing financial
g. The quaternary ammonium groups of CTAB molecules enhanced the interests or personal relationships that could have appeared to influence
PFOA uptake capacity of adsorbent upto three-fold. The K_AC and the work reported in this paper.
CTAB_K-AC showed 157 mg/g and 456 mg/g uptake of PFOA in the
batch isotherm experiment. Electrostatic interaction along with hydro Data availability
phobic interaction mainly governs the PFOA adsorptive removal from
the solution. With an increase in pH, the PFOA removal percent of K_AC No data was used for the research described in the article.
decreased (At pH 11, PFOA removal percent decreased from 95 % to 22
%) however, due to the presence of quaternary ammonium groups Acknowledgments
which remains permanently charged even at alkaline conditions, the
performance of CTAB_K-AC was not significantly affected by higher pH The Authors would like to thank the support provided by the Science
values. Both the synthesized adsorbents were found to be quite selective and Engineering Research Board (SERB), Government of India for the
Startup research grant (SRG/2020/000099). We would also like to
17
M.A.N. Shaikh et al. Journal of Water Process Engineering 54 (2023) 103965
thank Sophisticated Analytical Instrument Facility (SAIF) and Environ [21] E. Tyrode, M.W. Rutland, C.D. Bain, Adsorption of CTAB on hydrophilic silica
studied by linear and nonlinear optical spectroscopy, J. Am. Chem. Soc. 130 (51)
mental Science and Engineering Department at IIT Bombay for
(2008), https://doi.org/10.1021/ja805169z.
providing analytical facilities. [22] T. Mahmood, M.T. Saddique, A. Naeem, P. Westerhoff, S. Mustafa, A. Alum,
Comparison of different methods for the point of zero charge determination of NiO,
Appendix A. Supplementary data Ind. Eng. Chem. Res. 50 (17) (2011), https://doi.org/10.1021/ie200271d.
[23] W.A. Arnold, B. Arlene, J. Branyan, A.T. Bruton, C.C. Carignan, G. Cortopassi,
S. Datta, J. DeWitt, A. Doherty, R.U. Halden, H. Harari, E.M. Hartmann, T.
Supplementary data to this article can be found online at https://doi. C. Hrubec, S. Iyer, C.F. Kwiatkowski, J. LaPier, D. Li, L. Li, J.G.M. Ortiz,
org/10.1016/j.jwpe.2023.103965. A. Salamova, T. Schettler, R.P. Seguin, A. Soehl, R. Sutton, L. Xu, G. Zheng,
Quaternary ammonium compounds: a chemical class of emerging concern,
Environ. Sci. Technol. 57 (20) (2023) 7645–7665, https://doi.org/10.1021/acs.
References est.2c08244.
[24] K. Fujimoto, A. Yamawaki-Ogata, Y. Narita, Y. Kotsuchibashi, Fabrication of
[1] S. Deng, Y. Nie, Z. Du, Q. Huang, P. Meng, B. Wang, J. Huang, G. Yu, Enhanced cationic poly(vinyl alcohol) films cross-linked using copolymers containing
adsorption of perfluorooctane sulfonate and perfluorooctanoate by bamboo- quaternary ammonium cations, benzoxaborole, and carboxy groups, ACS Omega 6
derived granular activated carbon, J. Hazard. Mater. 282 (2015), https://doi.org/ (27) (2021), https://doi.org/10.1021/acsomega.1c02013.
10.1016/j.jhazmat.2014.03.045. [25] S. Deng, Q. Zhang, Y. Nie, H. Wei, B. Wang, J. Huang, G. Yu, B. Xing, Sorption
[2] B.O. Fagbayigbo, B.O. Opeolu, O.S. Fatoki, T.A. Akenga, O.S. Olatunji, Removal of mechanisms of perfluorinated compounds on carbon nanotubes, Environ. Pollut.
PFOA and PFOS from aqueous solutions using activated carbon produced from 168 (2012), https://doi.org/10.1016/j.envpol.2012.03.048.
Vitis vinifera leaf litter, Environ. Sci. Pollut. Res. 24 (14) (2017), https://doi.org/ [26] S. Deng, Y.Q. Zheng, F.J. Xu, B. Wang, J. Huang, G. Yu, Highly efficient sorption of
10.1007/s11356-017-8912-x. perfluorooctane sulfonate and perfluorooctanoate on a quaternized cotton
[3] J.M. Steigerwald, J.R. Ray, Adsorption behavior of perfluorooctanesulfonate prepared by atom transfer radical polymerization, Chem. Eng. J. 193–194 (2012),
(PFOS) onto activated spent coffee grounds biochar in synthetic wastewater https://doi.org/10.1016/j.cej.2012.04.005.
effluent, J. Hazard. Mate. Lett. 2 (2021), https://doi.org/10.1016/j. [27] Q. Yu, R. Zhang, S. Deng, J. Huang, G. Yu, Sorption of perfluorooctane sulfonate
hazl.2021.100025. and perfluorooctanoate on activated carbons and resin: kinetic and isotherm study,
[4] H. Wan, R. Mills, K. Qu, J.C. Hower, M.A. Mottaleb, D. Bhattacharyya, Z. Xu, Rapid Water Res. 43 (4) (2009), https://doi.org/10.1016/j.watres.2008.12.001.
removal of PFOA and PFOS via modified industrial solid waste: mechanisms and [28] Di Zhang, Q. Luo, B. Gao, S.Y.D. Chiang, D. Woodward, Q. Huang, Sorption of
influences of water matrices, Chem. Eng. J. 433 (2022), https://doi.org/10.1016/j. perfluorooctanoic acid, perfluorooctane sulfonate and perfluoroheptanoic acid on
cej.2021.133271. granular activated carbon, Chemosphere 144 (2016), https://doi.org/10.1016/j.
[5] M.N. Khan, A. Sarwar, Determination of points of zero charge of natural and chemosphere.2015.10.124.
treated adsorbents, Surf. Rev. Lett. 14 (3) (2007), https://doi.org/10.1142/ [29] Y. Wu, L. Qi, G. Chen, A mechanical investigation of perfluorooctane acid
S0218625X07009517. adsorption by engineered biochar, J. Clean. Prod. 340 (2022), https://doi.org/
[6] N.H. Subramanian, P. Manigandan, A. Wille, G. Radhakrishnan, Determination of 10.1016/j.jclepro.2022.130742.
perfluorooctanoate and perfluorooctanesulfonate inwater matrices by inline matrix [30] Dongqing Zhang, Q. He, M. Wang, W. Zhang, Y. Liang, Sorption of
elimination liquid chromatography with reversed phase separation and suppressed perfluoroalkylated substances (PFASs) onto granular activated carbon and biochar,
conductivity detection, J. Chromatogr. Sci. 49 (8) (2011), https://doi.org/ Environ. Technol. (United Kingdom) 42 (12) (2021), https://doi.org/10.1080/
10.1093/chrsci/49.8.603. 09593330.2019.1680744.
[7] H.C. Thomas, Heterogeneous ion exchange in a flowing system, J. Am. Chem. Soc. [31] Z. Wang, A. Alinezhad, S. Nason, F. Xiao, J.J. Pignatello, Enhancement of per- and
66 (10) (1944), https://doi.org/10.1021/ja01238a017. polyfluoroalkyl substances removal from water by pyrogenic carbons: tailoring
[8] P. Chularueangaksorn, S. Tanaka, S. Fujii, C. Kunacheva, Batch and column carbon surface chemistry and pore properties, Water Res. 229 (2023), 119467,
adsorption of perfluorooctane sulfonate on anion exchange resins and granular https://doi.org/10.1016/J.WATRES.2022.119467.
activated carbon, J. Appl. Polym. Sci. 131 (3) (2014), https://doi.org/10.1002/ [32] T. Gu, R. Zhang, S. Zhang, B. Shi, J. Zhao, Z. Wang, M. Long, G. Wang, T. Qiu,
app.39782. Z. Jiang, Quaternary ammonium engineered polyamide membrane with high
[9] B. Daas, U. Singh, B. Kumar, I. Mukherjee, Water quality status of Indian major positive charge density for efficient Li+/Mg2+separation, J. Membr. Sci. 659
Rivers with reference to agriculture and drinking purposes, in: (2022), https://doi.org/10.1016/j.memsci.2022.120802.
Geopanthropological Environment: An Appraisal, 2016. [33] M. Wang, A.A. Orr, J.M. Jakubowski, K.E. Bird, C.M. Casey, S.E. Hearon,
[10] Z.B. Zhang, X.Y. Liu, D.W. Li, T.T. Gao, Y.Q. Lei, B.G. Wu, J.W. Zhao, Y.K. Wang, P. Tamamis, T.D. Phillips, Enhanced adsorption of per- and polyfluoroalkyl
L. Wei, Effects of the ultrasound-assisted H3PO4 impregnation of sawdust on the substances (PFAS) by edible, nutrient-amended montmorillonite clays, Water Res.
properties of activated carbons produced from it, Xinxing Tan Cailiao/New Carbon 188 (2021), https://doi.org/10.1016/j.watres.2020.116534.
Materials 33 (5) (2018), https://doi.org/10.1016/S1872-5805(18)60349-X. [34] Z. Du, S. Deng, Y. Bei, Q. Huang, B. Wang, J. Huang, G. Yu, Adsorption behavior
[11] H. Liu, J. Zhang, N. Bao, C. Cheng, L. Ren, C. Zhang, Textural properties and and mechanism of perfluorinated compounds on various adsorbents-a review,
surface chemistry of lotus stalk-derived activated carbons prepared using different J. Hazard. Mater. 274 (2014) 443–454, https://doi.org/10.1016/j.
phosphorus oxyacids: adsorption of trimethoprim, J. Hazard. Mater. 235–236 jhazmat.2014.04.038.
(2012), https://doi.org/10.1016/j.jhazmat.2012.08.015. [35] J. Yu, L. Lv, P. Lan, S. Zhang, B. Pan, W. Zhang, Effect of effluent organic matter on
[12] P.V. Thitame, S.R. Shukla, Porosity development of activated carbons prepared the adsorption of perfluorinated compounds onto activated carbon, J. Hazard.
from wild almond shells and coir pith using phosphoric acid, Chem. Eng. Commun. Mater. 225–226 (2012), https://doi.org/10.1016/j.jhazmat.2012.04.073.
203 (6) (2016), https://doi.org/10.1080/00986445.2015.1104503. [36] D.N. Kothawala, S.J. Köhler, A. Östlund, K. Wiberg, L. Ahrens, Influence of
[13] Y. Gao, Q. Yue, B. Gao, A. Li, Insight into activated carbon from different kinds of dissolved organic matter concentration and composition on the removal efficiency
chemical activating agents: a review, Sci. Total Environ. 746 (2020), https://doi. of perfluoroalkyl substances (PFASs) during drinking water treatment, Water Res.
org/10.1016/j.scitotenv.2020.141094. 121 (2017), https://doi.org/10.1016/j.watres.2017.05.047.
[14] W. Hao, E. Björkman, M. Lilliestråle, N. Hedin, Activated carbons for water [37] Z. Du, S. Deng, S. Zhang, B. Wang, J. Huang, Y. Wang, G. Yu, B. Xing, Selective and
treatment prepared by phosphoric acid activation of hydrothermally treated beer high sorption of perfluorooctanesulfonate and perfluorooctanoate by fluorinated
waste, Ind. Eng. Chem. Res. 53 (40) (2014), https://doi.org/10.1021/ie5004569. alkyl chain modified montmorillonite, J. Phys. Chem. C 120 (30) (2016), https://
[15] S.A. Elfeky, S.E. Mahmoud, A.F. Youssef, Applications of CTAB modified magnetic doi.org/10.1021/acs.jpcc.6b04757.
nanoparticles for removal of chromium (VI) from contaminated water, J. Adv. Res. [38] N. Saeidi, F.D. Kopinke, A. Georgi, Understanding the effect of carbon surface
8 (4) (2017), https://doi.org/10.1016/j.jare.2017.06.002. chemistry on adsorption of perfluorinated alkyl substances, Chem. Eng. J. 381
[16] L.C. Wang, X. Jiong Ni, Y.H. Cao, G. Qun Cao, Adsorption behavior of bisphenol A (2020), https://doi.org/10.1016/j.cej.2019.122689.
on CTAB-modified graphite, Appl. Surf. Sci. 428 (2018), https://doi.org/10.1016/ [39] C. Zhao, C.Y. Tang, P. Li, P. Adrian, G. Hu, Perfluorooctane sulfonate removal by
j.apsusc.2017.07.093. nanofiltration membrane-the effect and interaction of magnesium ion / humic
[17] C. Patra, R. Gupta, D. Bedadeep, S. Narayanasamy, Surface treated acid-activated acid, J. Membr. Sci. 503 (2016), https://doi.org/10.1016/j.memsci.2015.12.049.
carbon for adsorption of anionic azo dyes from single and binary adsorptive [40] C. Wu, M.J. Klemes, B. Trang, W.R. Dichtel, D.E. Helbling, Exploring the factors
systems: a detail insight, Environ. Pollut. 266 (2020), https://doi.org/10.1016/j. that influence the adsorption of anionic PFAS on conventional and emerging
envpol.2020.115102. adsorbents in aquatic matrices, Water Res. 182 (2020), https://doi.org/10.1016/j.
[18] K. Weidemaier, H.L. Tavernier, M.D. Fayer, Photoinduced electron transfer on the watres.2020.115950.
surfaces of micelles, J. Phys. Chem. B 101 (45) (1997), https://doi.org/10.1021/ [41] M.A.E. de Franco, C.B. de Carvalho, M.M. Bonetto, R. de P. Soares, L.A. Féris,
jp972245c. Removal of amoxicillin from water by adsorption onto activated carbon in batch
[19] A.B.D. Nandiyanto, R. Oktiani, R. Ragadhita, How to read and interpret ftir process and fixed bed column: kinetics, isotherms, experimental design and
spectroscope of organic material, Indones. J. Sci. Technol. 4 (1) (2019), https:// breakthrough curves modelling, J. Clean. Prod. 161 (2017), https://doi.org/
doi.org/10.17509/ijost.v4i1.15806. 10.1016/j.jclepro.2017.05.197.
[20] M. Shirzad-Siboni, A. Khataee, A. Hassani, S. Karaca, Preparation, characterization [42] R. Radhika, T. Jayalatha, R.K. G., S. Jacob, R. Rajeev, B.K. George, Adsorption
and application of a CTAB-modified nanoclay for the adsorption of an herbicide performance of packed bed column for the removal of perchlorate using modified
from aqueous solutions: kinetic and equilibrium studies, C. R. Chim. 18 (2) (2015), activated carbon, Process Saf. Environ. Prot. 117 (2018), https://doi.org/10.1016/
https://doi.org/10.1016/j.crci.2014.06.004. j.psep.2018.04.026.
[43] Q. Hu, Y. Xie, Z. Zhang, Modification of breakthrough models in a continuous-flow
fixed-bed column: mathematical characteristics of breakthrough curves and rate
18
M.A.N. Shaikh et al. Journal of Water Process Engineering 54 (2023) 103965
profiles, Sep. Purif. Technol. 238 (2020), https://doi.org/10.1016/j. [55] F. Wang, K. Shih, Adsorption of perfluorooctanesulfonate (PFOS) and
seppur.2019.116399. perfluorooctanoate (PFOA) on alumina: Influence of solution pH and cations,
[44] R.C. Buck, J. Franklin, U. Berger, J.M. Conder, I.T. Cousins, P.De Voogt, A. Water Res. 45 (9) (2011), https://doi.org/10.1016/j.watres.2011.03.007.
A. Jensen, K. Kannan, S.A. Mabury, S.P.J. van Leeuwen, Perfluoroalkyl and [56] S. Deng, Q. Yu, J. Huang, G. Yu, Removal of perfluorooctane sulfonate from
polyfluoroalkyl substances in the environment: Terminology, classification, and wastewater by anion exchange resins: Effects of resin properties and solution
origins, Integ. Environ. Assess. Manag. 7 (4) (2011), https://doi.org/10.1002/ chemistry, Water Res. 44 (18) (2010), https://doi.org/10.1016/j.
ieam.258. watres.2010.06.038.
[45] R. Dhore, G.S. Murthy, Per/polyfluoroalkyl substances production, applications [57] B.O. Fagbayigbo, B.O. Opeolu, O.S. Fatoki, Adsorption of perfluorooctanoic acid
and environmental impacts, Biores. Technol. 341 (2021), https://doi.org/ (PFOA) and perfluorooctane sulfonate (PFOS) from water using leaf biomass (Vitis
10.1016/j.biortech.2021.125808. vinifera) in a fixed-bed column study, J. Environ. Hea. Sci. Eng. 18 (1) (2020),
[46] S. Garg, P. Kumar, V. Mishra, R. Guijt, P. Singh, L.F. Dumée, R.S. Sharma, A review https://doi.org/10.1007/s40201-020-00456-1.
on the sources, occurrence and health risks of per-/poly-fluoroalkyl substances [58] M. Hassan, Y. Liu, R. Naidu, J. Du, F. Qi, Adsorption of Perfluorooctane sulfonate
(PFAS) arising from the manufacture and disposal of electric and electronic (PFOS) onto metal oxides modified biochar, Environ. Technol. Innov. 19 (2020),
products. J. Wat. Proc. Eng. 38 (2020) https://doi.org/10.1016/j. https://doi.org/10.1016/j.eti.2020.100816.
jwpe.2020.101683. [59] E.F. Lessa, M.L. Nunes, A.R. Fajardo, Chitosan/waste coffee-grounds composite: An
[47] J. Glüge, M. Scheringer, I.T. Cousins, J.C. Dewitt, G. Goldenman, D. Herzke, efficient and eco-friendly adsorbent for removal of pharmaceutical contaminants
R. Lohmann, C.A. Ng, X. Trier, Z. Wang, An overview of the uses of per- And from water, Carbohyd. Poly. 189 (2018), https://doi.org/10.1016/j.
polyfluoroalkyl substances (PFAS), Environ. Sci.: Proc. Imp. 22 (12) (2020), carbpol.2018.02.018.
https://doi.org/10.1039/d0em00291g. [60] Y. Zhou, M. Xu, D. Huang, L. Xu, M. Yu, Y. Zhu, J. Niu, Modulating hierarchically
[48] G.M. Sinclair, S.M. Long, O.A.H. Jones, What are the effects of PFAS exposure at microporous biochar via molten alkali treatment for efficient adsorption removal
environmentally relevant concentrations? Chemosphere 258 (2020) https://doi. of perfluorinated carboxylic acids from wastewater, Sci. Total Environ. 757 (2021),
org/10.1016/j.chemosphere.2020.127340. https://doi.org/10.1016/j.scitotenv.2020.143719.
[49] T. Stahl, J. Heyn, H. Thiele, J. Hüther, K. Failing, S. Georgii, H. Brunn, Carryover of [61] T.H. Boyer, Y. Fang, A. Ellis, R. Dietz, Y.J. Choi, C.E. Schaefer, C.P. Higgins, T.
perfluorooctanoic acid (PFOA) and perfluorooctane sulfonate (PFOS) from soil to J. Strathmann, Anion exchange resin removal of per- and polyfluoroalkyl
plants, Arch. Environ. Cont. Toxic. 57 (2) (2009), https://doi.org/10.1007/ substances (PFAS) from impacted water: A critical review, Water Res. 200 (2021),
s00244-008-9272-9. https://doi.org/10.1016/j.watres.2021.117244.
[50] S. Lee, S. Kim, J. Park, H.J. Kim, G. Choi, S. Choi, S. Kim, S.Y. Kim, S. Kim, K. Choi, [62] M. Inyang, E.R.V. Dickenson, The use of carbon adsorbents for the removal of
H.B. Moon, Perfluoroalkyl substances (PFASs) in breast milk from Korea: Time- perfluoroalkyl acids from potable reuse systems, Chemosphere (2017) 184,
course trends, influencing factors, and infant exposure, Sci. Total Environ. 612 https://doi.org/10.1016/j.chemosphere.2017.05.161.
(2018), https://doi.org/10.1016/j.scitotenv.2017.08.094. [63] W. Guo, S. Huo, J. Feng, X. Lu, Adsorption of perfluorooctane sulfonate (PFOS) on
[51] U.S. EPA. (2016). Lifetime Health Advisories and health effects support documents corn straw-derived biochar prepared at different pyrolytic temperatures, J. Tai.
for perfluorooctanoic acid and perfluorooctane sulfonate. Https://Www.Epa.Gov/ Inst. Chem. Eng. 78 (2017), https://doi.org/10.1016/j.jtice.2017.06.013.
Sites/Production/Files/2016- 05/Documents/Pfoa_Pfos_Prepub_508.Pdf. [64] M.A. Islam, M. Asif, B.H. Hameed, Pyrolysis kinetics of raw and hydrothermally
[52] L. Ahrens, M. Bundschuh, Fate and effects of poly- and perfluoroalkyl substances in carbonized Karanj (Pongamia pinnata) fruit hulls via thermogravimetric analysis,
the aquatic environment: A review, Environ. Tox. Chem. 33 (9) (2014), https:// Bio. Technol. (2015), https://doi.org/10.1016/j.biortech.2014.11.115.
doi.org/10.1002/etc.2663. [65] M.A. Islam, S. Sabar, A. Benhouria, W.A. Khanday, M. Asif, B.H. Hameed,
[53] R. Naidu, P. Nadebaum, C. Fang, I. Cousins, K. Pennell, J. Conder, C.J. Newell, Nanoporous activated carbon prepared from karanj (Pongamia pinnata) fruit hulls
D. Longpré, S. Warner, N.D. Crosbie, A. Surapaneni, D. Bekele, R. Spiese, for methylene blue adsorption, J. Tai. Inst. Chem. Eng. 74 (2017), https://doi.org/
T. Bradshaw, D. Slee, Y. Liu, F. Qi, M. Mallavarapu, L. Duan, P. Nathanail, Per- and 10.1016/j.jtice.2017.01.016.
poly-fluoroalkyl substances (PFAS): Current status and research needs, Environ. [66] P.T. Scott, L. Pregelj, N. Chen, J.S. Hadler, M.A. Djordjevic, P.M. Gresshoff,
Technol. Innov. 19 (2020), https://doi.org/10.1016/j.eti.2020.100915. Pongamia pinnata: An Untapped Resource for the Biofuels Industry of the Future,
[54] K.U. Goss, The pKa values of PFOA and other highly fluorinated carboxylic acids, BioEnergy Res. 1 (1) (2008), https://doi.org/10.1007/s12155-008-9003-0.
Environ. Sci. Technol. 42 (2) (2008), https://doi.org/10.1021/es702192c. [67] Y. Cai, Q. Wang, B. Zhou, R. Yuan, F. Wang, Z. Chen, H. Chen, A review of
responses of terrestrial organisms to perfluorinated compounds, Sci. Total Environ.
793 (2021), https://doi.org/10.1016/j.scitotenv.2021.148565.
19