You are on page 1of 25

Journal of Environmental Chemical Engineering 9 (2021) 105784

Contents lists available at ScienceDirect

Journal of Environmental Chemical Engineering


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

Remediation of water from per-/poly-fluoroalkyl substances (PFAS) –


Challenges and perspectives
Shafali Garg a, Jingshi Wang b, Pankaj Kumar a, Vandana Mishra a, c, Hassan Arafat d, e,
Radhey Shyam Sharma a, c, *, Ludovic F. Dumée d, e, f, **
a
Bioresources and Environmental Biotechnology Laboratory, Department of Environmental Studies, University of Delhi, Delhi 110007, India
b
Deakin University, Geelong, Institute for Frontier Materials, Waurn Ponds, Victoria, Australia
c
Delhi School of Climate Change & Sustainability, Institute of Eminence, University of Delhi, Delhi 110007, India
d
Khalifa University, Department of Chemical Engineering, Abu Dhabi, United Arab Emirates
e
Center for Membrane and Advanced Water Technology, Khalifa University, Abu Dhabi, United Arab Emirates
f
Research and Innovation Center on CO2 and Hydrogen, Khalifa University, Abu Dhabi, United Arab Emirates

A R T I C L E I N F O A B S T R A C T

Keywords: Per-/poly-fluoroalkyl substances (PFAS) are an emerging class of environmental contaminants used as additives
PFAS to either enhance the thermo-chemical stability of products or alter the properties of surfaces. PFAS are
Treatment train amphiphilic molecules, composed of fluoro-alkyl chains terminated by specialized functional groups such as
Removal
carboxylic, sulphonic acids, phosphates, sulphonamides, and betaines; offering surfactant-like behavior, and
Degradation
Remediation, tandem technologies
making them highly persistent and mobile across all environmental compartments. The treatment of PFAS
contaminated water remains very complex due to the typically low concentrations and because of the complexity
of the wastewater matrix in which PFAS are present. Exposure to trace amounts of PFAS can cause severe health
impacts across all life forms. Trains of treatment or removal techniques must be employed to achieve higher
removal rates. This review assesses existing methods for PFAS capture, concentration, and degradation from
wastewaters. The performance and selectivity, as well as scalability and cost-effectiveness, of these techniques
are critically compared while operating limitations, as well as emerging solutions, are presented to evaluate the
combinatorial benefits of tandem operations for successful PFAS remediation. The discussion is then focused on

Abbreviations: ACs, Activated Carbons; AFFF, Aqueous Film-Forming Foams; AFPO, Ammonium Perfluorooctanoate; ASH, Air Sparged Hydrocyclon; BCF, Bio­
accumulation Factor; BDD, Boron-Doped Diamond; BiFeO3, Bismuth Ferrite; BiOCl, Bismuth Oxychloride; C.D., Current Density; CaCl2, Calcium Chloride; CaF2,
Calcium Fluoride; CBCs, Coconut-Based Carbons; CNTs, Carbon Nanotubes; DCMD, Direct Contact Membrane Distillation; DMAPAA-Q, (N-[3-(Dimethylamino)
Propyl]Acrylamide Methyl Chloride Quaternary;; EO, Electro-Oxidation; FTCAs, Fluorotelomer Carboxylic Acid; FTOH, Fluorotelomer Alcohol; FTS, Fluorotelomer
sulfonate; Ga2O3, Gallium Trioxide; GAC, Granular Activated Carbon; GO, Graphene Oxide; H2O2, Hydrogen Peroxide; HCl, Hydrochloric acid; HF, Hydrogen
fluoride; In2O3, Indium Trioxide; IrO2, Iridium Dioxide; KOC, Organic-Carbon Partition Coefficient; KOW, Octanol-Water Partition Coefficient; LCMS, Liquid Chro­
matography Mass Spectroscopy; LDH, Layered Double Hydroxides; LOQ, Limit of Quantification; MF, Microfiltration; MIPs, Molecularly Imprinted Polymers; MnOx,
Manganese Oxides; MQL, Method Quantification Limit; MWCNTs, Multi-Walled Carbon Nanotubes; Na2SO4, Sodium Sulfate; NaCl, Sodium Chloride; NaOH, Sodium
hydroxide; NF, Nanofiltration; O3, Ozone; OEP, Oxygen Evolution Potential; OH–, Hydroxyl Ion; PAC, Powdered Activated Carbon; PbO2, Lead Oxide; PC, Photo-
Catalysis; PDFA, Perfluorodecanoic Acid; PEC, Photo-Electrocatalytic; PFAAs, Perfluoroalkyl Acids; PFAS, Per-/Poly-Fluoroalkyl Substances; PFBA C3F7COOH,
Perfluorobutanoic Acid; PFBS, Perfluorobutane Sulfonate; PFCAs, Perfluoro Carboxylic Acids; PFCs, Perfluorinated Compounds; PFDA, Perfluorodecanoic acid;
PFDoDA, Perfluorododecanate; PFHpA C6F13COOH, Perfluoroheptanoic Acid; PFHpS, Perfluoroheptanesulfonic Acid; PFHxA C5F11COOH, Perfluorohexanoic Acid;
PFHxS, Perfluorohexane sulfonate; PFNA, Perfluorononanoic Acid; PFOA C7F15COOH, Perfluorooctanoic Acid; PFOS, Perfluorooctane Sulphonic Acid; PFOSA,
Perfluorooctane Sulfonamide; PFPeA C4F9COOH, Perfluoropentanoic Acid; PFPrA C2F5COOH, Perfluoropropionic Acid; PFSAs, Perfluoro Sulphonic Acids; PFTDA
PFTeDA, Perfluorotetradecanoic Acid; PFTrDA, Perfluorotridecanoic Acid; PFUnDA, Perfluoroundecanoic Acid; polyDADMAC, Polydiallyldimethyl Ammonium
Chloride; POPs, Persistent Organic Pollutants; POSA, Perfluorooctane sulfonamide; POSF, Perfluorooctane Sulfonyl Fluoride; rGO, Reduced Graphene Oxide; RO,
Reverse Osmosis; RuO2, Ruthenium oxide; S2O2– 8 , Persulfate; SnO2, Tin dioxide; SWCNTs, Single Walled Carbon Nanotubes; TFA CF3COOH, Trifluoroacetic Acid;
TiO2, Titanium dioxide; UF, Ultrafiltration; U.S. EPA, United States Environmental Protection Agency; WTPs, Water Treatment Plants; WWTPs, Wastewater
Treatment Plants; ZVI, Zero Valent Iron.
* Corresponding author at: Bioresources and Environmental Biotechnology Laboratory, Department of Environmental Studies, University of Delhi, Delhi 110007,
India.
** Corresponding author at: Khalifa University, Department of Chemical Engineering, Abu Dhabi, United Arab Emirates.
E-mail addresses: rssharma@es.du.ac.in (R.S. Sharma), ludovic.dumee@ku.ac.ae (L.F. Dumée).

https://doi.org/10.1016/j.jece.2021.105784
Received 26 April 2021; Received in revised form 15 May 2021; Accepted 29 May 2021
Available online 2 June 2021
2213-3437/© 2021 Elsevier Ltd. All rights reserved.
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

prospects for more cost-effective and scalable PFAS remediation solutions enabling the treatment of dilute and
complex water matrices, required to deal with these extremely persistent pollutants.

1. Introduction wastewater treatment from both processes and operational points of


view. First, general PFAS capture techniques will be discussed in terms
Per- or poly-fluoroalkyl substances (PFAS) represent a class of of their respective operability and overall performance, and limitations.
manmade organo-fluorine compounds composed of fully or partially Second, PFAS degradation techniques will be critically presented with
fluorinated alkyl chains terminated by polar acid groups such as the associated degradation pathways (See SI, Annexure 1). A focus on
carboxylate, sulphonate, or phosphonate [1]. The very stable C-F bond their ability to remediate low PFAS levels in complex water matrices will
provides PFAS with both tremendous thermal and chemical stabilities, be provided to develop a critical discussion on the techniques’ current
unique wettability, and friction properties, supporting their role as ad­ limitations. The discussion will then develop aspects related to the long-
ditives in the manufacturing of flame retardants such as aqueous term efficiency and cost-effectiveness of specific trains of combinatorial
film-forming foams, flame-resistant materials, across the solutions investigated to date to provide holistic PFAS removal strate­
semi-electronic, packaging, anti-adhesive, and metal finishing areas [2]. gies. The need for novel processes and materials solutions will be
The very stable C‒F bond provides PFAS with both tremendous thermal highlighted as the likely primary way forward to develop rapidly rele­
and chemical stabilities, unique wettability, and friction properties, vant approaches for the global remediation of PFAS-contaminated
supporting their role as additives in the manufacturing of flame-resistant wastewaters.
materials, across the semi-electronic, packaging, anti-adhesive, and
metal finishing areas [2]. The high bioaccumulation factor (BCF) of 2. PFAS capture/concentration technologies
PFAS owing to its strong protein-affinity coupled with its tremendous
persistence has resulted in PFAS toxicity in organisms across the biomes This section assesses the capture/concentration technologies of PFAS
globally [3]. with a focus on their efficacy to remediate both model and real effluents,
The fate and transport of PFAS in various environmental compart­ their applicability in Wastewater Treatment Plants (WWTPs), and pro­
ments remain unclear. Although long-chained PFAS (>11) were found vide solutions to overcome the challenges in their selectivity to capture
to adhere firmly to soil particles, shorter precursors tend to enter porous PFAS.
substrates and infiltrate groundwater runoffs, polluting both agricul­
tural soil and freshwater sources, from where it might enter food chain 2.1. Adsorption technologies
[4–6]. Multiple GenX PFAS and PFAS precursors are known to be vol­
atile and tend to escape into the atmosphere [7–9]. The partitioning Adsorption remains a widely used PFAS removal method, offering
coefficients of all PFAS compounds are not extensively known yet, but extremely high sorption rates of 2390 mg PFOS g-1 for IRA67/pH > 10
values of the octanol-water partition coefficient log Kow for Per­ (IERs) [27]. PFAS adsorption onto solid adsorbate is dictated by elec­
fluorooctane sulfonate (PFOS) and Perfluorooctanoic acid (PFOA) were trostatic, hydrogen bonding, and hydrophobic interactions, depending
reported at 4.49 and 4.81, respectively, while the organic-carbon on the water matrix composition and overall adsorbate properties. Ad­
partition coefficient (KOC) of these two compounds was reported at sorbates can be modified with a superficial monomolecular layer to
2.57 and 2.06 respectively, indicating high PFAS mobility both in water impart specific properties facilitating binding of the target PFAS mole­
and soil systems [10]. cules. PFAS, usually have negatively charged terminal groups at
A maximum permissible limit of 400 and 200 ng L–1 for PFOA and environmentally-relevant pHs, enabling its binding with positively
PFOS, respectively, in drinking water has been set by U.S. EPA, however, charged adsorbates [28]. The hydrophobic fluorinated alkyl chains
the industrial water samples often reveal much higher concentrations render PFAS extremely electronegative while the hydrophilic polar head
[11]. Thus, there is an urgent need to develop cost-effective solutions to group typically a carboxylic, sulphonic, or phosphonic acid, enables
manage PFAS containing-effluents, in order to reduce their health haz­ improved adhesion to adsorbates. Owing to these properties, PFAS form
ards towards exposed populations and ecosystems [12]. Treatment micelle, acting as surfactants in solution, which is often exploited for
technologies for PFAS were previously partly reviewed with a specific their successful removal [29]. Key materials used for PFAS adsorption
focus on either the source of the PFAS, such as groundwater [13] and soil include activated carbons (ACs), granular activated carbon (GAC)[13],
[14], or on the type of the treatment technologies used. Advanced oxi­ and powdered activated carbon (PAC); both of which offer high surface
dation/reduction processes (AOPs/ARPs) [15,16], adsorption [17], area, as well as organo-clay materials, minerals, bio-sorbents, molecu­
nanotechnological solutions with photocatalysis, electrochemical larly imprinted polymers (MIPs), nanomaterials such as carbon nano­
degradation, and adsorption [18], have been reviewed separately. tubes (CNTs), and ion-exchange resins (IERs) [30].
Degradation pathways by biological agents [19], for short and
ultra-short PFAS precursors [20], and by chemical and physical methods 2.1.1. Activated carbon
[21] have been summarized in different studies. The treatment train Capture by adsorption with carbon-based materials is the most
approaches or the combinatorial processes have been suggested as a way common practice for the treatment of PFOS or PFOA in contaminated
forward for the future PFAS treatment research for the effective and water. PFAS adsorption efficiency of over 99% was reported with acti­
efficient removal of these compounds from complex water systems [22]. vated carbons [31]. Surface modification plays an important role in
The importance of wastewater-based treatment techniques for PFAS achieving higher removal rates. Out of two types of commercial GACs,
have been assessed and highlighted in multiple previous studies charcoal-based (F400) and coconut-based (CBC), modification with a
[23–26]. However, a critical opinion discussing the different capture strong acid such as HCl increased the positive charge onto the surface of
and degradation technologies available, taking in consideration their the materials offering better adsorption efficiencies (7–8% for F400 and
efficiencies and challenges is missing, to link scalability to technological 6–9% for CBC) for negatively-charged PFAS, while treatment with
implementation towards the development of best remediation guide­ NaOH yielded no such benefits. The addition of heat-activated persulfate
lines and practices for PFAS treatment. (22–25% for F400 and 27–35% for CBC) and H2O2 and Fe (4–8% for
The purpose of this review is not only to guide on factors that affect F400 and 12–13% for CBC) reduced the adsorption of PFAS on both
treatment and review performance but also to develop a high-level types of GACs [32] (Table 1, Entries 1–6). The addition of a sulfur
analysis of the recurring challenges arising from PFAS contaminated polymer- polysulfide to PAC increased the capacity of the adsorption

2
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

Table 1
Case studies on adsorption and coagulation of PFAS.
Entry Treatment Substance Reaction kinetics Removal (%) Time Scale Reference
(qe) (mmol g–1)

Adsorption
1 Granular activated carbon (GAC) (pH= PFOS, PFOA 0.51 n.d 160,120 h Bench [35]
3.08, 3.0) 0.38
2 Powdered activated carbon (PAC) PFOS 0.69 56 4h Bench [35]
(pH= 3.0, 3.01) PFOA 0.79 34
3 GAC (1250 t, 24 filters, Chemviron F- PFOS Not studied 64 ± 11 n.d. Field- DWTP [71]
400 (Density 440 kg m-3; 12 filters), PFOA 45 ± 19
Norit ROW 0.8 (Density 381 kg m-3; 2
filters) and Norit 1240 EN (Density
485 kg m-3, 6 filters)
4 GAC+R.O. PFOS Not studied 86 ± 7 n.d. Field- DWTP [71]
PFOA 89 ± 22
5 GAC (Filtrasorb® 400, SA= 1050 m2 g- PFBA, PFPeA, Total PFAS PFOS (C6), PFTeDA (C13) 140 days Field- WTP [194]
1
) PFHxA, PFHpA, sorption= 25 µg g–1 and FOSA (C8)- ~80%,
PFOA, PFNA, PFDA, PFUnDA
PFUnDA, PFDoDA, (C10), PFDoDA (C11), PFDA
PFTeDA, PFBS, (C9), PFNA (C8)- 70–80%,
PFHxS, PFHxS (C6), PFOA (C7)-
PFOS, FOSA 60–70%, PFPeA (C4) and
PFBA (C3)- 14− 32%
6 Bamboo-derived AC (BAC) and resin PFHxA, PFHpA, PFOA: 2.82 (IRA67), 65–90 48 h Pilot- washing [195]
IRA67 PFOA 1.03 (BAC) wastewater
7 Anion exchange resin (Purolite® PFBA, PFPeA, Total PFAS PFUnDA (C10), 165 days Field- WTP [194]
A600, average diameter= 610 ± 90 PFHxA, PFHpA, sorption= 44 µg g–1 PFDoDA (C11), FOSA (C8),
mm) PFOA, PFNA, PFDA, and PFDA (C9)- 80–90%
PFUnDA, PFDoDA, and PFNA (C8) and PFTeDA
PFTeDA, PFBS, (C13)- 70–80%, PFBA (C3)
PFHxS, and PFPeA (C4)- 10–23%
PFOS, FOSA
8 IER- AI400 (pH=3.10, 3.02) PFOS, PFOA 0.36 n.d. 120 h Bench [35]
3.39
9 IER (Purolite®A520E, A600E, and PFOA, PFOS, PFBA, 0.006–0.093 100% up to a certain bed 70–80 h Bench [37]
A532E) and PFBS 0.009–0.105 volume (A600E,
0.016–0.130 A520E)
50–60 h
(A532E)
10 IER (Purolite® A592 and A860) PFAS including 45 ± 2 >98 (<10 ng L–1) PFSA = 25 Pilot- River [196]
Gen-X 62 ± 4 min (A592), water
69 ± 3 70 min (A860)
106 ± 3 Precursor
98 ± 5 PFAS = 50
84 ± 3 min (A592),
32 ± 3 100 min
(A860)
11 CNTs- 20 wt% graphene composite PFOA 2.415 × 10-3 96.9 4h Bench [41]
electrodes (BET SA=126.8 m2 g-1, pore
diameter=17.1 Ǻ, V=0.6)
12 MWNTs (Anode= Pt foil (20 mm × 40 PFOA 8.2 × 10-3 n.d. 3h Bench [42]
mm, cathode= Ti sheet (20 mm × 40 PFOS 6.5 × 10-3 2h
mm × 0.5 mm, SA= 6 cm2,
thickness=15 µm, V=0.6)
13 Polyaniline nanotubes (PANTs) (outer PFOS, 3.301 80, 40 h Bench [81]
diameter= 200 nm, inner diameter= PFOA 2.657 64
100 nm)
14 3-(trimethoxysily)propyl- PFOS 0.071 n.d. 0.083 Bench [47]
octadecyldimethyl-ammonium PFHxS 0.065
chloride (TPODAC) in a Si-O-Si PFBuS 0.065
framework (SA=140.6 ± 2.5 m2 g–1)
15 Chitosan crosslinked with PFOS 5.5 n.d. 100 h Bench [46]
epichlorohydrin (SA=14.2 m2 g–1, pore
volume= 9.11 cm3 g-1, pH=3.0)
16 Molecularly imprinted polymers PFOA (template), 0.016 90 0.83 Bench [197]
(MIPs) PFOS 0.013
Two functional monomers- 2-
(trifluoromethyl) acrylic acid (TFMAA)
and 4-vinyl pyridine (4-Vpy) and
ethylene glycol dimethacrylate
(EGDMA) as cross-linker in the
presence of azobisisobutyronitrile
(AIBN) (pH=2.0–5.0)
17 GO (SA= 434.6 m2 g–1) PFOA 0.039 × 10-3 60 48 h Bench and [48]
field- airport
2 -3
18 GO modified with FeO (SA=242.4 m PFOA 0.064 × 10 ~90 3–4 h Bench and [48]
g–1) field- airport
(continued on next page)

3
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

Table 1 (continued )
Entry Treatment Substance Reaction kinetics Removal (%) Time Scale Reference
(qe) (mmol g–1)

19 AC-alumina composite- RemB PFOA 0.062 × 10-3 90 3–4 h Bench and [48]
(SA=123.4 m2 g–1) field- airport
20 Organo-clay: PFOA, PFHxS (0.782 – hematite > 0.5 Bench [198]
Hematite (SA=9.9 m2 g–1, pH=5.9), PFOS, 0.927) × 10-3 > kaolinite >
kaolinite (SA= 23.11 m2 g-1, pH= 3.6), PFHxA, PFOS (0.588–0.624) montmorillonite
montmorillonite (SA= 67.52 m2 g-1, PFHxS × 10–3 ≫ PFOA
pH= 7.2) (0.251–2.707) ×
10–3 ≫ PFHxA
(0.108–0.188) ×
10–3
Coagulation
21 GO (adsorbent) + PFOS + Fulvic acid PFOS 0.061 ± 0.003 day–1 45.8 ± 1.8% 6 h (11.4 ± Bench- [199]
(flocculating agent) 0.5 days half- Organismic-
life in carp carp
intestine)
22 Alum (>60 mg L–1) PFOS, PFOA n.d. < 30 1.5 Bench [53]
23 Alum PFOA n.d. 42 1.5 Bench [57]
24 Ferric chloride (3–5 mg L–1) PFOA, PFOS n.d. 0 1.5 Bench [53]
25 FeCl3 PFOA n.d 4 1.5 Bench [57]
26 FeCl3 + FA (Conc.=50 mg L-1, pH=4) PFOA n.d. 47.6 (9.6) 0.5 Bench and [51]
PFOS 94.7 (14.3) Field
27 FeCl3 PFOS/PFOA n.d. 8/6 0.917 Bench [52]
Alum 23/18
PACl 35/27
(Conc.=5.0 mg L–1, pH=7.2) 65/72
Moringa oleifera (Conc.=30.0 mg L–1, 98/94
pH= 7.2)
M. oleifera + PAC (contact time=10
min)
Electro-coagulation
28 Iron electrode (C.D.=37.5 m cm–2, pH PFOA n.d. 93 1.5 Bench [57]
3.77, and mixing speed=180 rpm)
29 Iron electrode + H2O2 (C.D.=37.5 m PFOA n.d. >99 0.67 Bench [57]
cm–2, pH 3.77, and mixing speed=180
rpm)
30 Periodically reversed Al-Zn electrodes PFOS n.d. 100 0.167 Bench, Field- [58]
(Dimensions=60 × 40 × 2 mm, C. PFHxS 95.6 Groundwater
D.=12 V, pH 7.0, and mixing speed= PFBS 87.4
400 rpm (Reaction by-products O, Al,
C, N, Zn, and F elements)
31 Zn electrode+NaCl (SA= 200 cm2, PFOA 5.74 ± 0.08 96.7 0.167 Bench [200]
width 1 mm) (C.D.= 0.18 Wh L-1) PFOS 7.69 ± 0.04 3.6
Mg electrode+NaCl PFDA 11.3
Al electrode+NaCl PFNA 10.6
Fe electrode+NaCl PFHpA (PFDA > PFNA > PFOS >
PFBA PFOA >
PFHpA ≫PFBA)

process of PFOA reducing its concentration in the sample from 5000 µg Sulfonate (PFBS), and Perfluorohexane sulfonate (PFHxS)) except PFOS.
L-1 to 30 µg L-1 within 1 h while stirring. In contrast with PAC, GAC Sulfate as the mobile counter ion, worked more favorably for both types
adsorbed only 200 µg L-1 PFOA due to the lower specific surface area of resins [36] (Table 1, Entries 6–10). Increased hydrophobicity of the
offered by these materials. The use of PAC also significantly reduced the resins was found to play a major role in the adsorption of PFAS. Van der
generation of dust from the adsorbent degradation, which may block the Waal forces also play a role in the adsorption process and on assessment
pores on the membranes. PAC materials were found to generate dust with resins of varied hydrophobicity, the highest hydrophobic resin
concentrations of up to 11,146 µg m–3 for PFOA which was lowered to displayed the highest adsorption capacity, i.e., in the range of
only 159 µg m–3 upon addition of polysulfide, which helps prevent less 52.3–260.5 mg g–1. For scale-up operations, the volume of regeneration
frequent filter clogging, demonstrating the role and impact of surface of effluents requiring incineration after adsorption can be reduced by up
modifiers [33]. to 96.5% by using reverse osmosis technologies. Thin poly(amide)
membrane from GE Osmonics designed for brackish water operated at
2.1.2. Ion-exchange resins 20–22 bar pressure, gave an average water recovery rate of 65%. This
Ion exchange resins (IER) show higher adsorption capacity was followed by vacuum evaporation between 35 and 38 ◦ C. This
(210–2575 mg PFOS g–1 and 1206 mg PFOA g–1) as compared with non- reduced the overall operational costs [37]. Cationic polymer-based
ionic resins (37–41 mg PFOS g–1 and 38–46 mg PFOA g–1) [34,35]. hydrogels such as DMAPAA-Q (N-[3-(dimethylamino)propyl]acryl­
Different functional groups, polymer matrices of the resins, and porosity amide, methyl chloride quaternary) were also efficient (80–100%
levels dictate the suitability of resins for adsorbing PFAS [17]. Poly removal) in supporting the adsorption of legacy as well as Gen-X PFAS
(styrene) resins with type I quaternary ammonium functional groups and regeneration for up to 10 cycles resulted in no loss of efficiency [38].
showed higher adsorption capacities than poly(acrylic) resins exhibiting
similar functional groups for most PFAS (Perfluorobutanoic Acid 2.1.3. Carbon nanotubes
(PFBA), Perfluorohexanoic Acid (PFHxA), PFOA, Perfluorobutane The ability to alter the design of sorbents at the nanoscale has also led

4
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

to the development of more efficient adsorption materials offering larger 2.2. Coagulation
surface areas and enthalpies of adsorption, consistently limiting hys­
teresis of adsorption and greater cyclic performance recovery rates [18]. Coagulation is a conventional treatment method, used widely in
The presence of nano-sized bubbles increases the surface area onto WWTPs. The search for selecting coagulating agents for PFAS capture is
which the PFAS adsorbed allowing for more hydrophobic interactions directed around either primary coagulation strategies, whereby chem­
and subsequently, more adsorption capacity [39]. In a study conducted icals directly lead to PFAS coagulation, or secondary coagulation stra­
to understand the importance of surface roughness and nano-cavities, tegies, whereby PFAS is adsorbed onto discrete powder-like adsorbents,
under vacuum-degassing, 74–79% of the initial PFOS removal was which are thereafter coagulated. In addition, electro-coagulation, floc­
compromised with adsorption carried out with CNTs and graphite [40]. culation, and sedimentation have also been demonstrated as potential
Single-walled CNTs (SWCNTs) showed better sorption capacity (556 mg treatment routes to selectively capture PFAS. Charge neutralization
PFOS g–1, and 492 mg PFOA g–1) [41] than Multi-walled CNTs (MWNTs) resulting in separation via. settlement, floatation, or precipitation is
(2.69 mg PFOA g–1) [42]. This higher sorption capacity is facilitated by basically the working principle behind the four mechanisms discussed in
hydrophobic interactions, therefore, adsorbents functionalized with this section [51].
hydrophilic hydroxyl or carboxyl groups are added for aiding such
adsorption processes [43]. However, the electrostatic repulsion between 2.2.1. Conventional coagulants
negatively charged CNTs and anionic PFAS could act against the Synthetic (alum, ferric chloride, and polyaluminum chloride) and
adsorption process. Generally, with a decrease in carbon chain length natural (Moringa oleifera) coagulants were tested for their performance.
(<6C), adsorption decreases [44]. 28% amine-functionalized organic Conventional coagulants such as polyaluminum chloride or ferric chlo­
frameworks carried out >90% PFAS adsorption [45]. Thus, surface area, ride effectively neutralized between 0.1 and 1 mg L–1 of PFOS or PFOA
functional groups, and hydrophobicity are the most important de­ within 30 min of mixing in synthetic wastewater. These strategies are,
terminants of adsorbent performance. These studies are further sum­ however, often ineffective with industrial level concentrations of PFAS
marized in Table 1, Entries 11–13. (above 10 mg L–1). PFOA or PFOS removal for the concentration of ~60
By the application of an electric field, the adsorption capacity of the mg L–1 with alum only reached 20% efficiency in laboratory-scale con­
nanotubes, i.e., electrosorption to MWNTs sheets, increased. Following ditions. At 5 mg L–1, the efficiency of PFOS/PFOA removal was found to
pseudo-second-order kinetics, an increase of 41–60 folds for PFOS vary for different coagulant materials from polyaluminium chloride
(3.251 mg g–1) and PFOA (3.395 mg g–1) was observed when using (PaC) (35/27%) to alum (23/18%) and ferric chloride (8/6%).
continuous sheets supplied with electric field compared to discrete M. oleifera @[C0] = 30 mg L–1, coagulated 65% of PFOS and 72% of
MWNT powders (PFOS: 1.6 mg g–1 and PFOA: 0.911 mg g–1). This was PFOA, and conjugated with PaC in 10 min, the removal efficiencies
attributed to the improved electrostatic attraction between positively increased to 98% PFOS and 94% PFOA [52]. pH, non-organic matter,
charged electrodes and negatively charged PFOS– and PFO– [42]. and stirring time are deciding operating conditions for coagulants per­
formance, and further works on more complex and realistic water
2.1.4. Other adsorbents matrices are required [51,53].
Bio-sorbents such as chitosan, biochar, husks of rice and wheat, and Commercial coagulants such as cationic polymers polydiallyl
quaternized cellulose, extracted from cotton fibers, offer a promising dimethyl ammonium chloride (polyDADMAC) and polyamine, a co-
waste-to-resource approach that is both scalable and cost-effective, polymer of epichlorohydrin and dimethylamine also increase the sorp­
aligning well with circular economy strategies. Chitosan provided tion of PFAS to Aqueous film-forming foams (AFFF)-impacted grounds
good adsorption capacities of 645–2710 mg g− 1 for PFOS [46]. In a by a factor of 2.0–6.1 [54]. Conventional coagulants such as alum or
combined process, MIPs prepared with chitosan crosslinked with ferric chloride were ineffective in treating PFAS from WWTPs [55].
epichlorohydrin was used to permanently confine micelle arrays, Studies found that flocculation at WWTPs increased PFAS concentration
demonstrating the role of electrostatic interactions for PFAS capture, observed over a three days period. PFHxS increased by 300% by day 3,
with adsorption capacities of 105 mg g–1 for PFOS, 95 mg g–1 for PFHxS, Perfluoroheptanoic acid (PFHpA) by 2% by day 2, PFBA, Per­
and 65 mg g–1 perfluorobutane sulfonate (PFBS), reached equilibria fluoropentanoic acid (PFPeA), and Perfluorodecanoic acid (PFDA)
within 5 min. In terms of stability, the MIPs showed over 90% recovery decreased by 5–25% by day 2. This could be attributed to the effects of
in subsequent adsorption trials for up to 5 cycles with adsorption not filtration and precursor degradation [56]. These studies are summarized
going below 85% for all three types of PFAS [47]. in Table 1, Entries 21–27.
PFOS are known to adsorb well onto carbon materials, such as gra­
phene oxide (GO) (580 ± 205 mg g–1 or mg L–1 measured in terms of 2.2.2. Electrocoagulation
Freundlich affinity coefficient) [29,48]. Graphene oxide (GO) is a Electrocoagulation was found to facilitate the coagulation of PFOS or
unique monomolecular layer of graphene which can be functionalized PFOA rapidly by promoting sol-gel formation by neutralizing the
with various functional groups. The use of Mg2+ and Ca2+ cations in­ negative charge of PFAS by means of a sacrificial anode of iron/
creases the strength of binding between GO and PFOS by the bridging aluminum. These are means to generate hydrous ferric oxide/aluminum
effect [49]. Aluminium- and iron-based minerals such as alumina, he­ hydroxides, which neutralizes the electrostatic charge across the ad­
matite, goethite are able to bind to POPs by hydrophobic and π-π in­ sorbates, facilitating the coagulation. However, H+ generation at the
teractions [50]. The use of a commercial material, RemBind™ (RemB), cathode caused the floatation of the adsorbents, eventually requiring
an activated carbon/clay/alumina-based adsorbent, offered over 90% of regeneration or replacement. The iron electrode charged at a current
PFOS removal, which comparable to iron oxide modified GO composites density of 37.5 mA cm–2 was used to coagulate 93% of PFOA within 90
(FeG), but much greater removal than pristine GO (~60%) tested under min @3.77 initial pH and 180 rpm of mixing speed. The addition of
the same conditions. The pH and ionic strength of the solution also had H2O2 to this reaction resulted in the oxidative transformation of ferrous
limited impact on the behavior of most of these adsorbents but for GO, to ferric ions, which increased the removal to >99% within 40 min [57]
and electrostatic interactions were not found to affect extensively the (Table 1, Entries 28–31). Using the same electrode over many cycles,
binding capacity of the adsorbents. Hence, Ligand exchange mecha­ generates a passivation effect, which can be negated by periodic
nisms driven by Fe- and Al-minerals of FeG and RemB as well as hy­ reversing of electrode polarity. The use of switching polarity was
drophobic interactions were found to drive these alumina-minerals demonstrated for Zn-Al electrodes with reversing cycles every 10 min @
interactions and the functionalization of the GO with metal oxides hel­ 12.0 V, 7.0 pH, stirring speed of 400 rpm, and initial concentration of 1
ped to enhance the adsorption capacity of GO [48]. These studies are mg L–1. The removal of various PFAS compounds was demonstrated
summarized in Table 1, Entries 14–20. using this set-up, with 87.4% removal of PFBS, 95.6% of PFHxS, and

5
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

Table 2
Case studies using high-pressure membranes for PFAS concentration.
Entry Process Membrane code Compounds Treatment conditions Feed Efficiency (%) Scale Reference
concentration
(mg.L-1)

1 RO BW30, ESPA3, PFOS Cross-flow 0.5–1600 >99 Bench [65]


LFC3, and SG semiconductor
wastewater
pH 4
2 RO BW30, LFC1, PFOS Crossflow 10 >99 Bench [68]
LFC3, ESPA3, and synthetic
SG semiconductor
wastewater
pH 4
3 RO BW30, XLE, and PFHxA Crossflow 100 BW30, XLE, Bench [69]
SW30XLE synthetic industrial SW30XLE >99 at
effluents pH 7.1
pH 3.5 and 7.1
4 UF-RO n.d. PFOA and PFOS Crossflow 15–26 > 99 Bench [71]
Surface water quality 61–86 and Field
5 NF NF90, NF270, DK PFOS Crossflow 10 90–99 Bench [68]
synthetic
semiconductor
wastewater
pH 4
6 NF NF270, NF200, PFPnA, PFBS, PFHxA, PFHpA, PFHxS, Crossflow 1 For all PFAs Bench [74]
DK, DL PFOA, 6:2 FtS, PFNA, PFOS, FOSA, synthetic tertiary- >95
PFDA, PFUnA, PFDS, PFDoA, PFTeA treated wastewater 5> pH > 6;
pH 3 – 10 60–80
at pH 3
7 NF NF27, NTR-7450 PFHxA Cross-flow 100 70–99 Bench [76]
Synthetic water
pH 3.3, 7, and 10
8 NF NF270 PFOA Cross-flow 1 85–90 Bench [80]
Synthetic water
pH 7
9 NF NF90, NF270 PFHxA Crossflow 100 NF90 Bench [69]
synthetic industrial >97 at pH 3.5
effluents >99 at pH 7.1
pH 3.5 and 7.1 NF270
>80 at pH 3.5
>95 at pH 7.1

Notes: XLE: extra-low filtration, LFC: low fouling composite, SG: Singapore membrane, ESPA: energy saving polyamide, NF: nanofiltration, DL: dual-layer, ACM:
Actively Cool Membrane.

100% of PFOS within 10 min of operation from real groundwater [58]. rejection of PFPeA was found to drop from 85% to 58% when the feed
temperature was increased from 50 to 70 ◦ C, which was correlated with
2.3. Membrane filtration an increase in the evaporation rate of the feed solution.
However, high-pressure membrane processes, including nano­
Membrane filtration is a process whereby a semipermeable or porous filtration (NF) and reverse osmosis (RO) have been widely applied at
membrane is used to selectively separate a solvent from solutes [59]. bench and pilot scale for sequestering PFAS from wastewater. Few full-
Membrane filtration allows the opportunity to concentrate PFAS, which scale trials were performed to increase the rate of evaporation of the feed
offers the potential of PFAS reuse in industrial processes. The membrane solution over the past two decades since the pore size/molecular weight
separation process is driven by a pressure, electrical or temperature cut-off (MWCO) of these membranes is sufficient for the PFAS retention
gradient, leading to specific solvent permeation through the membrane [63,64]. The applications of RO and NF processes for PFAS treatment are
and solutes’ rejection capabilities. The solutes rejection mechanisms in summarized in Table 2.
membranes may, in fact, occur due to a combination of sieving and
charge attraction/repulsion mechanisms. The former mechanism is 2.3.1. Application of RO processes for PFAS treatment
typically applicable to particulate matter while the latter dominates the The first study on concentrating PFAS using four commercial RO
rejection of ions and small organic molecules [60]. membranes showed efficacy to remove >99% of PFOS (538 g mol− 1)
Although low-pressure membranes, such as microfiltration (MF) and across a concentration range of 0.5–1600 mg L− 1 from semiconductor
ultrafiltration (UF) membranes with large pore size distributions wastewater [65]. Although the rejection performance was enhanced
centered around 100–400 and ~10 nm, respectively, are unsuitable for with increasing concentration of PFOS in the feed. The PFOS rejection by
rejecting low molecular weight compounds such as PFAS [61], they commercially available membranes such as brackish water (BW30), low
remain useful during pre-treatment processes for concentrating PFAS fouling composite (LFC3), and Singapore membrane (SG) all reached
and to mitigate fouling at subsequent treatment processes. The appli­ >99.9% PFOS concentration, while a less significant enhancement of the
cation of MF membranes in the direct contact membrane distillation PFOS rejection was observed for the Energy-Saving Poly(amide)
process (DCMD) was reported as a pre-treatment for concentrating (ESPA3) membranes. The PFOS rejection performance was reconciled
PFPeA compounds from synthetic wastewater [62]. Since the DCMD with NaCl solutions’ rejection for these RO membranes, suggesting size
processes are driven by vapor pressure differences, the filtration trials exclusion as the controlling factor for rejecting PFOS and that the tighter
were conducted at different temperatures. At a feed temperature of RO membranes can effectively reject more PFOS molecules, which are
50 ◦ C, the rejection of PFPeA reached the highest value of 85%. The governed by solubility and diffusivity of the solute and solvent. Similar

6
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

Fig. 1. Comparison of technology performance


in terms of removal efficiency achieved to
capture or degrade PFAS, with the treatment
trains or hybrid systems both in the lab- and at
field-scale. Data recorded in Tables 1–6.
Missing bars represent lack of data in the
respective research articles. From the box-plots,
it is observable that there is consistently high
removal shown by treatment train approaches,
affirming their usefulness for upscalability and
achieving high efficiency. Notes: AC: Activated
carbon, IR: Ion-exchange resin, NA: Nano­
adsorbent, OT: Others, CC: Conventional coag­
ulant, EC: Electrocoagulant, RO: Reverse
Osmosis, FL: Filtration, NF: Nanofiltration, OR: Oxidation/reduction, SC: Sonochemical, PT: Photo/Thermal, BT: Biological treatment, ET: Enzymatic treatment, UV:
Ultraviolet, AF: Air-fractionation, OZ: Ozonation, RP: Reduction processes, AR: Aeration, EO: Electro-oxidation, PC: Photocatalysis, UF: Ultrafiltration, TP: Water
treatment plant, OP: Oxidation processes, ZI: Zero-valent iron, BC: Biological activated carbon.

mechanism has also been reported for pharmaceuticals [66]. Theoreti­ Ultraviolet-advanced oxidation processes (RO-UV/AOPs) showing
cally, negatively charged membranes shall reject negatively charged complete mineralization of the PFAA substances, though the content
organic compounds more effectively owing to electrostatic repulsions within the post-treatment RO concentrates [72] was not explicitly
[67]. The ESPA3 exhibited a more negative streaming potential charge mentioned in the study, but it highlights the usefulness of hybrid tech­
at pH 4, while both BW30 and LFC3 were similar to each other and nologies (Table 6).
exhibited a less negative absolute charge at the same pH. The results
suggest that the impact of electrostatic repulsions on PFOS rejection is 2.3.2. Application of NF processes for PFAS treatment
not as significant as size exclusion mechanisms when using RO mem­ The retention performance of NF membranes, semipermeable
branes for PFAS removal. Using RO membranes for treating PFOS mol­ membranes with looser microstructures compared to RO membranes, is
ecules at lower concentrations (10 mg L− 1) [68] and for treating PFAS also governed by size exclusion of compounds. NF270 membranes were
with shorter-chain such as PFHxA also confirmed similar behaviors [69] used to reject nine PFAS, including PFBA, PFPeA, PFHxA, PFOA, per­
(Table 2, Entries 1–5). fluorononanoic acid (PFNA), PFDA, PFBS, PFHxS, and PFOS [73].
Another series of studies evaluated the normalized initial RO fluxes Retention by the NF270 membranes was achieved for PFBS and PFHxS,
to differentiate the effect of initial flux from that of the membrane ma­ exhibiting a molecular weight of 300 and 314 g mol–1, for which the
terials properties showed superior removal efficiencies of >99% for rejection dropped to 93% and 97%, respectively [73]. For lower mo­
PFOS with SG, LFC1, LFC3, BW30, and ESPA3 RO membranes [68]. lecular weight PFAS compounds, only 70% of PFPeA (264 g mol–1) was
From an operation point of view, the filtration conducted at higher rejected by NF270 in a separate study [74]. Based on the size exclusion
applied pressures led to a severe degree of fouling, hindering the passage mechanisms, the MWCO of NF membranes, defined for rejection at or
of both water and PFOS molecules, and subsequently rejecting more above 80%, therefore appears to lay around 300 g mol–1.
PFOS molecules over time. Furthermore, LFC1, LFC3, and ESPA3 with a The performance of NF membranes to reject PFAS compounds may
surface roughness of 135.8 ± 12.8, 108.4 ± 12.4, and 181.9 ± 26.1 nm, be enhanced further by controlling electrostatic repulsions between the
respectively were found to be more easily and rapidly fouled than SG contaminants and the membrane surface [67,75]. The impact of the pH
and BW30 with a surface roughness of 17.3 ± 3.0 and 68.3 ± 12.5 nm, on the retention of anionic Perfluorooctane sulfonamide (POSA) [74],
respectively. The results imply that RO membranes with rougher sur­ and PFHxA molecules [69,76] by negatively charged NF membranes
faces were more susceptible to PFOS adsorption than those with smooth were reported for several bench-scale studies. In these studies, PFOA,
surfaces by providing more surface areas, in line with silica colloidal POSA, and PFHxA dissociated into negative ions at the pH range of 3–10.
matter fouling, a relevant comparison point to PFAS given molecular At lower pH range, for the NF270 membranes, exhibiting an isoelectric
weights and native compounds charge [70]. point at around pH 4, the surface charge of the membranes became less
RO membranes were also often applied in combination with other negative to neutral, which led to the reduction of solute rejection and
techniques, improved PFAS removal process efficiencies across several hence adsorption of PFAS compounds. As the solution pH increased
pilot plants and a drinking water treatment plant (DWTP). UF pre- toward the alkaline range (pH 6–10), the membrane surface with
treatment prior to RO units was tested to remove PFOS and PFOA dissociated carboxylic functional groups became strongly negatively
molecules. The UF-RO treatment trains were able to reject >99% of charged, which increased PFAS rejection from electrostatic repulsion
PFOS and PFOA owing to the RO barriers [71]. In both pilot and between anionic PFAS molecules and the membrane surface [69, 74, 76,
full-scale operations similarly, tertiary effluents containing per­ 77]. Furthermore, the membrane materials can be engineered to
fluoroalkyl acids (PFAAs) were treated by RO followed with enhance the fouling resistance against PFAS compounds. The BW30

Fig. 2. Comparison of technology performance


in terms of time taken to capture or degrade
PFAS, with the treatment trains or hybrid sys­
tems both in the lab- and at field-scale. Data
recorded in Tables 1–6. Missing bars represent
lack of data in the respective research articles.
From the box-plots, it is observable that some of
the adsorption, aeration and biological tech­
nologies show extremely high removal times
and with the exception of field studies with
water treatment plants (WTPs), most of the
treatment train studies bring down the removal
time significantly.

7
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

membranes with amine enriched surfaces would be more negatively add to the cost of the process and make electrocoagulation an energy-
charged under alkaline conditions, thus attributing to a stronger repul­ intensive process [84]. Sedimentation, however, has no effect on the
sive interaction with the dissociated PFAS ions and subsequently resist concentration of PFAS in WWTPs [85,86]. Hence, ineffective for treat­
PFAS adsorption [78] (Table 2, Entries 6–9). ment of PFAS.
Membrane fouling caused by the adsorption of PFAS is also Although novel adsorbent materials and especially nanomaterials
inevitable during NF filtration processes, which subsequently in­ are promising adsorbent materials, limitations of their scalability, and
fluences rejection and permeation. A significant 35% permeation potentially toxic side effects from the bare adsorbents if released in
reduction was observed during the filtration experiments using 9 decontaminated waters, represent a major hindrance to their develop­
PFAS compounds [73]. Smaller PFAS molecules, such as PFHxS and ment. Furthermore, adsorption and coagulation are often coupled with
PFBS, remained undetected at the end of filtration tests in the other concentration techniques such as membrane filtration to maximize
permeate solution, indicating that the fouling layer hindered the removal.
passage of PFAS molecules and likely concentrated them in the cake NF membranes with looser microstructures may be more energy-
layer onto the membrane surface. Such rejection enhancements were efficient towards PFAS rejection, compared to RO membranes. How­
caused by the entrapment of PFAS within the selective poly(amide) ever, the MWCO of NF membranes is larger than that of RO limiting
layer of the membrane material [68,79]. The cake layer formation, their applicability to reject low molecular weight PFAS compounds.
however, led to a decline in the PFAS rejection capability over time, i. Furthermore, unlike RO membranes of which rejection mechanisms
e., membrane fouling. are governed predominantly by size exclusion, the performance of NF
The selective layer of NF membranes could also be tailored to feature membranes are under significant influence of the chemistry of PFAS
nanoscale surface pores across the material surface to minimize inor­ contaminated feed such as pH, ionic strength, and the content of dis­
ganic fouling [80]. The water flux of such NF membranes declined by solved organic matters [68,74]. Although the retention performance of
18% at the end of 20 h scaling tests with a mixture of the PFOA and salt RO at a large scale remained at the same level as bench-scale studies,
solutions containing sodium chloride (NaCl), calcium chloride (CaCl2), as a non-destructive method for PFAS treatment, disposal of the PFAS
and sodium sulfate (Na2SO4). The virgin NF270 experienced a much concentrated brine has always been an issue [87]. At the pilot scale,
higher flux decline by 28% compared to the modified membranes the retention performance of the NF membrane, especially in combi­
(Table 2). The operating parameters in terms of removal and time taken nation with other treatment processes such as electro-oxidation (EO)
are compared in Figs. 1 and 2. and photocatalysis, cannot match the hybrid approaches conducted at
the bench scale (>93–99%) [73,88,89]. These challenges and research
2.4. Limitations and directions for PFAS capture technologies needs are summarized in Table 7. Most of the current studies are done
on a laboratory scale and lack the field application studies. Systematic
Adsorption is one of the most commonly applied techniques for PFAS studies to develop ultra-selective PFAS capture pathways or materials
capture since performing at low operating costs, and offering high are therefore required.
removal efficiencies up to 2710 mg PFOS g–1 for chitosan [46],
2390 mg PFOS g–1 for IERs [27], and 1651 mg PFOS g–1 for CNTs [81], 3. PFAS degradation technologies
respectively. The simplicity of the design and both the modulation and
regeneration potential (especially MIPs and CNTs) of the process also This section focuses on recent advances in degradation, decomposi­
favor the implementation of these materials. However, low adsorption tion or mineralization of PFAS, mostly achieved by oxidation/reduction
and coagulation rates may detrimentally extend the operating times, up of PFAS, covered as the physicochemical processes of catalysis, wet
to 240 h for GAC, 168 h for IERs, 10 h for CNTs, MIPs, and bio-sorbents, oxidation-reduction, sonochemical, photo/thermal as well as the
thus requiring thermal regeneration of adsorbent such as GAC with biochemical processes, which involve biologically derived catalysts or
temperatures above 600 ◦ C, creates waste residuals to dispose off biological systems leading to PFAS degradation. The mechanism of
exhausted carbon, process optimization is necessary in terms of pH, degradation of PFOA and the various reaction intermediaries have also
temperature, and contact time [17]. Different absorption chemistry is been discussed.
observed for varied conditions of pH. Sorbents, such as activated carbon
materials, are expensive and regeneration is not always feasible [82]. 3.1. Catalysis
The nature of the performance achieved ranged widely based on the
different types and expected interaction mechanisms between the sur­ The catalytic degradation of PFOA compounds has been studied
face of the adsorbents and coagulants and the PFAS solutes. Initial using both photocatalytic and electrocatalytic approaches. The strong
concentration of the PFAS solute also plays a critical role in determining bond dissociation energy of the C‒F bonds (<544 kJ mol–1) represents a
the removal efficiency. A direct comparison of the performance of the key challenge to mineralize PFOA efficiently. This section provides links
two main types of adsorbates is not entirely valid but specific isotherms to the properties of materials and kinetics of reactions and the primary
of adsorption as well as reaction kinetics should be standard methods to catalytic degradation pathways of PFOA compounds are discussed. Po­
evaluate interplays and link operating conditions to capture outputs. tential new materials that can be used in this application as well as
Resins and bio-sorbents were found to be much better adsorbents scalable routes towards energy-efficient PFOA treatment will be
compared to AC, which are nevertheless the most reported materials in introduced.
literature due to their low cost and accessibility. PAC has 100 times more
surface area than GAC but they are not reusable whereas GAC, even 3.1.1. Photocatalysis
though more expensive can undergo regeneration, a couple of times Photolysis was initially considered as a potential option since used
[83]. However, the regeneration costs of ACs hinder their application in extensively for the degradation of a variety of toxic agrochemicals,
daily-use processes and new regeneration strategies, specific to PFAS pharmaceuticals, and other industrial compounds [60,90]. However,
contaminations, must be developed. direct photolysis of different PFOA compounds in water with UV-C
Conventional coagulation, flocculation, and sedimentation tech­ (254 nm) even up to 300 min of exposure showed very low degrada­
niques are not able to achieve more than 20% removal with real tion [91]. The photon energy of UV-C of 254 nm (471 kJ mol–1) remains
wastewater. Pollutants are transferred to another phase without being insufficient to split the C-F bond [21]. Although shorter wavelength,
degraded, regeneration is neither practiced nor feasible with floccula­ below 200 nm, would yield photon energies of >600 kJ mol–1, the
tion. Spent coagulants are often removed after usage and chemicals operating costs involved with such UV sources, and yet low efficacy, are
added each time while changing wastewater load. Sacrificial electrodes typically considered prohibitive for scale-up applications. Thus, the

8
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

Table 3
Photo-, electro- and photoelectro- catalytic performance towards the degradation of PFOA materials. All kinetic degradation rates reported were normalized for
comparison purposes. Most of the tests done to date reported data on perfluorooctanoic acid compounds [55]. However, [136,201] can be referred to for better
understanding.
Entry PFOA Materials details and Catalytic operation Major by-products Kinetic degradation Comments on performance Reference
properties conditions besides H+ and F- rate (min-1)

Photocatalysis
1 PFOA Titanium dioxide 254 nm @ 400 W and 25 oC; PFHpA, PFHxA, F- 2 and 5 Limited impact of the TNTs on [95]
nanotubes (TNTs), 9 nm 50 mg L− 1 PFOA ions the rate
inner diameter dispersed
in solution
2 PFOA Cu and Fe doped 254 nm @ 400 W and 25 oC, PFHpA, PFHeA, 1 and 3.1 Increased rate with metal [202]
titanium dioxide thin 0.5 g L− 1 catalyst, varied pH; PFPeA, PFBA, doping but no significant
films dispersed in 50 mg L− 1 PFOA PFPrA, TFA changes
solution
3 PFOA Titanium dioxide Pb 254 nm @ 400 W and 25 oC; PFHpA with a 2 and 85 Pd doping more efficient than [99]
doped thin films catalyst concentration from reduced number previous metal doping
dispersed in solution 0.1 to 3 g L− 1; 50 mg L− 1 of the CF2 units;
PFOA PFHpA, PFHeA,
PFPeA, PFBA,
PFPrA, and TFA
4 PFOA TiO2-rGO 200 and 600 nm wavelength Not assessed but 0.2 and 1.9 TiO2 alone was worse than [93]
@ 150 W, catalyst theorized photolytic reference. The level
concentration from 0.05 to of TiO2 doping on the graphene
0.5 g L− 1. 50 mg L− 1 PFOA was not evaluated
5 PFOA Pt, Pd, Ag modified TiO2 125 W at 365 nm, 5.3 Not assessed 20 and 121 Doping levels were not assessed. [96]
mWcm-2; catalyst 0.5 g L− 1; Unclear if Pt had the same
PFOA 60 mg L− 1 atomic level of doping as others
6 PFOA In2O3 nanostructures 15 W @ 254 nm; catalyst Not assessed 5 and 1323; best The geometry and size of the [92]
concentration 0. 5 g L− 1; performance for the particles were varied by
PFOA concentration of porous spheres adjusting the co-solvent addition
30 mg L− 1 compared to dense during hydrothermal synthesis.
cubes Variations attributed to the level
of O vacancies
7 PFOA In2O3 and Titanium 23 W @ 254 nm; PFOA 100 Not assessed 4 and 25 Tests were done with model [203]
oxide μmol L− 1 mixed with 0.05 g secondary effluents showed
L− 1 of much less degradation
photocatalyst capability; reduction by ~ 6
times
8 PFOA ZnO/rGO nanocomposite 1 g L− 1 of ZnO/rGO; PFOA PFPrA, PFBA, 33–100 Strong adsorption component [204]
concentration 100 mg L− 1; PFHpA, (LCMS (up to 35% of initial PFOA
irradiation density 1464.6 data) concentration after 1 h) not
Wm− 2 deducted; (values corrected I
rate column)
9 PFOA Pb-BiFeO3/rGO 5 W irradiation @ 254 nm; Not assessed 1–13 The density of nanoparticles on [205]
10–100 mg L–1 PFOA; the rGO was not evaluated. Best
50–400 mg L–1 performance achieved for mid-
range rGO doping amount
(0.5 %)
10 PFOA TiO2–MWCNT 300 W @ 365 nm; PFOA at PFPA, PFBA, 3–17 Very large loadings of catalysts. [206]
30 mg L–1 and 1.6 g L–1 of PFPeA, PFHxA, Limited evaluation of titania
TiO–2MWCNT catalysts PFHpA (LCMS) loadings
11 PFOA Bismuth oxychloride 254 nm @ 10 W; 1 g L− 1 of Not assessed 21–78 The level of vacancies and zeta [186]
nanosheets catalyst; 20 μmol L− 1 potential directly impact
performance. Radicals
quenchers affect the stability of
the reaction and reduction
kinetics
12 PFOA surface defective BiOCl 300 W with broad sunlight- Not assessed 4–17 Catalyst is unstable and [185]
like spectrum with an dissolves over time
intensity of 105 mW cm-2;
PFOA 50 μmol L− 1; 1 g L− 1 of
catalyst
13 PFOA Titania 23 W, emitting 254 nm; TFA, PFPrA, 5–50 Radicals quenchers affect the [207]
6 mW. cm− 2; PFBA, PFPeA, rate of degradation dramatically
PFHxA, PFHpA,
PFOA
14 PFOA Phosphotungstic acid PFOA 5–40 mg⋅L− 1; PFOA, PFHpA, 5–56 Kinetics offers a lower pH [208]
(H3PW12O40) 8 W @ 254 nm; PFHxA, PFPeA, dependence to pH compared to
PFBA other materials.
PFPrA, TFA
15 PFOA Ga2O3 PFOA 75 PFOA, PFHpA, 8–30 Nitrogen bubbling supports the [97]
μmol L− 1 with 0.5 g L− 1
PFHxA, PFPeA, kinetics while O2 bubbling
β-Ga2O3 powder; PFBA slowed it down
254 nm @ 15 W; PFPrA, TFA
16 PFOA MnOx vacancies rich PFHpA; PFHxA; 135–220 at pH 3.8 Synergistic •OH and H+ [184]
doped In2O3 PFPeA; PFBA degradation pathway.
(continued on next page)

9
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

Table 3 (continued )
Entry PFOA Materials details and Catalytic operation Major by-products Kinetic degradation Comments on performance Reference
properties conditions besides H+ and F- rate (min-1)

1 g L− 1 catalyst PFOA
50 mg L–1; 100 mW cm–2
(500 W nominal) @ 254 nm
17 PFOA Photocatalyst- UV–vis irradiation with CO2, F– 1.63 µmol h–1 No consumption of catalyst [102]
heteropolyacid tungstic Xe-Hg lamp (200 W) takes place
Electrocatalysis
18 PFOA Mixed composite thin 100 mg L− 1 Not assessed 10–64 [209]
film materials with Ti/ PFOA concentration at a
SnO2-Sb, Ti/SnO2-Sb/ constant current density of
PbO2 and Ti/SnO2-Sb/ 10 mA cm-2 with a
MnO2 10 mmol L− 1 NaClO4
supporting electrolyte
solution; Current densities
ranging from 5,
10, 20, 30 and 40
10 mA cm–2
19 PFOA Boron-doped diamond pH 3; − 0.5 to − 0.6 V PFOA, PFHpA, 46–103 Changes in Fe and Mn oxidation [108]
anode; FeMn-doped PFHxA, PFPeA, states and densities visible after
carbon aerogel cathode PFBA reaction
PFPrA, TFA
20 PFHxS, Titanium suboxide 5 mA cm− 2; 1–15 mA cm–2; Not assessed Individual (5×10–6 A systematic study looking at [210]
PFBA, electrodes SA of 468.05 cm2 to 3.5×10–8 m s–1) the impact of mixing various
PFBS, Mixture (1 × 10-5 to PFOA and PFOS compounds
PFPeA, 8 × 10–8 m s–1) together on selective
PFHpA, degradation
PFHxA
21 PFOA Ti/SnO2-Sb/PbO2-Zr 5–30 mA PFOA, PFHpA, 110–410 Impact of magnetic stirring [211]
electrode cm− 2; 30–75 ◦ C; 100 mg L− 1
PFHxA, PFPeA, beneficial to the
PFOA; 7.5–16 V applied PFBA homogenization and
voltage PFPrA, TFA defluorination process
22 PFOS F and Sb co-doped Ti/ PFOA 100 mg L–1; pH 3; PFOA, PFHpA, 150–350 Enhanced affinity through [212]
SnO2 electrode with Sn- 3–5 V; 1–17 mA cm–2; PFHxA, PFPeA, fluoride doping
Sb interlayer stirring PFBA
PFPrA, TFA
Photo-electro catalysis
23 PFOA rGO/BiOCl composite 14 W @ 254 nm; bias Not assessed 0.9 (EC), 7.9 (PC), 6.7 times increase in [107]
powders/assemblies voltage 40.1 (PEC ozone), performance with combinatorial
of + 2.0 V; 72.8 (PEC with processes; 15% loss in 4 cycles
1
Ozone dosage of 25 mg h− Ozone)
bubbled in oxygen at
10 mg L− 1 in PFOA
solution at pH 3

photolytic degradation kinetics of model PFOA and PFOS solutions, with generation and radical formation [95–97].
reported chain lengths of 6–10 carbons, was poor (10–7 to 10–5 min–1) Generating vacancies across single metal oxides leading to electron-
(Table 5, Entries 12–14). However, the photocatalytic degradation ki­ holes into the surface microstructure of the photocatalyst was also
netics of PFOA in either complex model solutions, with mixed PFOA demonstrated to lead to enhanced reactivities with titanium sub-oxide
sources and other contaminants mimicking wastewaters, or real efflu­ as well as In2O3 materials [98]. Oxygen vacancies generation was
ents were also low on the order of 10–9 to 10–8 min–1, demonstrating the found to increase the reactivity towards PFOA by 20–100 folds for both
need for more efficient and sustainable radicals generation pathways titanium oxide and indium oxide photocatalysts [96,99]. The highest
[92,93]. reported catalytic rate, to-date, was achieved for hollow indium oxide
Photo-catalysis (PC), however, involving heterogeneous photo­ spheres with a few percent of oxygen vacancies which reached
catalysts was found to decompose PFOA compounds at longer light 1323 min-1 for PFOA. Other ways of improving photocatalysis have been
wavelength than photolysis and to provide greater yields and thus much discussed in Section 6.
faster kinetics (Table 3, Entries 1–17). During heterogeneous PC, the Under environmental conditions, PFOS is not known to undergo
photocatalyst will absorb light and generate electron transfers towards photolysis, and the half-life of PFOA was estimated to be 349 days [100].
the surface of the semiconductor. The energy input will either split In lab-scale studies, direct photolysis with H2O2 and UV–visible irradi­
water, generating radicals, or directly degrade absorbed contaminants ation produced CO2, F– (~27–57%), and shorter-chained PFAS,
onto the catalyst surface. The first studies on PC of PFOA compounds degrading 42–56% PFAS [101]. The use of heteropolyacid photocatalyst
were carried out in the early 1990 s with titania macroparticles slurry under tungstic irradiation of 200 W xenon-mercury lamp resulted in the
reactors, generated by sol-gel, and led to rates of reactions in the order of complete mineralization of PFOA (≈100%) into CO2 and F–. No photo­
10–5 to 10–6 min–1 much slower compared to rates reported for other catalyst was used in this process, leaving behind the entire photocatalyst
contaminants, such as pharmaceuticals, dyes or other persistent organic for reuse [102]. Calcium (Ca2+) present across calcium gluconate can be
pollutants (POPs) [90,94]. Stemming from this study, nanoscale single used, in reactions where fluoride is generated as a reaction metabolite,
semiconductors, such as titanium dioxide (TiO2) nanoparticles and to form CaF2 to capture F–, which is environmentally benign [103].
nanotubes, gallium oxide (Ga2O3) nanosheets, and indium trioxide Within 60 min, ≈100% PFOA was degraded to SO4 and O2 anions, e–,
(In2O3) nanoparticles were developed for slurry reactor applications and and shorter chained intermediates when irradiated with UV with Ga2O3
found to be up to 1000 times more effective for PFOA treatment than and peroxymonosulfate as co-catalysts [104].
direct photolysis due to the larger surface area offered for photo-electron

10
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

Fig. 3. Comparison of photocatalytic or electrocatalytic performance in terms of wavelength of light and power consumption linking it with the corresponding
degradation rate has been shown. Data recorded in Table 3. Power of approximately 250 W and wavelength of visible light range between 380 and 600 nm are
effective in photoelectrocatalytic removal.

3.1.2. Electro-catalysis split in the presence of Fe2+ ions, have been demonstrated with porous
Electrocatalysis (EC), an effective advanced oxidation process, has carbon electrodes, offering excellent pore interconnectivity and
been successfully used for the degradation of micropollutants in conductive properties for the reactions [109]. However, adding Fe to
wastewater including dyes, other POPs, and more recently PFOA com­ porous carbons was found to largely alter the 2-electron oxygen reduc­
pounds [105]. Electrochemical treatment of aqueous pollutants allows tion pathway into a 4-electron pathway, resulting in much slower H2O2
for fast oxidation performance in relatively mild operating conditions yield rates. Other metals, including Mn, have been used in such reactions
and does not require extensive chemical pre-treatments, limiting the but still await their use for PFOA treatment.
environmental footprint of the technology [106].
The cost-effectiveness of the EC process, however, remains low due 3.1.3. Photo-electro catalysis
to the high energy requirements, relatively unstable electrode materials Recently, the development of photo-electrocatalytic (PEC) reactors,
over time due to pitting and corrosion issues [105, 106]. Several side combining the benefits and reactivities of both photo and electro­
reactions will occur simultaneously during any EC process, including a catalytic processes have emerged as efficient alternatives to AOP for
hydrogen evolution reaction (HER) at the cathode and an oxygen evo­ POPs remediation. PEC is particularly promising since offering a de-
lution reaction at the anode. Although the rate of hydroxyl radicals multiplied rate of degradation compared to single catalytic process
generation increases semi-linearly with respect to the current density, or (Table 3, Entry 23).
applied voltage across the electrode stack, while the system’s power The application of PEC was demonstrated for the degradation of
consumption will increase as the square of the voltage, making the EC PFOA with reduced graphene oxide (rGO)-BiOCl composites. The com­
process often limited to low operating voltages [107]. A strategy to bination of PC and EC into a PEC system led to a significant increase in
overcome this energy loss at high voltages, while still benefiting from the catalytic performance by up to 6.7 times [107]. The direct rate for EC
the high yield of radical generation from EC, is to develop combinatorial was 0.9 min–1, while this increased to 7.9 min–1 for PC. PEC reached a
processes where EC operated at low current densities is used in parallel promising value of 40.1 min–1 while the addition of O3 to the process
to either hydrothermal or ultrasonic processes. doped the reaction up to 72.8 min–1. No measurable degradation of the
To date, the focus of PFOA treatment with EC has however been on materials was reported confirming that although the density of radicals
the development of more stable materials, able to operate over a larger generated was strongly increased, this did not come at the expense of
range of current densities without undergoing either dramatic integrity surface damage or poisoning of the materials. A challenge with BiOCl is,
failures or surface passivation/oxidation (Table 3, Entries 18–22). PFOA however, its dissolution over time limiting long-term performance and
was degraded in a system composed of boron-doped diamond (BDD) leading to Bi3+ ions release in the solution [110]. The comparison on the
anodes and Ti/SnO2-Sb-Bi cathodes [108]. The reaction mechanism was performance parameters of the EC and PC processes in terms of current
linked to favorable electron transfer from PFOA to BDD anode while density and light wavelength employed has been shown in Fig. 3.
other anode materials, such as Pt, Ti/IrO2, Ti/RuO2, and Ti/IrO2-RuO2
were found to be inefficient to eliminate PFOA. Although BDD elec­ 3.2. Wet Oxidation-reduction reactions
trodes offer better performance and well-established chemical and
electrochemical stabilities, long life, and high oxygen evolution poten­ Oxidation-reduction reactions involve electron transfer from an
tial (OEP), its manufacturing costs are redibitory to its scalable appli­ electron donor (reductant) to an electron acceptor (oxidant), which re­
cation. Conductive materials were therefore, considered as substrates for sults in either partial to complete mineralization of the target contami­
BDD deposition, including various metals and semiconductors. Although nant directly or through the formation of free radicals that degrade the
titanium metal offers advantages in terms of cost and scalability, issues target contaminants indirectly [16]. PFAS are highly stable molecules,
related to interface delamination and compatibility between the BDD thus degradation via free radicals is typically more efficient, rather than
and the TiO2 native oxide layer have limited their development. It is also the direct electron transfer. Fenton agents, heat-activated persulfate,
difficult to grow in-situ BDD on titanium due to the strong H2 uptake bimetals, sub-critical water, and reduction with elemental iron- Zero
observed by titanium metal during plasma-growth of BDD, leading to Valent Iron (ZVI) are the most commonly used reagents [111]. Fluorine
fracture and degradation of the titanium matrix. Several metal oxides has a reduction potential of 3.6 V, making it the most potent inorganic
such as PbO2, SnO2, RuO2, and TiO2 have been demonstrated as active oxidant; thus, mere one-electron oxidants are ineffective in defluori­
anode surface materials for EC because they exhibit high oxygen evo­ nating PFAS [112]. This is also, why hydroxyl-based (•OH) oxidation
lution potential leading to larger densities of hydroxyl radical genera­ technologies are mildly effective for PFAS oxidation. AOP with UV
tion as well as relatively high electron transfer capabilities. irradiation and use of catalyst- H2O2 or S2O2-8 resulted in < 10% of
Electro-Fenton reactions, relying on H2O2 generation and subsequent PFOA and PFNA and 10–50% of PFOS removal, tested with PFAS

11
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

Table 4
Case studies on Physicochemical degradation of PFAS.
Entry Treatment Substance Conditions Removal (%) Time Degradation by-products Scale Reference
(Initial
concentrations)

Wet oxidation-reduction processes


1 Heat-activated PFOS pH ≤ 3, activation No effect 6h C4-C7 homologs of PFOA Bench [119]
persulphate PFOA temperatures of up to 60 ◦ C, 86–98
oxidation S2O2–
8 concentrations of up
to 84 mM
2 UV-activated GenX – <5, 3h –CF2 removal at every step Bench [121]
persulfate PFOA 27 with the generation of SO2–4
and OH–
-
3 UV-sulfite GenX – >90 2h F , PFPrA, TPA Bench [121]
4 Ozonation PFOS, Flow rate= 2 mg L–1 81.08, n.d. All were detectable. Field- WTP [56]
(H2O2 + UV) PFOA 25 Ozonation is ineffective in
+ WTP PFAS degradation
processes
(Adsorption +
filtration)
5 ZVI + biochar PFOA, PFHpA, Hardwood-oak based 17–94 120 days F–, removal efficiencies Bench [213]
PFHxA, biochar, pH= 8.5 depend on carbon chain
PFPeA, PFOS, length, 0–17% removal for
PFBA, PFHpS, short-chained PFCAs (C6-
PFHxS, PFBS C7), 20–70% for short-
chained PFSAs
6 γ-ray irradiation PFOA 50 kGy, Ar saturated 100 2–3 h C6F13-CF(H)-COOH Bench [164]
with a 60Co (10 mg L− 1) atmosphere, pH 6.7, 50 PFOA substituted with alkyl
source, Catalysts- t-butanol/H2O2 groups
Electron beam EB: 100 kGy, pH 12.5 C7F13-COOH,
(EB) CF3CHCH2COH
irradiations (CH3)2COOH
PFOA substituted
with different
alkyl groups
7 γ-ray irradiation PFOA 96 Gy/min, pH 13, N2 Degradation 12 h PFBA, PFPeA, PFHxA Bench [214]
with a60Co (20 mg L− 1) saturated atmosphere, rate- 0.67 h–1,
source Catalysts- t-butanol/H2O2 > 99
8 Plasma-based PFOS, PFOA with Ni-Cr and Al ring as 96–98 1h PFHpA, PFHxA, PFPeA, Bench [215]
reactor electrodes PFBA
PFHxS and PFBS (from PFOS
only)
Sonochemical degradation
9 Sonochemical AFFF (ANSUL 500 kHz or 1 MHz PFSA- 90.5, 13 h F– and SO2–
4 Bench [130]
reactor with 12 and 3 M) PFCA- 26.6,
transducers (P. FTS- 38.4
D.=12 kW,
Initial pH=
<4.0)
10 Sonochemical K-PFOS 400 kHz 96.9 Up to 4 h PFOA, PFHpA, PFHxA Bench [134]
500 93.8
1000 91.2
11 Sonochemical PFOA (240 nM) 358 kHz Not calculated 14 min Not studied Bench [127]
(A 20 kHz PFOS (200 nM) 358 17 min
horned PFBA (470 nM) 610 33 min
transducer PFBS (300 nM) 610 33 min
placed PFHA (320 nM) 610 19 min
perpendicular to PFHS (230 nM) 610 26 min
an Allied Signal
– Elac Nautik
ultrasonic
transducer, P.
D.=250 W L-1,
T = 10 ◦ C)
12 Sonochemical PFBS (300 nM) 358 KHz 66 42.3 min Not studied Bench [128]
(An allied Signal PFBA (470 nM) 57 57.2 min
– Elac Nautik PFHS (250 nM) 86 23.2 min
ultrasonic PFHA (320 nM) 90 16.8 min
transducer, P.
D.=250 W L-1,
T = 10 ◦ C)
13 Sonochemical PFOA 358 kHz 99 16.9 min CO, CO2, formaldehyde, Bench [135]
(An allied Signal (10,000 nM) 87 25.7 min carbonyl fluoride, HF
– Elac Nautik PFOS
ultrasonic (10,000 nM)
transducer, P.
(continued on next page)

12
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

Table 4 (continued )
Entry Treatment Substance Conditions Removal (%) Time Degradation by-products Scale Reference
(Initial
concentrations)

D.= 250 W L-1,


T = 10 ◦ C)
14 Sonolysis + O3 PFOS, 354 or 612 kHz 61, Up to 2 h Pseudo-first-order rate Field- Landfill [133]
PFOA 56 kinetics if followed, not groundwater
studied
Photo/Thermal
15 UV irradiation AFFF feed water UV intensity= 210–280 nm 16.8 20 min Limitedly studied, Bench [124]
PFHxA
17 UV–vis PFOA No photo or chemical 89.5 72 h CO2, F–, and short-chain Bench [102]
irradiation with catalyst perfluorocarboxylic acids
tungstic Xe-Hg
lamp (200 W)
18 UV–vis with PFOA Chemical catalyst- H2O2 35.5 24 h CO2, F– Bench [102]
tungstic Xe-Hg
lamp
19 Heat AFPO 196–234 ◦ C > 99 2.2–43.8 min, 1-H-perfluoroheptane Bench [138]
350–400 ◦ C, follows first- 0.2 s (90–95%)
order kinetics
20 Heat PFOA 355–385 ◦ C, follows first- ~100 2 s, T1/2 1-H-perfluoroheptane, SiF4 Bench [137]
order kinetics = 26 min
@ 385 ◦ C
21 Heat PFAS 350 ◦ C 99.91 10–14 days Volatile PFAS captured with PFAS- [140]
400 ◦ C 99.998 scrubber impacted soil,
in-situ thermal
treatment
22 Heat PFOS 726.85 ◦ C >99 0.2 s PFOA and SO2 Bench [136]

23 Heat 9PFAS- 350 ◦ C 43 (fortified), 75 min Not studied Bench and [216]
fortified and field 450 ◦ C 71 (fortified field
contaminated 550 ◦ C clay), 87
(fortified sand),
79 (field)
> 99 (field and
fortified)
71–99 (field)
24 Heat PFOS 350 ◦ C > 99 (with 1–30 min Defluorination, CaF2 Bench [217]
425 ◦ C lime)
> 99 (without
lime)
25 Heat+ Ca(OH)2 PFOS 300 − 600 ◦ C, 70–71, 15 min CaF2 and Ca5(PO4)3F Field- Sludge [165]
700 − 900 ◦ C 68
26 Heat+ Ca PFOS 300–900 ◦ C >90 (PFOS) 15 min CaF2, –CF, –CF3, –C3F3, PFAS waste [142]
(OH)2/CaCO3/ PFOA 450 ◦ C with Ca(OH)2 >80 (PFHxS) –C3F5
CaO PFHxS >50 (PFOA)
>45 (FOSA)

27 Heat+GAC+N2/ PFCAs 200–700 ◦ C, follows first- >99.99 (at n.d. Volatile organofluorine Bench [141]
∑7
O2/CO2/air 3PFSAs order kinetics 700 ◦ C) species, F–
atmosphere PFECA
28 Smoldering PFAS-impacted >900 ◦ C 16 <140 min Altered, shortened, or, Bench and [139]
column + GAC soil 44 volatile perfluorinated field
PFAS on GAC compounds

contaminated surface and groundwater [113] (Table 4, entries 1–8). 56 ng L–1 [120].
H2O2 [102], Na2S2O8 [114], O3 [115], Na/KMnO4 [116] were shown PFOS and PFOA underwent reductive dehalogenation by zero-valent
to successfully decompose PFOA into F– and CO2. OH– ions were also metals, such as iron with sub- and super-critical water at a temperature
generated, which helps in further oxidation. Na2S2O8 degrades PFOS of over >250 ◦ C, and pressurized between 0.78 and 0.81 MPa, forming
into PFOA with the help of SO•– 4 and (•OH) which further decomposes fluorine and sulfate [114]. Aqueous electrons (e–aq) have emerged as the
with the loss of ‒CF2 at every step until full degradation (defluorination) promising and powerful reductive agents produced in UV/sulfite [121]
[117]. H2O2 and humic acid were used as catalysts in such reactions. or UV/iodide [122] systems to decompose PFAS especially PFOS, PFOA,
PFOA was oxidized with heat-activated persulfate at and PFHxA. TiO2 due to its strong catalytic ability, hydrophilicity,
T = 50 ◦ C/pH = 2/t = 100 h, whereas PFOS did not oxidize even at chemical, structural stability, durability, nontoxicity, and economic
90 ◦ C. PFOA, likely, breaks down as follows with persulfate: PFOA, considerations coupled with morphological purposiveness, is another
PFHpA, PFHxA, PFPeA, PFBA, Perfluoropropionic Acid (PFPrA), Tri­ promising catalyzing agent [99]. Defluorination of PFOS occurs with
fluoroacetic Acid (TFA), and so on [118]. By-products have not been vitamin B12 as a catalyst and Ti(III) as a reducing agent measured in
consistently studied in these studies. In a bench-scale study, testing the terms of F– released with branched PFOS releasing much higher F–
efficiency of heat-activated persulfate towards the decomposition of (71%) as compared to linear (18%). This occurs because of the
PFOA and PFOS, PFOS remained unaffected, whereas PFOA showed defluorination reaction in which radical intermediates are formed
98% removal within 6 h. Undesirable shorter-chained intermediates, around the carbons, breaking the C‒C bond [123].
chlorates, and metals were also flocculated in the process [119]. In During ozonation processes, O3 is injected directly into the waste
another field study, oxidative conversion resulted in an increase of stream [124]. O3 failed to degrade PFOS or PFOA in a peroxone process
PFCAs in the urban run-off by a median of 69% between 2.8 and at 4–5 pH. Although alkaline ozonation removed 71% of PFOA and 76%

13
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

of PFOS in a peroxone-assisted process increasing the removal to 99% [137,138]. The remediation of PFAS-impacted soils by smoldering with
for PFOA and 89% for PFOS within 4 h of operation [115]. Similarly, a and without GAC-mixed soil was evaluated. Up to 44% of degradation of
process called ozofractionation with O3 injected into a foaming frac­ PFAS was achieved with GAC compared to only 16%, without GAC. The
tionator instead of directly in air exploits the surfactant properties and total organic fluoride analysis confirmed that GAC-application enhanced
foaming ability of PFAS towards water remediation. The hydrophilic the rate of PFAS volatilization from soils and water, which supported the
polar head and hydrophobic carbon tail helped in micelle formation. On decomposition of PFOA at low operating temperatures (down to
sparging O3, a PFAS foam was formed which was separated from the 200 ◦ C). Only sub-1% levels of PFAS were emitted as quantifiable PFAS
water stream. Ozofractionation achieved a 99.99% reduction of PFAS after treatment [139]. PFSAs (400–450 ◦ C) require higher temperatures
from 28.8 to <0.002 ng L–1, in this manner [125]. than PFCAs (350 ◦ C) to completely volatilize [140]. In an N2 atmosphere
and at 700 ◦ C, >99.99% of PFOA and PFOS were decomposed to yield
3.3. Sonochemical degradation fluoride ions as major quantifiable by-product [141].
PFOS decomposition takes place by hydrogen fluoride (HF) loss
Upon application of an ultrasonic field (>16 kHz), the soundwaves resulting in an α-sultone formation that decomposes spontaneously to
cause the bubbles present in the solution to reach their maximal radial form PFOA and SO2. The PFOS decomposition did not begin with C‒S
expansion and undergo a quasi-adiabatic compression. This energy is bond fission. The terminal functional groups (‒SO3H, ‒COOH, ‒
converted into kinetic energy of the molecule, which is released when SO2NH2, etc.) detach entirely during PFAS incineration. The acid head
the bubble collapses to burst. This phenomenon is called cavitation. group of PFOS was destroyed at 1000 K, and PFOS degraded following
Temperatures close to 5000 K are generated in this process, resulting in first-order kinetics with a half-life of 0.2 s [136]. The lime treatment of
the pyrolysis of any chemicals, such as PFAS, present in the solution PFOS between 350 and 900 ◦ C, resulted in the formation of CaF2, ‒CF, ‒
[126]. In this reaction, PFAS is mineralized by two processes, including, CF3, ‒C3F3, ‒C3F5 [142]. Shorter, altered, or more volatile forms of PFAS
the cleavage of ionic functional groups by pyrolytic action followed by were hypothesized to be formed in the process with the potential of
the defluorination resulting in smaller molecular weight PFAS com­ long-range transport [139]. To minimize the spread of volatile pre­
pounds formation. The frequencies used range from 202 to 1500 kHz cursors of PFAS, a scrubber is recommended to be used at thermal
with power densities of 83–333 W L–1 [127–132]. The cavitating bub­ mineralization sites. Toxic greenhouse gases such as dioxins, furans, CF4,
ble’s number, radius, temperature, pressure, dynamics, and symmetry and C2F6 were produced alongside CO2 during this operation in the field
are determined by the applied frequency [128]. The number of active [143,144].
bubbles per unit time and the mass transfer to the bubble surface of the
PFAS species determine the rate of degradation of PFAS. 3.5. Biological degradation
In a lab-scale study conducted, at 400 kHz, 96.9% of degradation
was observed over 4 h. The hydrated electron release mechanism has WWTPs may serve as the largest target sites for large-scale opera­
been suggested for such a reaction resulting in defluorination of PFOS, tions for the treatment of PFAS via microbial- or phyto-degradation.
accompanied by sulfate release. CO and CO2 were also produced during Wetlands are known to represent physicochemical conditions of
this degradation mechanism [133] (Table 4, Entries 9–14). To enhance WWTPs, therefore, also serve as a repository of microbes and plants
degradation, in an experiment, dual-frequency exposure of useful in bioremediation of targeted contaminants. This section dis­
20 + 202 kHz and 20 + 610 kHz was studied. Enhancements of 12% cusses the potential of microbes and wetland plants, besides enzymes
and 23% were achieved for PFOS and PFOA degradation respectively at and protein to treat PFAS from wastewaters.
20 + 202 kHz. However, there was relatively no enhancement at a fre­
quency of 20 + 610 kHz. On single frequency exposure for up to 4 h, 3.5.1. Wetlands plants and microbes for bioremediation
tested at various frequencies of 400, 500, and 1000 kHz, 96.9%, 93.8%, Though initially microbial degradation was suggested to be of
and 91.2% respectively of PFOS was degraded. At higher frequencies up limited importance for PFAS degradation [19,145], recent evidence
to 1000 kHz, the degradation efficiency decreases to ~91% because of suggests enrichment of different bacterial phyla in the order of Proteo­
the overall reduction in the intensity of cavitation frequency collapse bacteria > Actinobacteria > Verrucomicrobia > Bacteroidetes > Cyano­
[134]. At 500 kHz, 90.5% of Perfluoro Sulphonic Acids (PFSAs), 26.6% bacteria > Firmicutes in water-soil-plant systems fortified with varying
of Perfluoro Carboxylic Acids (PFCAs), and 38.4% of Fluorotelomer concentrations of eight PFAS, indicating their relevance in biodegra­
sulfonate (FTS) were degraded with 12 transducers in 13 h [129]. At dation [146]. Acidimicrobium A6, an Actinobacteria, follows reductive
higher initial concentrations of 10,000 nM, up to 99% of PFOA and 87% defluorination and removes ~60% of PFOA/PFOS (100 mg L–1) within
of PFOS were degraded at 358 kHz [135]. In a field-scale study of 100 days [147]. In aerobic conditions, the presence of N-ethyl per­
groundwater impacted with PFAS, PFOS and PFOA were reduced by fluorooctane sulfonamide (N-EtFOSA)-enriched Methylocaldum in wet­
61% and 56% respectively by sonolysis, carried out in the presence of lands as potential N-EtFOSA-degrader, whereas under anaerobic
other organic contaminants at a frequency of 354/612 kHz and power conditions showed a dominance of Methanomethylovorans [148].
density of 250 W L–1 [133]. Both aerobic and anaerobic degradation pathways of 8:2 Fluo­
rotelomer Alcohol (FTOH) have been elucidated [149]. Acidimicrobium
3.4. Photo/Thermal degradation Gordonia, Desulfococcus, Pseudomonas, and some other lineages of bac­
teria use fluorotelomers such as 6:2 FTS as sulfur sources [150]. PFAS
Thermolysis involves using high temperatures to break the C‒C or C‒ show resistance to anaerobic bacteria and anaerobic bacteria are im­
F bonds of the PFAS chain to form shorter chained moieties which will mune to PFAS toxicity and thus, they continue their methanogenic ac­
eventually be degraded like the photolytic degradation products. tivity in PFAS contaminated sites [151]. AFFF/PFAS mixture also failed
Incineration remains the most commonly applied thermal degradation to inhibit microbial methane generation, and most PFAS sorbed onto the
technique [136]. Other pathways utilize chemicals and microbes and did not undergo any decomposition [152].
critical/sub-critical ions to support thermal dissociation of PFAS, and Plants with varying capabilities of uptake, translocation, and
the most efficient studies are discussed in Table 4, Entries 15–28. biotransformation of PFAS from industrial effluent, contaminated irri­
Incineration of PFAS such as PFOA, PFOS, and Ammonium Per­ gation water, and landfill leachates have been reported. The fate of PFAS
fluorooctanoate (AFPO) takes place successfully at temperatures as high in plants remains a net outcome of complex interactions between
as 600–1000 ◦ C. The estimated half-life of PFOA was 26 min at 385 ◦ C, physicochemical properties of PFAS, the physiological capability of the
while ~99% was degraded within 9 h leaving 1-H-perfluoroheptane as plant species, and abiotic factors. The PFAS concentration ratio in the
the major by-product and Perfluoro-1-heptene as the minor by-product plant to environmental matrices (wastewater, surface water, lake, and

14
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

Table 5
Case studies on biological/biochemical degradation of PFAS.
Entry Treatment Substance Conditions Removal Time Degradation by-products Scale Reference
(%)

Organismic
1 Phanaerochete 6:2 FTOH, Lab-cultured 50, 28 F–, 5:3 polyfluorinated acid (40%), Bench [218]
chrysosporium 8:2 FTOH 70 days 5:2 sFTOH (10%), PFHxA (4%), and
others
(about 1% each). Fewer metabolites
were produced after 8:2 FTOH
degradation, such as 7:2
sFTOH (6%), PFOA (5%), 7:2 Ft
ketone (3%), and others (<1%
each).
2 Ti(III)-citrate with Branched An aqueous solution with pH 9.0 71 7 F- Bench [123]
vitamin B12 as a catalyst PFOS and at 70 ◦ C days
from cyanocobalamin
3 Acidimicrobium sp. strain PFOS, PFOA The initial concentration of <60 100 F–, shorter PFAS, acetate, Fe (III) Bench [147]
(A6) 0.1 mg L–1 and 100 mg L–1, ferric days
ions as the catalyst, hydrogen as an
electron donor
4 2 strains of Pseudomonas PFHxS, 8:2 Field-isolated strains, 40 ± 3 10 Bioaccumulation, F–, shorter PFAS Bench [219]
sp. FTOH Alkanotrophic conditions using days
octanol
Enzymatic
5 Laccase-mediated PFOA Batch-reactions ~50 157 pseudo-first-order with rate Bench [159]
oxidative humification days constant of 0.0044 day–1, No PFCAs,
F–, shorter-chained aliphatic
compounds
6 Horseradish peroxidase PFOA Batch reactions 68 6h F–, and shorter-chained aliphatic Bench [158]
+ 4-methoxy phenol compounds
7 Hemp protein (Cannabis PFOS+PFHxS Pump and treat batch reactions 97.9 1h Not studied Bench [161]

sativa) PFAS 95.2
8 Soy and pea protein 97
82
9 Egg protein 48
30
10 Lupin protein 85
73
11 Whey protein 29
9

agriculture fields) in yam roots (0.53–0.59) and sugarcane stems or removal. Horseradish peroxidases co-polymerize phenol and PFOA in
(0.65–0.67) confirmed significant uptake of PFBS and PFHpA, respec­ the presence of H2O2 from an aqueous solution and about 68% removal
tively [153]. Bioaccumulation and translocation in roots and shoots of of PFOA was achieved [158]. Laccase also reduces 50% PFOA in 157
Juncus effusus varied with the carbon chain length in a greenhouse study. days. Both peroxidases and laccase are humification enzymes, which
J. effusus-soil interfaces together accumulated 82.8% of PFAAs (C4 to degraded PFAS following either radical abstraction of double bonds or
C8) from aqueous solution in 21 days [154]. Wheat, soybean, and Kolbe’s decarboxylation in a stepwise conversion of ‒CF2 moiety
pumpkin; grown in hydroponics; effectively metabolized N-EtFOSA into generating CO2 and F– [159]. Enzymatic degradation yields partially
FOSA, Perfluorooctane sulfonamide (PFOSA), PFOS, and other fluorinated shorter-chained alcohols and aldehydes, lacking
shorter-chain PFAS, PFHxS, and PFBS in roots and shoots with shorter-chained PFCAs. Peroxidase/laccase-catalyzed oxidative humifi­
biotransformation pathways different from animals and microbes [155]. cation reactions are prevalent in the environment, therefore, serve as
Different wetland plants, Alisma orientale, Arundo donax, Canna indica, environmentally benign methods for PFAS treatment [160].
Cyperus alternifoliusm, C. papyrus, Thalia dealbata, Phragmites australis, Protein powder of hemp (Cannabis sativa) has shown >98% removal
Pontederia cordata, show potential to accumulate PFOS mainly in roots of PFOS and PFHxS from real contaminated groundwater in a pump and
(48.8− 95.8%), while PFOA mostly in shoots (29.3− 77.4%). Localiza­ treat system in a contact time of 1 h/pH 4–6, where salinity exerted a
tion of PFAS confirmed that uptake and translocation of PFAS occur both positive impact on removal rate [161]. Hemp protein adsorbs PFAS
across cell walls and/or intercellular spaces, and across plasma mem­ comparable to a GAC: Norit® (<1% difference for total PFAS) but re­
branes [156]. Another study from riparian wetland showed high affinity mains ~2.7 times cheaper than GAC (hemp protein: US$45 kg–1 vs GAC:
of Cyperus congestus, Eichhornia crassipes, Persicaria amphibia, Phragmites US$125 kg–1) [162]. Thus, the combinatorial or tandemly applied
australis, Polygonum salicifolium, Ruppia maritime, and Schoenoplectus technologies have come up as a solution to the problems of applicability
corymbosus to PFOA accumulation in the range 11.7–38 ng g–1, with a and scalability (Table 5, Entries 1–11).
BCF range of 0.05–0.37 [157].
Taking clues from emerging reports of wetland plants and microbes
as PFAS accumulators and degraders, a constructed wetland may serve 3.6. Limitations and directions for PFAS degradation technologies
as an integrated system relying on physicochemical and biological
processes for large-scale, eco-friendly treatment of PFAS, combining Chemical degradation technologies are mainly hindered by in­
benefits from different types of treatment methods. terferences by ions and their untargeted actions. The need to maintain
conditions of pH and temperature, slow reaction rates, as well as the
3.5.2. Enzymes and proteins for bioremediation addition of chemicals further add to the composition load of the water
Both enzymes and proteins have shown promise in PFAS degradation are however limiting the extension of these technologies. Additionally,
the reaction intermediates and metabolites produced could be more

15
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

Table 6
Case studies on treatment train approach for PFAS treatment.
Entry Treatment Substance Conditions Removal (%) Time Degradation Scale Reference
by-products

1 UV-Air fractionation AFFF feed water UV intensity = 210–280 nm, 81.3% PFAS in total, 20 min A limited study, Bench [124]
airflow and 80% PFHxS, PFHxA
rate of 30 L min− 1 and feed 100% PFHpS, 87.0%
flow rate of 1.4 L min− 1 PFOS, 16.7% PFHxA,
50.0% PFHpA, and
91% PFOA
2 UV/O3 AFFF feed water UV intensity = 210–280 nm, 73 20 min A limited study, Bench [124]
Ozone generator PFHxA
wavelength = 185 nm
3 UV-O3 - air AFFF feed water UV intensity = 210–280 nm, > 95 20 min A limited study, Bench [124]
fractionation Ozone generator PFHxA
wavelength= 185 nm, air
flow
rate= 30 L min− 1 and feed
flow= 1.4 L min− 1
4 WTP + RO + advanced PFOS, PFOA UF= 0.04 nm pore size, <MQL n.d. Could not be Field- [56]
oxidation RO= >300 MW substances detected, hence WTP
rejected not studied
5 GAC + ZVI + V-UV + PFOA [GAC]= 12.5 mg L–1/[Fe0]= 50 5–7 h C7, C6, C5, C4, Bench [176]
Fenton 7.5 gL-1/[H2O2]= and C3
22.8 mmol L–1, UV @ 185 nm
6 UV + iodide + Humic PFOS pH= 10, [HA]= 1 mg L–1, 55.6 1.5 I2, I–, I–3, IO–3, Bench [220]
acid [PFOS]= 0.03 mM HOI
7 GAC + AER + NF PFAS 270 Da cut-off of N.F. >99 35 weeks Not studied Pilot [221]
8 Aeration + duckweed PFAS The initial concentration of >95 (80% of PFOS 7 h +2 weeks Not studied Bench [178]
200 ng L–1, pH 2.3, aeration with aeration)
of 7 h
9 O3 + GAC+ PFAS such as Regenerated GAC with 10–99 90 days Not studied Pilot [222]
biodegradation PFBS, PFHpA, 0.65 O3
PFNA, and
PFOA and flame
retardants
10 GAC vs. Superfine Contaminated The SPAC has a mean particle For 77, 57, 200, 42 Concentration Pilot [223]
powder activated groundwater, diameter of 0.88 µm groundwater = GAC (h) study,
carbon (SPAC) + AFFF feed water and it was delivered as slurry = 6.2 Not studied
ceramic microfiltration having a density of 100 g- SPAC = 11
(CMF), SA of SPAC SPAC L–1, membrane flux For fire training
= 927 ± 40 m2 g–1, 60–65 LMH with a crossflow grounds =
dosage = 500 mg L–1 velocity of 0.18 m s–1 SPAC = 2990
(µg PFAA g–1 AC)
11 Adsorption + PC Iron hydr Initial PFOA conc.= 200 µgL- 95.2% 1h C7, C6, C5, C4, Bench [224]
1
(oxide) & , Adsorbent conc.= 1 gL-1 in photodegraded, C3, and so on,
carbon spheres 1:1 ratio, pH= 7, t = 4 h, 57.2% defluorinated OH–
solar light through quartz
photo-reactor while stirring
at 200 rpm
Entry Process Membrane code Compounds Treatment Feed Efficiency (%) Scale Reference
conditions concentration
(mg L–1)
12 NF + GAC NF270 PFBA, PFPeA, PFHxA, PFOA, Crossflow 0.001 93–99 in 8–33 Bench [73]
PFNA, PFBS, PFHxS, and synthetic days
PFOS, and PFDA groundwater
pH 3.5 and 7.1
13 NF + EO NF270 PFHxA Crossflow 64 and 204 89.1/88.7 in Bench [88]
Synthetic water 0.9–4.4 h and
pH 2–11 field
14 NF + 2540-ACM5- TSF PFOA Crossflow 0.005 and 0.1 66.9 in 1 min Field [89]
photocatalysis 2.5 in. (6.4 cm) in Synthetic water
+ UF diameter pH 2–11
15 RO-UV + AOP n.d. PFPnA, PFBS, PFHxA, PFHpA, Crossflow 52–227 >99 (Time = n. Bench [72]
PFHxS, PFOA, 6:2 FtS, PFNA, Tertiary effluent d.) and
PFOS, PFDA field
16 O3 + PFPnA, PFHxA, PFHpA, and 53, 30, 53, and
Biologically PFOA 24 respectively
activated for 25–35 min
carbon (BAC) +
GAC

Notes: WTP processes include coagulation/flocculation and sedimentation, then ultra-filtration (UF), reverse osmosis (RO), advanced oxidation, and finally
stabilization.

16
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

toxic than the primary PFAS target [19]. degradation techniques such as EC, AOP, advanced reduction processes
The slow kinetics of PFOA degradation was found to be largely (ARPs), and sonochemical has proven to radically improve the rate of
dependent on the light wavelength used as well as on the overall energy PFAS treatment. These are known as treatment train processes and are
input provided to the system, due to the very high C-F bond energy of the used to achieve high in-situ treatment efficiency [174]. In a pilot-scale
system. Although the range of non-terminal by-products produced was study, a waste disposal site was used as a source of PFAS. GAC was
found to be typically limited, their stability and typical low concentra­ used to adsorb PFOA, PFOS, and PFHxA followed by thermolysis to
tions, make their degradation extremely challenging. In addition, the obtain fluorine recoveries of 51%, 70%, and 74% respectively, up from
release of fluoride ions, leading to HF formation is a major hazard and no 30%, 72%, and 46% without GAC. The use of a catalyst- NaOH further
study to date looked at synergistic degradation/recovery strategies to improved the recovery by 74%, 90%, and 91% respectively for PFOA,
remove fluoride acids while they form. Further materials development, PFOS, and PFHxA [175]. These studies are summarized in Table 6, En­
particularly regarding the surface area and oxygen vacancy control tries 1–16.
should also be considered with stabilization strategies involving metal The use of bimetals and zero-valent metals along with adsorption
doping or nano-texturation. techniques for capture are often recommended. The use of membranes is
Advanced oxidation and reduction processes are able to achieve often also desirable to achieve higher removal efficiencies followed by
partial or complete degradation with efficiencies between 55% and PFAS degradation techniques [13]. Oxidation by ZVI and adsorption by
100% [163]. The oxidation of F– is thermodynamically unfavorable due GAC were combined with photocatalytic degradation by a vacuum
to the high electronegativity of the element. PFOA or PFOS have no UV-Fenton system at 185 nm to degrade PFOA. Generation of aqueous
hydrogen that can be acted upon by hydroxyl radicals to form less electrons helps in oxidation of iodide, which drives defluorination of
thermodynamically stable hydrogen ions, thus simple oxidation PFOA. From an initial concentration of 10 mg L–1, PFOA was effectively
methods are ineffective on PFAS. However, AOPs with persulfate and defluorinated to a final concentration of 3.23 mg L–1 with [GAC] =
ionization radiations have been proven to be effective in degrading PFAS 12.5 mg L–1/[Fe0] = 7.5 gL-1/[H2O2] = 22.8 mmol L–1. Intermediates
[164]. Similarly, ARPs with activated persulfate and aqueous electrons such as C7, C6, C5, C4, and C3 were generated in the process by the
have proven to be more useful for PFAS degradation. photochemical activity of H2O2 generated OH– [176]. The addition of
Incineration is a common and well-studied PFAS remediation KOH and removal of the water from sand resulted in complete
method, which can guarantee the complete decomposition of a large defluorination in a ball milling experiment with AFFF-impacted soils.
range of PFAS compounds. However, extremely high-temperature gen­ 98% of PFOS and 99% of PFOA were removed in 4 h. Up to 96% removal
eration also entails heavy capital investment on the equipment. The was also achieved on-site at a firefighting training area with this tech­
generation and potential release of toxic fluorinated and greenhouse nique [162]. The continuous application of a mechano-chemical and a
gases such as hexafluoroethane and tetrafluoromethane with warming hydrothermal process resulted in the adsorption of PFOA in a fixed bed
potential of 5700 and 11,900 and half-lives of 50,000 and 10,000 years reactor. Mg-Al-layered double hydroxides- LDH-1 and LDH-2 were able
respectively require further cleaning technologies such as scrubbing and to remove 90% and 98% of PFOA within 2 h and 40 min respectively
electrostatic precipitation, which might concur additional input costs. with an initial concentration of 20 mg L–1 of PFOA, volume of 20 mL,
However, the use of Ca(OH)2 or persulfate can bring down the release of and 0.05 g of adsorbent. This increased its potential for field application
toxic gases from both sludge and solid waste to produce environmental- [177].
friendly CaF2 [118, 142, 165]. Phytoremediation efficiency of 14.4% with duckweed was also
Sonochemical degradation enables the generation of temperatures achieved by adding an aeration step with more than 95% PFAS removal
above 500 ◦ C that are sufficient to degrade a broad range of PFAS from deionized water with 7 h of aeration [178]. In a patented tech­
compounds, which reaction pathways are sustained for nano-seconds. In nology, coagulation combined with air-sparging and hydrocyclone was
dilute concentrations of about 1 µM, sonochemical degradation works used to report 80–97% removal from AFFF-contaminated groundwater.
well but it is rendered unsuitable for real wastewater samples, where the The water was chemically treated with a metal coagulant and a polymer
water mixture is complex. The interferences in real wastewater bring and then it was sent to an air-sparged hydrocyclone (ASH) chamber
down the efficiency of such degradation processes. Sonochemical where air and water fall at a tangential angle to form bubbles or foam.
degradation must be coupled with other processes such as ozonation and This foam is then clarified of water in the subsequent steps [179]. The
filtration to make it self-sustaining or achieve higher removal rates. performance comparison of different capture, degradation, and combi­
Bioremediation is a promising field as it is an environmentally natorial technologies is represented in Figs. 1–3.
benign and economically cheap method. However, biodegradation The limitations of the different treatment processes are coupled in
processes for PFAS treatment are underdeveloped and understudied the treatment train approaches. ASH is desirable and patented because it
currently. It is hampered by extremely long reaction times (2 days to 2 can handle large volumes and other kinds of wastewater, with short
years), a very slow decomposition rate (36–60 years for reaction yield of retention times [180]. However, mass balance loss in the form of volatile
25% for PFCAs) [166], low removal rates (0–68%) [167,168]. The precursors occurred during the thermal treatment of the PFAS foam
half-life of precursors (1200–1700 years for poly(acrylate) and 24–281 [175]. High capital investment is guaranteed, however, without a 100%
years for poly(urethane)) [167], degradation rates, and mass balance guarantee of PFAS removal. High removal rates are achieved with
differ greatly because of environmental conditions. An estimate of the filtration during which sludge and oil-based wastes are generated in the
PFOA half-lives in these conditions was predicted to range between 10 ASH treatment process [180]. Energy- and cost-intensive incinerators,
and 17 years and 870–1400 years [169,170] thus hindering the scal­ sonicators, and electrochemical equipment, and loss of adsorbents are
ability of such processes (Summarized in Table 7, also Figs. 1 and 2). other shortcomings of such kind of processes [175].

4. Treatment train approaches 5. Future perspectives

Tandem use of the aforementioned technologies has proven to be Capture technologies are dominated by activated carbons, which
more effective in the case of treatment of POPs such as PFAS given its despite efficiently capturing PFAS compounds, were found to suffer
unique stability and recalcitrance [171–173]. To solve the limitations of high-cost penalties with respect to their regeneration, due to very robust
capital investment, time consumption, and efficacy of removal, treat­ PFAS materials interactions. The cost of the adsorbents also remains a
ment train approaches play a significant role in achieving higher hindrance, and significant cost-reductions could be achieved through
removal rates with wide-field applicability. The use of capture tech­ more sustainable materials development. Cost-analysis of the technol­
niques such as adsorption, filtration, and coagulation coupled with ogies should also be evaluated systematically to understand the

17
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

Table 7
The major limitations, advantages, and removals of various PFAS capture, concentration, and degradation technologies as concluded from the present study.
Type Technology Governing parameters Advantages Challenges Common Challenges Additional
References

Capture Adsorption Geometrical: Pore shape, GAC adsorb longer-chained Regeneration is difficult (ACs, Charge of the material
size, and surface area PFAS better biomaterials, and minerals) Transient charge on
Chemical: Molecular Adsorption capacities (g-1 of Regeneration process power- or PFAS molecules [17]
weight, charge, and PFOS/PFOA) pressure-driven (MIPs, CNTs and Effluent properties such
functional group PAC: 1.2–375 mg, IER: IERs) as pH, temperature,
2.5–3067 mg, electrical conductivity,
Chitosan: 645–2580 mg, CNTs: organic matter,
140–1651 mg inorganic
Removal efficiencies of 23% for Frequent replacement of ions [225]
bio-adsorbents and minerals to adsorbents
>99% for IERs, CNTs and ACs
Coagulation Stirring duration Ease of WWTP implementation Dose (~60 mg L–1 alum
remediates only 20% PFOS/
PFOA)
Coagulant dose Swift action in 0.167–1.5 h Corroding of electrodes (H+)
Electrode configuration 99% removal efficiency with Disposing-off used coagulants [192]
electrocoagulation Sludge management
Concentration Membrane Electrostatic repulsion NF (MWCO of 200–1000 Da), Pore size changes due to [88]
filtration more powerful than size Flux >95% changing pH
exclusion
High-pressure driven Effect of functional groups and
membranes (RO, NF), electrostatic repulsion
more effective
Pore-size RO (MWCO <100 Da), Flux Handling and disposal of [75]
Surface roughness >99% concentrates
Surface zeta potential Ease of membrane tuning and
Isoelectric point fabrication
Degradation Catalysis Oxygen vacancies Degradation rate- Poisoning of the catalytic system Formation of shorter-
0.2–1323 min–1 over time chained
Zeta potential Presence of oxygen holes and Doping or stress of degraded precursors and by-
porous structural design helps products/ions products
Radicals quenchers Removal of 80 to >99% Formation of GHGs and [108]
dependent on metal-oxide HF
doping Need of pre-filtration of
In2O3 better than TiO2 effluent (except for
Sonochemical Reactor configuration Up to 98% removals Presence of contaminants in incineration and in [133]
water such as bicarbonates and wetlands
VOCs hinder acoustic waves
propagation
Sonication time Additives, e.g., persulfate ups Energy consumption
removal
Irradiation density Dual frequencies ups removal Upscaling
Ultrasonic frequencies Treatment trains with ozonation Noise pollution
or UV
Power densities Useful for in-situ groundwater Aging of the ultrasound
Atmospheric conditions remediation sonicators
Initial [PFAS]
Operating pH
Temperature
Additives
Photo/ Temperature Low operating times of 0.02 s to Low thermalization potential of [165,175]
Thermal 75 min Volatile PFAS
Irradiation frequency Addition of Ca(OH)2 prevents High operation cost
CO2 leaching into the
atmosphere
Exposure time Supercritical water (94.8% Production of CF3H (Global
PFHxA) remediates more than warming potential:14,800)
subcritical water (83.6% PFHxS)
High mineralization rates
High removal efficiency
Oxidation/ Additive concentration Supercritical water (94.8% OH- less effective in PFAS [16]
reduction Reaction duration PFHxA) remediates > subcritical reduction
water (83.6% PFHxS)
PFAS functional group Highly reductive hydrated e– Presence of Cl- and solids in
water reduces mineralization by
25%
Reaction duration ZVI, fenton agents and electron Variable degradation rate:
Solution chemistry: pH, beams increase removals 0.041 min–1 with persulfate to
T, DO, and TDS 0.00023 s–1 with heat-activated
Organic and inorganic persulfate
constituents: HA, nitrate,
and sulfate
Biological Environment friendly [171]
(continued on next page)

18
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

Table 7 (continued )
Type Technology Governing parameters Advantages Challenges Common Challenges Additional
References

Seeding density of Toxicity or inactivation of


culture microbes due to harmful PFAS
metabolites
Effluents toxicity Cost-effective Disposal issues and trophic
transfer of bio-concentrated and
bio-accumulated PFAS in plants
Bioaccumulation Upscalable via. Economical Extremely slow degradation
potential options such as constructed rates
wetlands (0.003 min–1 for HRP,
0.0044 day–1 for laccase)
Initial [PFAS] Removal efficiencies of 18–96% Enzymatic degradation: 1–6 h [158]
Enzymatic removal >> faster than
Organismic removal Organismic accumulation/
Time of enzymatic degradation: breakdown: Up to 100 days
1–6 h

scalability of the approach as well as the feasibility of scaled-up opera­ semiconductor decoration [93, 189–191]. The presence of graphene was
tions and reactor design. The use of biomaterials (US$246 ton–1 for found to support electron conduction and therefore results in increased
activated biochar) over synthetic ($1500 ton-1 for activated carbon) yields of reaction. Key factors such as light source and wavelength, re­
could help in cost-cutting [181]. Hemp protein powder shows not only action temperature, solution pH, dissolved oxygen, and dissolved
high efficacy to remove PFAS but also remains cost-effective comparing organic matter present in the solution, as well as the presence of radical
with commercial adsorbents such as GAC. scavengers will also affect the rate of PFOA compounds degradation. The
Most membrane separation systems have emerged as promising al­ control of the process conditions is therefore as critical as the design of
ternatives on model solutions (Figs. 1–3) yet remain largely non- proper photo-catalysts.
competitive on large scales since only allowing for several factors of Electrocoagulation [58], AOPs, and ARPs [16] have proven to be
concentrations. RO still proves to be more cost-effective ($2 m–3), effective in remediation of waters from PFAS such as electrocoagulation
calculated with the average price of electricity at 8.14 cents KW h–1. removes 65% of fluorine from water but also offer the additional benefit
Conventional water treatment methods are still a mere $0.25 m–3, of treating other water quality parameters such as 60% of total organic
making it the most wide-reaching treatment practice [127]. The inte­ carbons (TOC) is treated with electrocoagulation which is of major
gration of membranes with catalysts however represents a promising concern for water quality keepers [192]. Fe2+ as the activation agent has
alternative to generate PFAS lean streams in a single step. proven to be most economical compared to other activation agents
Future degradation-focused research directions should focus on of­ [117]. The research on cost-effectiveness of oxidation-reduction tech­
fering greater yields of radical generation through more resilient and nologies is highly limited and more research is required to make these
efficient catalysts engineering while reducing the sensitivity of the cat­ methods scalable. Similarly, sono-chemistry and biological methods for
alysts to specific adsorption of radical scavengers. Strategies to develop PFAS treatment, although promising, are yet in their nascent stages and
high surface area photo-electrocatalytic reactors (PECRs) and photo- further demonstrations are required to establish the conversion mech­
electrocatalytic membrane reactors (PECMRs) should also be further anisms and the applicability of these methods to real wastewaters. These
investigated to allow for developing sustainable and cost-effective con­ limitations may be overcome by implementing combinatorial processes
centration/degradation processes. The development of advanced photo, such as dual-frequency systems for sono-chemistry are known to reduce
electrocatalytic materials, and membrane reactors has a great potential the cost by 23% to $7.5 m–3 from $10 m–3.
to play, beyond the sole scope of PFAS, across a range of POPs. Given the highly localized nature of PFAS contamination and varied
The presence of oxygen vacancies will deeply affect the local atomic source water compositions, it appears unlikely that a single technology
microstructure and electronic properties along the surface of the mate­ will emerge as a universal solution for PFAS contaminated source wa­
rial, therefore allowing for direct fine-tuning of their intrinsic physico­ ters. Further combinations of technologies, including coagulation/floc­
chemical properties in PEC processes [182,183]. The charge generation culation or air scouring with adsorbent or membrane materials, may
and separation, as well as subsequent surface reactions, may be equally offer technical advances supporting lower operating costs and
controlled through precise defect engineering. Particularly, the presence selective recovery of PFAS [54]. It is also important not to downplay the
of vacancies will (i) extend the light absorption range of the material, role played by other compounds present and mixed with PFAS com­
enabling its operation across a wider range of wavelength and poten­ pounds [193]. Water recovery as an obvious point represents a first
tially within the visible range, (ii) promote charge-carrier separation, critical milestone, but other compounds mixed with PFAS should also be
critical to reducing the probability of electron-hole recombination, as valorized towards a zero-liquid discharge (ZLD) approach. At this
well as (iii) improve surface reactions in terms of yield and selectivity by expense only, would such technologies, applied to PFAS or other POPs,
affecting the specific adsorption of contaminants across the surface of may allow for a truly circular economy strategy.
the materials [182]. The dependence of the rate of reaction on the level
of oxygen vacancies was found to be clear for several vacancy-induced 6. Conclusions
materials for PFOA degradation [184–186]. A challenge with
oxygen-vacancy engineering however lies with the relatively short span This review critically analyzed the core challenges such as efficiency,
stability of these defects. Oxidation reactions may occur over time and time consumption, influence of effluent and technical properties and
routes to stabilize the vacancies were demonstrated through specific their limitations thereof, related to the development of some of the most
metal doping of the microstructure of the semiconductors [187,188]. important water remediation techniques related to PFAS compounds’
The development of composite nano-structured photocatalysts for capture, concentration, and degradation (Table 7). The efficiency of
both slurry reactors and surface reactors has allowed for a further in­ these remediation strategies is primarily hampered by the low concen­
crease in photo-electron generation and recombination. Graphene oxide tration of PFAS compounds, the complexity of the source waters matrix,
and its derivatives were used on several occasions as support for and new technologies to selectively concentrate PFAS are required as

19
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

pretreatment steps prior to PFAS degradation. Tandem, combinatorial, [9] J. Roth, I. Abusallout, T. Hill, C. Holton, U. Thapa, D. Hanigan, Release of volatile
per- and polyfluoroalkyl substances from aqueous film-forming foam, Environ.
or hybrid technologies, developed as trains of sequential treatment ap­
Sci. Technol. Lett. 7 (2020) 164–170, https://doi.org/10.1021/acs.
proaches are very likely one of the most promising ways forward to estlett.0c00052.
increase the efficiency of the remediation steps by reducing the need of [10] J. Fu, Y. Gao, L. Cui, T. Wang, Y. Liang, G. Qu, B. Yuan, Y. Wang, A. Zhang,
pretreatment, operational durations and costs, and to limit the release of G. Jiang, Occurrence, temporal trends, and half-lives of perfluoroalkyl acids
(PFAAs) in occupational workers in China, Sci. Rep. 6 (2016) 38039, https://doi.
toxic by-products into the environment. org/10.1038/srep38039.
[11] U.S. EPA, EPA’s per- and polyfluoroalkyl substances (PFAS) Action Plan. http
s://www.epa.gov/sites/production/files/2019-02/documents/pfas_action_pla
CRediT authorship contribution statement n_021319_508compliant_1.pdf (Accessed 25 April 2021).
[12] S. Garg, P. Kumar, V. Mishra, R. Guijt, P. Singh, L.F. Dumée, R.S. Sharma,
Shafali Garg: Conceptualization, Methodology, Writing - original A review on the sources, occurrence and health risks of per-/poly-fluoroalkyl
substances (PFAS) arising from the manufacture and disposal of electric and
draft. Jingshi Wang: Data curation, Writing - review & editing. Pankaj
electronic products, J. Water Process Eng. 38 (2020), 101683, https://doi.org/
Kumar: Visualization, Data curation. Hassan Arafat: Visualization, 10.1016/j.jwpe.2020.101683.
Data curation. Vandana Mishra: Supervision, Writing - review & edit­ [13] K.H. Kucharzyk, R. Darlington, M. Benotti, R. Deeb, E. Hawley, Novel treatment
ing. Radhey Sharma: Supervision, Writing - review & editing. Ludovic technologies for PFAS compounds: a critical review, J. Environ. Manag. 204
(2017) 757–764, https://doi.org/10.1016/j.jenvman.2017.08.016.
Dumee: Supervision, Validation, Writing - review & editing. [14] R. Mahinroosta, L. Senevirathna, A review of the emerging treatment
technologies for PFAS contaminated soils, J. Environ. Manag. 255 (2020),
109896, https://doi.org/10.1016/j.jenvman.2019.109896.
Declaration of Competing Interest [15] J. Cui, P. Gao, Y. Deng, Destruction of per- and polyfluoroalkyl substances (PFAS)
with advanced reduction processes (ARPs): a critical review, Environ. Sci.
Technol. 54 (2020) 3752–3766, https://doi.org/10.1021/acs.est.9b05565.
The authors declare that they have no known competing financial [16] M. Trojanowicz, A. Bojanowska-Czajka, I. Bartosiewicz, K. Kulisa, Advanced
interests or personal relationships that could have appeared to influence oxidation/reduction processes treatment for aqueous perfluorooctanoate (PFOA)
the work reported in this paper. and perfluorooctanesulfonate (PFOS) – a review of recent advances, Chem. Eng.
J. 336 (2018) 170–199, https://doi.org/10.1016/j.cej.2017.10.153.
[17] D.Q. Zhang, W.L. Zhang, Y.N. Liang, Adsorption of perfluoroalkyl and
Acknowledgments polyfluoroalkyl substances (PFASs) from aqueous solution – a review, Sci. Total
Environ. 694 (2019), 133606, https://doi.org/10.1016/j.scitotenv.2019.133606.
[18] W. Zhang, D. Zhang, Y. Liang, Nanotechnology in remediation of water
A/Prof. Ludovic Dumee acknowledges the Australian Research contaminated by poly- and perfluoroalkyl substances: a review, Environ. Pollut.
Council (ARC) for his Discovery Early Career Research Award (DECRA) 247 (2019) 266–276, https://doi.org/10.1016/j.envpol.2019.01.045.
[19] J. Liu, S.M. Avendaño, Microbial degradation of polyfluoroalkyl chemicals in the
2018. LFD acknowledges the support from Khalifa University through
environment: a review, Environ. Int. 61 (2013) 98–114, https://doi.org/10.1016/
project RC2-2019-007. The authors acknowledge the support of Deakin j.envint.2013.08.022.
University and the Micro-Nano Group (Prof. Lingxue Kong). The support [20] F. Li, J. Duan, S. Tian, H. Ji, Y. Zhu, Z. Wei, D. Zhao, Short-chain per- and
extended by the Ministry of Education, Government of India to the Delhi polyfluoroalkyl substances in aquatic systems: occurrence, impacts and
treatment, Chem. Eng. J. 380 (2020), 122506, https://doi.org/10.1016/j.
School of Climate Change & Sustainability, Institute of Eminence, Uni­ cej.2019.122506.
versity of Delhi is duly acknowledged by RSS and VM. SG and PK extend [21] N. Merino, Y. Qu, R.A. Deeb, E.L. Hawley, M.R. Hoffmann, S. Mahendra,
thanks to University Grants Commission for Junior Research Fellowship Degradation and removal methods for perfluoroalkyl and polyfluoroalkyl
substances in water, Environ. Eng. Sci. 33 (2016) 615–649, https://doi.org/
and University of Delhi. 10.1089/ees.2016.0233.
[22] D. Lu, S. Sha, J. Luo, Z. Huang, X. Zhang Jackie, Treatment train approaches for
the remediation of per- and polyfluoroalkyl substances (PFAS): a critical review,
Appendix A. Supporting information
J. Hazard. Mater. 386 (2020), 121963, https://doi.org/10.1016/j.
jhazmat.2019.121963.
Supplementary data associated with this article can be found in the [23] O.S. Arvaniti, A.S. Stasinakis, Review on the occurrence, fate and removal of
perfluorinated compounds during wastewater treatment, Sci. Total Environ., 524-
online version at doi:10.1016/j.jece.2021.105784.
525 (2015) 81–92, https://doi.org/10.1016/j.scitotenv.2015.04.023.
[24] T.L. Coggan, D. Moodie, A. Kolobaric, D. Szabo, J. Shimeta, N.D. Crosbie, E. Lee,
References M. Fernandes, B.O. Clarke, An investigation into per- and polyfluoroalkyl
substances (PFAS) in nineteen Australian wastewater treatment plants (WWTPs),
Heliyon 5 (2019) 02316, https://doi.org/10.1016/j.heliyon.2019.e02316.
[1] K. Kannan, Perfluoroalkyl and polyfluoroalkyl substances: current and future
[25] S. Chen, Y. Zhou, J. Meng, T. Wang, Seasonal and annual variations in removal
perspectives, Environ. Chem. 8 (2011) 333–338, https://doi.org/10.1071/
efficiency of perfluoroalkyl substances by different wastewater treatment
EN11053.
processes, Environ. Pollut. 242 (2018) 2059–2067, https://doi.org/10.1016/j.
[2] M. Kotthoff, J. Müller, H. Jürling, M. Schlummer, D. Fiedler, Perfluoroalkyl and
envpol.2018.06.078.
polyfluoroalkyl substances in consumer products, Environ. Sci. Pollut. Res. 22
[26] M.G. Kibambe, M.N.B. Momba, A.P. Daso, M.A.A. Coetzee, Evaluation of the
(2015) 14546–14559, https://doi.org/10.1007/s11356-015-4202-7.
efficiency of selected wastewater treatment processes in removing selected
[3] C. Lau, K. Anitole, C. Hodes, D. Lai, A. Pfahles-Hutchens, J. Seed, Perfluoroalkyl
perfluoroalkyl substances (PFASs), J. Environ. Manag. 255 (2020), 109945,
acids: a review of monitoring and toxicological findings, Toxicol. Sci. 99 (2007)
https://doi.org/10.1016/j.jenvman.2019.109945.
366–394, https://doi.org/10.1093/toxsci/kfm128.
[27] Y. Gao, S. Deng, Z. Du, K. Liu, G. Yu, Adsorptive removal of emerging
[4] V. Gellrich, T. Stahl, T.P. Knepper, Behavior of perfluorinated compounds in soils
polyfluoroalky substances F-53B and PFOS by anion-exchange resin: a
during leaching experiments, Chemosphere 87 (2012) 1052–1056, https://doi.
comparative study, J. Hazard. Mater. 323 (2017) 550–557, https://doi.org/
org/10.1016/j.chemosphere.2012.02.011.
10.1016/j.jhazmat.2016.04.069.
[5] F. Ebrahimi, A.J. Lewis, C.M. Sales, R. Suri, E.R. McKenzie, Linking PFAS
[28] M.L. Brusseau, The influence of molecular structure on the adsorption of PFAS to
partitioning behavior in sewage solids to the solid characteristics, solution
fluid-fluid interfaces: Using QSPR to predict interfacial adsorption coefficients,
chemistry, and treatment processes, Chemosphere 271 (2021), 129530, https://
Water Res. 152 (2019) 148–158, https://doi.org/10.1016/j.watres.2018.12.057.
doi.org/10.1016/j.chemosphere.2020.129530.
[29] L. Liu, Y. Liu, B. Gao, R. Ji, C. Li, S. Wang, Removal of perfluorooctanoic acid
[6] T.M.H. Nguyen, J. Bräunig, K. Thompson, J. Thompson, S. Kabiri, D.A. Navarro,
(PFOA) and perfluorooctane sulfonate (PFOS) from water by carbonaceous
R.S. Kookana, C. Grimison, C.M. Barnes, C.P. Higgins, M.J. McLaughlin, J.
nanomaterials: a review, Crit. Rev. Environ. Sci. Technol. 50 (2020) 2379–2414,
F. Mueller, Influences of chemical properties, soil properties, and solution pH on
https://doi.org/10.1080/10643389.2019.1700751.
soil–water partitioning coefficients of per- and polyfluoroalkyl substances
[30] D.Q. Zhang, W.L. Zhang, Y.N. Liang, Adsorption of perfluoroalkyl and
(PFASs), Environ. Sci. Technol. 54 (2020) 15883–15892, https://doi.org/
polyfluoroalkyl substances (PFASs) from aqueous solution – a review, Sci. Total
10.1021/acs.est.0c05705.
Environ. 694 (2019), 133606, https://doi.org/10.1016/j.scitotenv.2019.133606.
[7] S. Joudan, J.J. Orlando, G.S. Tyndall, T.C. Furlani, C.J. Young, S.A. Mabury,
[31] M. Sörengård, E. Östblom, S. Köhler, L. Ahrens, Adsorption behavior of per- and
Atmospheric fate of a new polyfluoroalkyl building block,
polyfluoralkyl substances (PFASs) to 44 inorganic and organic sorbents and use of
C3F7OCHFCF2SCH2CH2OH, Environ. Sci. Technol. (2021), https://doi.org/
dyes as proxies for PFAS sorption, J. Environ. Chem. Eng. 8 (2020), 103744,
10.1021/acs.est.0c07584.
https://doi.org/10.1016/j.jece.2020.103744.
[8] Q. Wang, Y. Ruan, H. Lin, P.K.S. Lam, Review on perfluoroalkyl and
[32] D.P. Siriwardena, M. Crimi, T.M. Holsen, C. Bellona, C. Divine, E. Dickenson,
polyfluoroalkyl substances (PFASs) in the Chinese atmospheric environment, Sci.
Changes in adsorption behavior of perfluorooctanoic acid and
Total Environ. 737 (2020), 139804, https://doi.org/10.1016/j.
perfluorohexanesulfonic acid through chemically-facilitated surface modification
scitotenv.2020.139804.

20
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

of granular activated carbon, Environ. Eng. Sci. 36 (2019) 453–465, https://doi. [54] Y.H. Aly, D.P. McInnis, S.M. Lombardo, W.A. Arnold, K.D. Pennell, J. Hatton, M.
org/10.1089/ees.2018.0319. F. Simcik, Enhanced adsorption of perfluoro alkyl substances for in situ
[33] N.A. Lundquist, M.J. Sweetman, K.R. Scroggie, M.J.H. Worthington, L.J. Esdaile, remediation, Environ. Sci.: Water Res. Technol. 5 (2019) 1867–1875, https://doi.
S.F.K. Alboaiji, S.E. Plush, J.D. Hayball, J.M. Chalker, Polymer supported carbon org/10.1039/C9EW00426B.
for safe and effective remediation of PFOA- and PFOS-contaminated water, ACS [55] T.D. Appleman, C.P. Higgins, O. Quiñones, B.J. Vanderford, C. Kolstad, J.
Sustain. Chem. Eng. 7 (2019) 11044–11049, https://doi.org/10.1021/ C. Zeigler-Holady, E.R.V. Dickenson, Treatment of poly- and perfluoroalkyl
acssuschemeng.9b01793. substances in U.S. full-scale water treatment systems, Water Res. 51 (2014)
[34] S.T.M.L.D. Senevirathna, Development of effective removal methods of PFCs 246–255, https://doi.org/10.1016/J.WATRES.2013.10.067.
(perfluorinated compounds) in water by adsorption and coagulation, Kyoto [56] J. Thompson, G. Eaglesham, J. Reungoat, Y. Poussade, M. Bartkow, M. Lawrence,
University, 2010. https://repository.kulib.kyoto-u.ac.jp/dspace/bitstream/2433/ J.F. Mueller, Removal of PFOS, PFOA and other perfluoroalkyl acids at water
126798/2/D_SENEVIRATHNA_THENNAKOON_MUDIYANSELAGE_LALANTHA_ reclamation plants in South East Queensland Australia, Chemosphere 82 (2011)
DHARSHANA_SENEVIRATHNA.pdf (Accessed 30 December 2020). 9–17, https://doi.org/10.1016/j.chemosphere.2010.10.040.
[35] Q. Yu, R. Zhang, S. Deng, J. Huang, G. Yu, Sorption of perfluorooctane sulfonate [57] B. Yang, Y. Han, Y. Deng, Y. Li, Q. Zhuo, J. Wu, Highly efficient removal of
and perfluorooctanoate on activated carbons and resin: kinetic and isotherm perfluorooctanoic acid from aqueous solution by H2O2-enhanced
study, Water Res. 43 (2009) 1150–1158, https://doi.org/10.1016/j. electrocoagulation-electroflotation technique, Emerg. Contam. 2 (2016) 49–55,
watres.2008.12.001. https://doi.org/10.1016/j.emcon.2016.04.001.
[36] L. Laura del Moral, Y.J. Choi, T.H. Boyer, Comparative removal of Suwannee [58] J. Bao, W.-J. Yu, Y. Liu, X. Wang, Z.-Q. Liu, Y.-F. Duan, Removal of
River natural organic matter and perfluoroalkyl acids by anion exchange: impact perfluoroalkanesulfonic acids (PFSAs) from synthetic and natural groundwater by
of polymer composition and mobile counterion, Water Res. 178 (2020), 115846, electrocoagulation, Chemosphere 248 (2020), 125951, https://doi.org/10.1016/
https://doi.org/10.1016/j.watres.2020.115846. j.chemosphere.2020.125951.
[37] A. Zaggia, L. Conte, L. Falletti, M. Fant, A. Chiorboli, Use of strong anion [59] L.F. Greenlee, D.F. Lawler, B.D. Freeman, B. Marrot, P. Moulin, Reverse osmosis
exchange resins for the removal of perfluoroalkylated substances from desalination: water sources, technology, and today’s challenges, Water Res. 43
contaminated drinking water in batch and continuous pilot plants, Water Res. 91 (2009) 2317–2348, https://doi.org/10.1016/j.watres.2009.03.010.
(2016) 137–146, https://doi.org/10.1016/j.watres.2015.12.039. [60] P. Kumari, N. Bahadur, L.F. Dumée, Photo-catalytic membrane reactors for the
[38] M. Ateia, M. Arifuzzaman, S. Pellizzeri, M.F. Attia, N. Tharayil, J.N. Anker, remediation of persistent organic pollutants – a review, Sep. Purif. Technol. 230
T. Karanfil, Cationic polymer for selective removal of GenX and short-chain PFAS (2020), 115878, https://doi.org/10.1016/j.seppur.2019.115878.
from surface waters and wastewaters at ng/L levels, Water Res. 163 (2019), [61] Y.-T. Tsai, A. Yu-Chen Lin, Y.-H. Weng, K.-C. Li, Treatment of perfluorinated
114874, https://doi.org/10.1016/j.watres.2019.114874. chemicals by electro-microfiltration, Environ. Sci. Technol. 44 (2010)
[39] F. Liu, L. Hua, W. Zhang, Influences of microwave irradiation on performances of 7914–7920, https://doi.org/10.1021/es101964y.
membrane filtration and catalytic degradation of perfluorooctanoic acid (PFOA), [62] X. Chen, A. Vanangamudi, J. Wang, J. Jegatheesan, V. Mishra, R. Sharma, S.
Environ. Int. 143 (2020), 105969, https://doi.org/10.1016/j. R. Gray, J. Kujawa, W. Kujawski, F. Wicaksana, L.F. Dumée, Direct contact
envint.2020.105969. membrane distillation for effective concentration of perfluoroalkyl substances –
[40] P. Meng, S. Deng, X. Lu, Z. Du, B. Wang, J. Huang, Y. Wang, G. Yu, B. Xing, Role impact of surface fouling and material stability, Water Res. 182 (2020), 116010,
of air bubbles overlooked in the adsorption of perfluorooctanesulfonate on https://doi.org/10.1016/j.watres.2020.116010.
hydrophobic carbonaceous adsorbents, Environ. Sci. Technol. 48 (2014) [63] D. Banks, B.-M. Jun, J. Heo, N. Her, C.M. Park, Y. Yoon, Selected advanced water
13785–13792, https://doi.org/10.1021/es504108u. treatment technologies for perfluoroalkyl and polyfluoroalkyl substances: a
[41] Z. Niu, Y. Wang, H. Lin, F. Jin, Y. Li, J. Niu, Electrochemically enhanced removal review, Sep. Purif. Technol. 231 (2020), 115929, https://doi.org/10.1016/j.
of perfluorinated compounds (PFCs) from aqueous solution by CNTs-graphene seppur.2019.115929.
composite electrode, Chem. Eng. J. 328 (2017) 228–235, https://doi.org/ [64] F. Li, J. Duan, S. Tian, H. Ji, Y. Zhu, Z. Wei, D. Zhao, Short-chain per- and
10.1016/j.cej.2017.07.033. polyfluoroalkyl substances in aquatic systems: occurrence, impacts and
[42] X. Li, S. Chen, X. Quan, Y. Zhang, Enhanced adsorption of PFOA and PFOS on treatment, Chem. Eng. J. 380 (2020), 122506, https://doi.org/10.1016/j.
multiwalled carbon nanotubes under electrochemical assistance, Environ. Sci. cej.2019.122506.
Technol. 45 (2011) 8498–8505, https://doi.org/10.1021/es202026v. [65] C.Y.Y. Tang, Q.S. Fu, A.P. Robertson, C.S. Criddle, J.O. Leckie, Use of reverse
[43] P.N. Omo-Okoro, A.P. Daso, J.O. Okonkwo, A review of the application of osmosis membranes to remove perfluorooctane sulfonate (PFOS) from
agricultural wastes as precursor materials for the adsorption of per- and semiconductor wastewater, Environ. Sci. Technol. 40 (2006) 7343–7349, https://
polyfluoroalkyl substances: a focus on current approaches and methodologies, doi.org/10.1021/es060831q.
Environ. Technol. Innov. 9 (2018) 100–114, https://doi.org/10.1016/j. [66] A.M. Comerton, R.C. Andrews, D.M. Bagley, The influence of natural organic
eti.2017.11.005. matter and cations on the rejection of endocrine disrupting and pharmaceutically
[44] S. Deng, Q. Zhang, Y. Nie, H. Wei, B. Wang, J. Huang, G. Yu, B. Xing, Sorption active compounds by nanofiltration, Water Res. 43 (2009) 613–622, https://doi.
mechanisms of perfluorinated compounds on carbon nanotubes, Environ. Pollut. org/10.1016/j.watres.2008.11.003.
168 (2012) 138–144, https://doi.org/10.1016/j.envpol.2012.03.048. [67] Y. Yoon, P. Westerhoff, S.A. Snyder, E.C. Wert, Nanofiltration and ultrafiltration
[45] W. Ji, L. Xiao, Y. Ling, C. Ching, M. Matsumoto, R.P. Bisbey, D.E. Helbling, W. of endocrine disrupting compounds, pharmaceuticals and personal care products,
R. Dichtel, Removal of GenX and perfluorinated alkyl substances from water by J. Membr. Sci. 270 (2006) 88–100, https://doi.org/10.1016/j.
amine-functionalized covalent organic frameworks, J. Am. Chem. Soc. 140 memsci.2005.06.045.
(2018) 12677–12681, https://doi.org/10.1021/jacs.8b06958. [68] C.Y. Tang, Q.S. Fu, C.S. Criddle, J.O. Leckie, Effect of flux (transmembrane
[46] Q. Zhang, S. Deng, G. Yu, J. Huang, Removal of perfluorooctane sulfonate from pressure) and membrane properties on fouling and rejection of reverse osmosis
aqueous solution by crosslinked chitosan beads: sorption kinetics and uptake and nanofiltration membranes treating perfluorooctane sulfonate containing
mechanism, Bioresour. Technol. 102 (2011) 2265–2271, https://doi.org/ wastewater, Environ. Sci. Technol. 41 (2007) 2008–2014, https://doi.org/
10.1016/j.biortech.2010.10.040. 10.1021/es062052f.
[47] F. Wang, X. Lu, K.M. Shih, P. Wang, X. Li, Removal of perfluoroalkyl sulfonates [69] Á. Soriano, D. Gorri, A. Urtiaga, Selection of high flux membrane for the effective
(PFAS) from aqueous solution using permanently confined micelle arrays removal of short-chain perfluorocarboxylic acids, Ind. Eng. Chem. Res. 58 (2019)
(PCMAs), Sep. Purif. Technol. 138 (2014) 7–12, https://doi.org/10.1016/j. 3329–3338, https://doi.org/10.1021/acs.iecr.8b05506.
seppur.2014.09.037. [70] C.Y. Tang, T.H. Chong, A.G. Fane, Colloidal interactions and fouling of NF and RO
[48] S. Lath, D.A. Navarro, D. Losic, A. Kumar, M.J. McLaughlin, Sorptive remediation membranes: a review, Adv. Colloid Interface Sci. 164 (2011) 126–143, https://
of perfluorooctanoic acid (PFOA) using mixed mineral and graphene/carbon- doi.org/10.1016/j.cis.2010.10.007.
based materials, J. Environ. Chem. 15 (2018) 472–480, https://doi.org/10.1071/ [71] C. Flores, F. Ventura, J. Martin-Alonso, J. Caixach, Occurrence of perfluorooctane
EN18156. sulfonate (PFOS) and perfluorooctanoate (PFOA) in N.E. Spanish surface waters
[49] R. Zhang, Z. Hu, H. Wei, S. Zhang, X. Meng, Adsorption of perfluorooctane and their removal in a drinking water treatment plant that combines conventional
sulfonate on carbonized poly-melamine-formaldehyde sponge, Sci. Total Environ. and advanced treatments in parallel lines, Sci. Total Environ. 461–462 (2013)
727 (2020), 138626, https://doi.org/10.1016/j.scitotenv.2020.138626. 618–626, https://doi.org/10.1016/j.scitotenv.2013.05.026.
[50] M.S. Hellsing, S. Josefsson, A.V. Hughes, L. Ahrens, Sorption of perfluoroalkyl [72] C.M. Glover, O. Quiñones, E.R.V. Dickenson, Removal of perfluoroalkyl and
substances to two types of minerals, Chemosphere 159 (2016) 385–391, https:// polyfluoroalkyl substances in potable reuse systems, Water Res. 144 (2018)
doi.org/10.1016/j.chemosphere.2016.06.016. 454–461, https://doi.org/10.1016/j.watres.2018.07.018.
[51] Y. Bao, J. Niu, Z. Xu, D. Gao, J. Shi, X. Sun, Q. Huang, Removal of perfluorooctane [73] T.D. Appleman, E.R.V. Dickenson, C. Bellona, C.P. Higgins, Nanofiltration and
sulfonate (PFOS) and perfluorooctanoate (PFOA) from water by coagulation: granular activated carbon treatment of perfluoroalkyl acids, J. Hazard. Mater.
mechanisms and influencing factors, J. Colloid Interface Sci. 434 (2014) 59–64, 260 (2013) 740–746, https://doi.org/10.1016/j.jhazmat.2013.06.033.
https://doi.org/10.1016/j.jcis.2014.07.041. [74] E. Steinle-Darling, M. Reinhard, Nanofiltration for trace organic contaminant
[52] B.K. Pramanik, S.K. Pramanik, F. Suja, A comparative study of coagulation, removal: Structure, solution, and membrane fouling effects on the rejection of
granular- and powdered-activated carbon for the removal of perfluorooctane perfluorochemicals, Environ. Sci. Technol. 42 (2008) 5292–5297, https://doi.
sulfonate and perfluorooctanoate in drinking water treatment, Environ. Technol. org/10.1021/es703207s.
36 (2015) 2610–2617, https://doi.org/10.1080/09593330.2015.1040079. [75] J. Heo, L.K. Boateng, J.R.V. Flora, H. Lee, N. Her, Y.-G. Park, Y. Yoon,
[53] F. Xiao, M.F. Simcik, J.S. Gulliver, Mechanisms for removal of perfluorooctane Comparison of flux behavior and synthetic organic compound removal by
sulfonate (PFOS) and perfluorooctanoate (PFOA) from drinking water by forward osmosis and reverse osmosis membranes, J. Membr. Sci. 443 (2013)
conventional and enhanced coagulation, Water Res. 47 (2013) 49–56, https:// 69–82, https://doi.org/10.1016/j.memsci.2013.04.063.
doi.org/10.1016/j.watres.2012.09.024. [76] C. Zeng, S. Tanaka, Y. Suzuki, S. Fujii, Impact of feed water pH and membrane
material on nanofiltration of perfluorohexanoic acid in aqueous solution,

21
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

Chemosphere 183 (2017) 599–604, https://doi.org/10.1016/j. [99] M.-J. Chen, S.-L. Lo, Y.-C. Lee, J. Kuo, C.-H. Wu, Decomposition of
chemosphere.2017.05.132. perfluorooctanoic acid by ultraviolet light irradiation with Pb-modified titanium
[77] X. Hang, X. Chen, J. Luo, W. Cao, Y. Wan, Removal and recovery of dioxide, J. Hazard. Mater. 303 (2016) 111–118, https://doi.org/10.1016/j.
perfluorooctanoate from wastewater by nanofiltration, Sep. Purif. Technol. 145 jhazmat.2015.10.011.
(2015) 120–129, https://doi.org/10.1016/j.seppur.2015.03.013. [100] S. Taniyasu, N. Yamashita, E. Yamazaki, G. Petrick, K. Kannan, The
[78] R. Reis, L.F. Dumee, L. He, F. She, J.D. Orbell, B. Winther-Jensen, M.C. Duke, environmental photolysis of perfluorooctanesulfonate, perfluorooctanoate, and
Amine enrichment of thin-film composite membranes via low pressure plasma related fluorochemicals, Chemosphere 90 (2013) 1686–1692, https://doi.org/
polymerization for antimicrobial adhesion, ACS Appl. Mater. Interfaces 7 (2015) 10.1016/j.chemosphere.2012.09.065.
14644–14653, https://doi.org/10.1021/acsami.5b01603. [101] H. Hori, Y. Nagano, M. Murayama, K. Koike, S. Kutsuna, Efficient decomposition
[79] C. Bellona, M. Marts, J.E. Drewes, The effect of organic membrane fouling on the of perfluoroether carboxylic acids in water with a combination of persulfate
properties and rejection characteristics of nanofiltration membranes, Sep. Purif. oxidant and ultrasonic irradiation, J. Fluor. Chem. 141 (2012) 5–10, https://doi.
Technol. 74 (2010) 44–54, https://doi.org/10.1016/j.seppur.2010.05.006. org/10.1016/j.jfluchem.2012.05.012.
[80] C. Boo, Y. Wang, I. Zucker, Y. Choo, C.O. Osuji, M. Elimelech, High performance [102] H. Hori, E. Hayakawa, H. Einaga, S. Kutsuna, K. Koike, T. Ibusuki, H. Kiatagawa,
nanofiltration membrane for effective removal of perfluoroalkyl substances at R. Arakawa, Decomposition of environmentally persistent perfluorooctanoic acid
high water recovery, Environ. Sci. Technol. 52 (2018) 7279–7288, https://doi. in water by photochemical approaches, Environ. Sci. Technol. 38 (2004)
org/10.1021/acs.est.8b01040. 6118–6124, https://doi.org/10.1021/es049719n.
[81] C. Xu, H. Chen, F. Jiang, Adsorption of perflourooctane sulfate (PFOS) and [103] S.M. Vijayan, T. Göen, K. Dennerlein, R.E. Horch, I. Ludolph, H. Drexler, S. Kilo,
perfluorooctanoate (PFOA) on polyaniline nanotubes, Colloids Surf. A Calcium, magnesium and aluminium ions as decontaminating agents against
Physicochem. Eng. Asp. 479 (2015) 60–67, https://doi.org/10.1016/j. dermal fluoride absorption following hydrofluoric acid exposure, Toxicol. Vitr. 71
colsurfa.2015.03.045. (2021), 105055, https://doi.org/10.1016/j.tiv.2020.105055.
[82] Y. Li, D.P. Oliver, R.S. Kookana, A critical analysis of published data to discern the [104] B. Xu, J.L. Zhou, A. Altaee, M.B. Ahmed, M.A.H. Johir, J. Ren, X. Li, Improved
role of soil and sediment properties in determining sorption of per and photocatalysis of perfluorooctanoic acid in water and wastewater by Ga2O3/UV
polyfluoroalkyl substances (PFASs), Sci. Total Environ. 628–629 (2018) 110–120, system assisted by peroxymonosulfate, Chemosphere 239 (2020), 124722,
https://doi.org/10.1016/j.scitotenv.2018.01.167. https://doi.org/10.1016/j.chemosphere.2019.124722.
[83] F.J. Real, F.J. Benitez, J.L. Acero, F. Casas, Adsorption of selected emerging [105] H. Jin, C. Guo, X. Liu, J. Liu, A. Vasileff, Y. Jiao, Y. Zheng, S.-Z. Qiao, Emerging
contaminants onto PAC and GAC: equilibrium isotherms, kinetics, and effect of two-dimensional nanomaterials for electrocatalysis, Chem. Rev. 118 (2018)
the water matrix, J. Environ. Sci. Health, Part A 52 (2017) 727–734, https://doi. 6337–6408, https://doi.org/10.1021/acs.chemrev.7b00689.
org/10.1080/10934529.2017.1301751. [106] D. Li, J. Qu, The progress of catalytic technologies in water purification: a review,
[84] M.B. Ahmed, M.M. Alam, J.L. Zhou, B. Xu, M.A.H. Johir, A.K. Karmakar, M. J. Environ. Sci. 21 (2009) 713–719, https://doi.org/10.1016/S1001-0742(08)
S. Rahman, J. Hossen, A.T.M.K. Hasan, M.A. Moni, Advanced treatment 62329-3.
technologies efficacies and mechanism of per- and poly-fluoroalkyl substances [107] Z. Li, S. Li, Y. Tang, X. Li, J. Wang, L. Li, Highly efficient degradation of
removal from water, Process Saf. Environ. Prot. 136 (2020) 1–14, https://doi. perfluorooctanoic acid: an integrated photo-electrocatalytic ozonation and
org/10.1016/j.psep.2020.01.005. mechanism study, Chem. Eng. J. 391 (2020), 123533, https://doi.org/10.1016/j.
[85] B.R. Shivakoti, S. Fujii, M. Nozoe, S. Tanaka, C. Kunacheva, Perfluorinated cej.2019.123533.
chemicals (PFCs) in water purification plants (WPPs) with advanced treatment [108] Q. Wang, M. Liu, H. Zhao, Y. Chen, F. Xiao, W. Chu, G. Zhao, Efficiently
processes, Water Supply 10 (2010) 87–95, https://doi.org/10.2166/ degradation of perfluorooctanoic acid in synergic electrochemical process
ws.2010.707. combining cathodic electro-Fenton and anodic oxidation, Chem. Eng. J. 378
[86] S. Takagi, F. Adachi, K. Miyano, Y. Koizumi, H. Tanaka, I. Watanabe, S. Tanabe, (2019), 122071, https://doi.org/10.1016/j.cej.2019.122071.
K. Kannan, Fate of perfluorooctanesulfonate and perfluorooctanoate in drinking [109] Y. Liu, S. Chen, X. Quan, H. Yu, H. Zhao, Y. Zhang, Efficient mineralization of
water treatment processes, Water Res. 45 (2011) 3925–3932, https://doi.org/ perfluorooctanoate by electro-fenton with H2O2 electro-generated on
10.1016/j.watres.2011.04.052. hierarchically porous carbon, Environ. Sci. Technol. 49 (2015) 13528–13533,
[87] V.A. Arias Espana, M. Mallavarapu, R. Naidu, Treatment technologies for aqueous https://doi.org/10.1021/acs.est.5b03147.
perfluorooctanesulfonate (PFOS) and perfluorooctanoate (PFOA): A critical [110] W.-W. Liu, R.-F. Peng, Recent advances of bismuth oxychloride photocatalytic
review with an emphasis on field testing, Environ. Technol. Innov. 4 (2015) material: property, preparation and performance enhancement, J. Electron. Sci.
168–181, https://doi.org/10.1016/j.eti.2015.06.001. Technol. 18 (2020), 100020, https://doi.org/10.1016/j.jnlest.2020.100020.
[88] Á. Soriano, D. Gorri, A. Urtiaga, Efficient treatment of perfluorohexanoic acid by [111] P.M. Dombrowski, P. Kakarla, W. Caldicott, Y. Chin, V. Sadeghi, D. Bogdan,
nanofiltration followed by electrochemical degradation of the NF concentrate, F. Barajas-Rodriguez, S.Y.D. Chiang, Technology review and evaluation of
Water Res. 112 (2017) 147–156, https://doi.org/10.1016/j.watres.2017.01.043. different chemical oxidation conditions on treatability of PFAS, Remediation 28
[89] A. Boonya-Atichart, S.K. Boontanon, N. Boontanon, Study of hybrid membrane (2018) 135–150, https://doi.org/10.1002/rem.21555.
filtration and photocatalysis for removal of perfluorooctanoic acid (PFOA) in [112] P. Wardman, Reduction potentials of one-electron couples involving free radicals
groundwater, Water Sci. Technol. 2017 (2018) 561–569, https://doi.org/ in aqueous solution, J. Phys. Chem. Ref. Data 18 (1989) 1637–1755, https://doi.
10.2166/wst.2018.178. org/10.1063/1.555843.
[90] C. Dang, F. Sun, H. Jiang, T. Huang, W. Liu, X. Chen, H. Ji, Pre-accumulation and [113] American Water Works Association, Perfluorinated compounds treatment and
in-situ destruction of diclofenac by a photo-regenerable activated carbon fiber removal. https://www.awwa.org/Portals/0/AWWA/Programs/AWWAPFCFact
supported titanate nanotubes composite material: Intermediates, DFT calculation, SheetTreatmentandRemoval.pdf (Accessed 30 December 2020).
and ecotoxicity, J. Hazard. Mater. 400 (2020), 123225, https://doi.org/10.1016/ [114] H. Hori, Y. Nagaoka, M. Murayama, S. Kutsuna, Efficient decomposition of
j.jhazmat.2020.123225. perfluorocarboxylic acids and alternative fluorochemical surfactants in hot water,
[91] R.R. Giri, H. Ozaki, T. Morigaki, S. Taniguchi, R. Takanami, UV photolysis of Environ. Sci. Technol. 42 (2008) 7438–7443, https://doi.org/10.1021/
perfluorooctanoic acid (PFOA) in dilute aqueous solution, Water Sci. Technol. 63 es800832p.
(2011) 276–282, https://doi.org/10.2166/wst.2011.050. [115] A.Y.-C. Lin, S.C. Panchangam, C.-Y. Chang, P.K.A. Hong, H.-F. Hsueh, Removal of
[92] Z. Li, P. Zhang, T. Shao, J. Wang, L. Jin, X. Li, Different nanostructured In2O3 for perfluorooctanoic acid and perfluorooctane sulfonate via ozonation under
photocatalytic decomposition of perfluorooctanoic acid (PFOA), J. Hazard. alkaline condition, J. Hazard. Mater. 243 (2012) 272–277, https://doi.org/
Mater. 260 (2013) 40–46, https://doi.org/10.1016/j.jhazmat.2013.04.042. 10.1016/j.jhazmat.2012.10.029.
[93] B. Gomez-Ruiz, P. Ribao, N. Diban, M.J. Rivero, I. Ortiz, A. Urtiaga, [116] C.S. Liu, K. Shih, F. Wang, Oxidative decomposition of perfluorooctanesulfonate
Photocatalytic degradation and mineralization of perfluorooctanoic acid (PFOA) in water by permanganate, Sep. Purif. Technol. 87 (2012) 95–100, https://doi.
using a composite TiO2− rGO catalyst, J. Hazard. Mater. 344 (2018) 950–957, org/10.1016/j.seppur.2011.11.027.
https://doi.org/10.1016/j.jhazmat.2017.11.048. [117] L. Yang, L. He, J. Xue, Y. Ma, Z. Xie, L. Wu, M. Huang, Z. Zhang, Persulfate-based
[94] C.A. Martín, M.A. Baltanás, A.E. Cassano, Photocatalytic decomposition of degradation of perfluorooctanoic acid (PFOA) and perfluorooctane sulfonate
chloroform in a fully irradiated heterogeneous photoreactor using titanium oxide (PFOS) in aqueous solution: Review on influences, mechanisms and prospective,
particulate suspensions, Catal. Today 27 (1996) 221–227, https://doi.org/ J. Hazard. Mater. 393 (2020), 122405, https://doi.org/10.1016/j.
10.1016/0920-5861(95)00191-3. jhazmat.2020.122405.
[95] Y.-C. Chen, S.-L. Lo, J. Kuo, Effects of titanate nanotubes synthesized by a [118] S. Park, L.S. Lee, V.F. Medina, A. Zull, S. Waisner, Heat-activated persulfate
microwave hydrothermal method on photocatalytic decomposition of oxidation of PFOA, 6:2 fluorotelomer sulfonate, and PFOS under conditions
perfluorooctanoic acid, Water Res. 45 (2011) 4131–4140, https://doi.org/ suitable for in-situ groundwater remediation, Chemosphere 145 (2016) 376–383,
10.1016/j.watres.2011.05.020. https://doi.org/10.1016/j.chemosphere.2015.11.097.
[96] X. Li, P. Zhang, L. Jin, T. Shao, Z. Li, J. Cao, Efficient photocatalytic [119] T.A. Bruton, D.L. Sedlak, Treatment of perfluoroalkyl acids by heat-activated
decomposition of perfluorooctanoic acid by indium oxide and its mechanism, persulfate under conditions representative of in situ chemical oxidation,
Environ. Sci. Technol. 46 (2012) 5528–5534, https://doi.org/10.1021/ Chemosphere 206 (2018) 457–464, https://doi.org/10.1016/j.
es204279u. chemosphere.2018.04.128.
[97] B. Zhao, M. Lv, L. Zhou, Photocatalytic degradation of perfluorooctanoic acid [120] E.F. Houtz, D.L. Sedlak, Oxidative conversion as a means of detecting precursors
with beta-Ga2O3 in anoxic aqueous solution, J. Environ. Sci. 24 (2012) 774–780, to perfluoroalkyl acids in urban runoff, Environ. Sci. Technol. 46 (2012)
https://doi.org/10.1016/s1001-0742(11)60818-8. 9342–9349, https://doi.org/10.1021/es302274g.
[98] X. Pan, M.-Q. Yang, X. Fu, N. Zhang, Y.-J. Xu, Defective TiO2 with oxygen [121] Y. Bao, S. Deng, X. Jiang, Y. Qu, Y. He, L. Liu, Q. Chai, M. Mumtaz, J. Huang,
vacancies: Synthesis, properties and photocatalytic applications, Nanoscale 5 G. Cagnetta, G. Yu, Degradation of PFOA substitute: GenX (HFPO–DA ammonium
(2013) 3601–3614, https://doi.org/10.1039/C3NR00476G. salt): oxidation with UV/persulfate or reduction with UV/sulfite? Environ. Sci.
Technol. 52 (2018) 11728–11734, https://doi.org/10.1021/acs.est.8b02172.

22
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

[122] Y. Qu, C. Zhang, F. Li, J. Chen, Q. Zhou, Photo-reductive defluorination of [146] D. Zhang, W. Zhang, Y. Liang, Bacterial community in a freshwater pond
perfluorooctanoic acid in water, Water Res. 44 (2010) 2939–2947, https://doi. responding to the presence of perfluorooctanoic acid (PFOA), Environ. Technol.
org/10.1016/j.watres.2010.02.019. 41 (2020) 3646–3656, https://doi.org/10.1080/09593330.2019.1616828.
[123] V. Ochoa-Herrera, R. Sierra-Alvarez, A. Somogyi, N.E. Jacobsen, V.H. Wysocki, J. [147] S. Huang, P.R. Jaffé, Defluorination of perfluorooctanoic acid (PFOA) and
A. Field, Reductive defluorination of perfluorooctane sulfonate, Environ. Sci. perfluorooctane sulfonate (PFOS) by acidimicrobium sp. strain A6, Environ. Sci.
Technol. 42 (2008) 3260–3264, https://doi.org/10.1021/es702842q. Technol. 53 (2019) 11410–11419, https://doi.org/10.1021/acs.est.9b04047.
[124] X. Dai, Z. Xie, B. Dorian, S. Gray, J. Zhang, Comparative study of PFAS treatment [148] Y. Tingru, Occurrence and biodegradation study of perfluoroalkyl and
by UV, UV/ozone, and fractionations with air and ozonated air, Environ. Sci. polyfluroalkyl substances in a tropical constructed wetland, National University
Water Res. Technol. 5 (2019) 1897–1907, https://doi.org/10.1039/ of Singapore, 2018. https://scholarbank.nus.edu.sg/handle/10635/150335
C9EW00701F. (Accessed 25 April 2021).
[125] J. Horst, J. McDonough, I. Ross, M. Dickson, J. Miles, J. Hurst, P. Storch, Water [149] M. Sáez, P. de Voogt, J.R. Parsons, Persistence of perfluoroalkylated substances in
treatment technologies for PFAS: the next generation, Groundw. Monit. closed bottle tests with municipal sewage sludge, Environ. Sci. Pollut. Res. 15
Remediat. 38 (2018) 13–23, https://doi.org/10.1111/gwmr.12281. (2008) 472–477, https://doi.org/10.1007/s11356-008-0020-5.
[126] H. Cao, W. Zhang, C. Wang, Y. Liang, Sonochemical degradation of poly- and [150] D.M. O’Carroll, T.C. Jeffries, M.J. Lee, S.T. Le, A. Yeung, S. Wallace, N. Battye, D.
perfluoroalkyl substances – a review, Ultrason. Sonochem. 69 (2020), 105245, J. Patch, M.J. Manefield, K.P. Weber, Developing a roadmap to determine per-
https://doi.org/10.1016/j.ultsonch.2020.105245. and polyfluoroalkyl substances-microbial population interactions, Sci. Total
[127] T. Campbell, M.R. Hoffmann, Sonochemical degradation of perfluorinated Environ. 712 (2020), 135994, https://doi.org/10.1016/j.scitotenv.2019.135994.
surfactants: power and multiple frequency effects, Sep. Purif. Technol. 156 (2015) [151] V. Ochoa-Herrera, J.A. Field, A. Luna-Velasco, R. Sierra-Alvarez, Microbial
1019–1027, https://doi.org/10.1016/j.seppur.2015.09.053. toxicity and biodegradability of perfluorooctane sulfonate (PFOS) and shorter
[128] T.Y. Campbell, C.D. Vecitis, B.T. Mader, M.R. Hoffmann, Perfluorinated chain perfluoroalkyl and polyfluoroalkyl substances (PFASs), Environ. Sci.:
surfactant chain-length effects on sonochemical kinetics, J. Phys. Chem. A 113 Process. Impacts 18 (2016) 1236–1246, https://doi.org/10.1039/C6EM00366D.
(2009) 9834–9842, https://doi.org/10.1021/jp903003w. [152] N.J.M. Fitzgerald, H.R. Temme, M.F. Simcik, P.J. Novak, Aqueous film forming
[129] V.L. Gole, A. Fishgold, R. Sierra-Alvarez, P. Deymier, M. Keswani, Treatment of foam and associated perfluoroalkyl substances inhibit methane production and
perfluorooctane sulfonic acid (PFOS) using a large-scale sonochemical reactor, co-contaminant degradation in an anaerobic microbial community, Environ. Sci.
Sep. Purif. Technol. 194 (2018) 104–110, https://doi.org/10.1016/j. Process. Impacts 21 (2019) 1915–1925, https://doi.org/10.1039/C9EM00241C.
seppur.2017.11.009. [153] W. Wang, G. Rhodes, J. Ge, X. Yu, H. Li, Uptake and accumulation of per- and
[130] V.L. Gole, R. Sierra-Alvarez, H. Peng, J.P. Giesy, P. Deymier, M. Keswani, Sono- polyfluoroalkyl substances in plants, Chemosphere 261 (2020), 127584, https://
chemical treatment of per- and poly-fluoroalkyl compounds in aqueous film- doi.org/10.1016/j.chemosphere.2020.127584.
forming foams by use of a large-scale multi-transducer dual-frequency based [154] W. Zhang, D. Zhang, D.V. Zagorevski, Y. Liang, Exposure of Juncus effusus to seven
acoustic reactor, Ultrason. Sonochem. 45 (2018) 213–222, https://doi.org/ perfluoroalkyl acids: Uptake, accumulation and phytotoxicity, Chemosphere 233
10.1016/j.ultsonch.2018.02.014. (2019) 300–308, https://doi.org/10.1016/j.chemosphere.2019.05.258.
[131] L. Rodriguez-Freire, N. Abad-Fernández, R. Sierra-Alvarez, C. Hoppe-Jones, [155] S. Zhao, T. Zhou, B. Wang, L. Zhu, M. Chen, D. Li, L. Yang, Different
H. Peng, J.P. Giesy, S. Snyder, M. Keswani, Sonochemical degradation of biotransformation behaviors of perfluorooctane sulfonamide in wheat (Triticum
perfluorinated chemicals in aqueous film-forming foams, J. Hazard. Mater. 317 aestivum L.) from earthworms (Eisenia fetida), J. Hazard. Mater. 346 (2018)
(2016) 275–283, https://doi.org/10.1016/j.jhazmat.2016.05.078. 191–198, https://doi.org/10.1016/j.jhazmat.2017.12.018.
[132] L. Rodriguez-Freire, R. Balachandran, R. Sierra-Alvarez, M. Keswani, Effect of [156] T.-T. Wang, G.-G. Ying, L.-Y. He, Y.-S. Liu, J.-L. Zhao, Uptake mechanism,
sound frequency and initial concentration on the sonochemical degradation of subcellular distribution, and uptake process of perfluorooctanoic acid and
perfluorooctane sulfonate (PFOS), J. Hazard. Mater. 300 (2015) 662–669, perfluorooctane sulfonic acid by wetland plant Alisma orientale, Sci. Total
https://doi.org/10.1016/j.jhazmat.2015.07.077. Environ. 733 (2020), 139383, https://doi.org/10.1016/j.scitotenv.2020.139383.
[133] J. Cheng, C.D. Vecitis, H. Park, B.T. Mader, M.R. Hoffmann, Sonochemical [157] J.B.N. Mudumbi, S. Ntwampe, F.M. Muganza, O. Okonkwo, Susceptibility of
degradation of perfluorooctane sulfonate (PFOS) and perfluorooctanoate (PFOA) riparian wetland plants to perfluorooctanoic acid (PFOA) accumulation, Int. J.
in landfill groundwater: Environmental matrix effects, Environ. Sci. Technol. 42 Phytoremediat. 16 (2014) 926–936, https://doi.org/10.1080/
(2008) 8057–8063, https://doi.org/10.1021/es8013858. 15226514.2013.810574.
[134] R. James Wood, T. Sidnell, I. Ross, J. McDonough, J. Lee, M.J. Bussemaker, [158] L.M. Colosi, R.A. Pinto, Q. Huang, W.J. Weber, Peroxidase-mediated degradation
Ultrasonic degradation of perfluorooctane sulfonic acid (PFOS) correlated with of perfluorooctanoic acid, Environ. Toxicol. Chem. 28 (2009) 264–271, https://
sonochemical and sonoluminescence characterisation, Ultrason. Sonochem. 68 doi.org/10.1897/08-282.1.
(2020), 105196, https://doi.org/10.1016/j.ultsonch.2020.105196. [159] Q. Luo, J. Lu, H. Zhang, Z. Wang, M. Feng, S.-Y.D.Y.D. Chiang, D. Woodward,
[135] C.D. Vecitis, H. Park, J. Cheng, B.T. Mader, M.R. Hoffmann, Kinetics and Q. Huang, Laccase-catalyzed degradation of perfluorooctanoic acid, Environ. Sci.
mechanism of the sonolytic conversion of the aqueous perfluorinated surfactants, Technol. Lett. 2 (2015) 198–203, https://doi.org/10.1021/acs.estlett.5b00119.
perfluorooctanoate (PFOA), and perfluorooctane sulfonate (PFOS) into inorganic [160] S. Garg, P. Kumar, S. Singh, A. Yadav, L.F. Dumée, R.S. Sharma, V. Mishra,
products, J. Phys. Chem. A 112 (2008) 4261–4270, https://doi.org/10.1021/ Prosopis juliflora peroxidases for phenol remediation from industrial wastewater
jp801081y. – an innovative practice for environmental sustainability, Environ. Technol.
[136] M.Y. Khan, S. So, G. da Silva, Decomposition kinetics of perfluorinated sulfonic Innov. 19 (2020), 100865, https://doi.org/10.1016/j.eti.2020.100865.
acids, Chemosphere 238 (2020), 124615, https://doi.org/10.1016/j. [161] B.D. Turner, S.W. Sloan, G.R. Currell, Novel remediation of per- and
chemosphere.2019.124615. polyfluoroalkyl substances (PFASs) from contaminated groundwater using
[137] P.J. Krusic, A.A. Marchione, D.C. Roe, Gas-phase NMR studies of the thermolysis Cannabis sativa L. (hemp) protein powder, Chemosphere 229 (2019) 22–31,
of perfluorooctanoic acid, J. Fluor. Chem. 126 (2005) 1510–1516, https://doi. https://doi.org/10.1016/j.chemosphere.2019.04.139.
org/10.1016/j.jfluchem.2005.08.016. [162] L.P. Turner, B.H. Kueper, K.M. Jaansalu, D.J. Patch, N. Battye, O. El-Sharnouby,
[138] P.J. Krusic, D.C. Roe, Gas-phase NMR technique for studying the thermolysis of K.G. Mumford, K.P. Weber, Mechanochemical remediation of
materials: Thermal decomposition of ammonium perfluorooctanoate, Anal. perfluorooctanesulfonic acid (PFOS) and perfluorooctanoic acid (PFOA) amended
Chem. 76 (2004) 3800–3803, https://doi.org/10.1021/ac049667k. sand and aqueous film-forming foam (AFFF) impacted soil by planetary ball
[139] A.L. Duchesne, J.K. Brown, D.J. Patch, D. Major, K.P. Weber, J.I. Gerhard, milling, Sci. Total Environ. (2020), 142722, https://doi.org/10.1016/j.
Remediation of PFAS-contaminated soil and granular activated carbon by scitotenv.2020.142722.
smoldering combustion, Environ. Sci. Technol. 54 (2020) 12631–12640, https:// [163] H. Tian, J. Gao, H. Li, S.A. Boyd, C. Gu, Complete defluorination of perfluorinated
doi.org/10.1021/acs.est.0c03058. compounds by hydrated electrons generated from 3-indole-acetic-acid in
[140] E. Crownover, D. Oberle, M. Kluger, G. Heron, Perfluoroalkyl and polyfluoroalkyl organomodified montmorillonite, Sci. Rep. 6 (2016) 32949, https://doi.org/
substances thermal desorption evaluation, Remediation 29 (2019) 77–81, https:// 10.1038/srep32949.
doi.org/10.1002/rem.21623. [164] M. Trojanowicz, I. Bartosiewicz, A. Bojanowska-Czajka, K. Kulisa, T. Szreder,
[141] F. Xiao, P.C. Sasi, B. Yao, A. Kubátová, S.A. Golovko, M.Y. Golovko, D. Soli, K. Bobrowski, H. Nichipor, J.F. Garcia-Reyes, G. Nałęcz-Jawecki, S. Męczyńska-
Thermal stability and decomposition of perfluoroalkyl substances on spent Wielgosz, J. Kisała, Application of ionizing radiation in decomposition of
granular activated carbon, Environ. Sci. Technol. Lett. 7 (2020) 343–350, https:// perfluorooctanoate (PFOA) in waters, Chem. Eng. J. 357 (2019) 698–714,
doi.org/10.1021/acs.estlett.0c00114. https://doi.org/10.1016/j.cej.2018.09.065.
[142] F. Wang, X. Lu, X.-y Li, K. Shih, Effectiveness and mechanisms of defluorination of [165] F. Wang, K. Shih, X. Lu, C. Liu, Mineralization behavior of fluorine in
perfluorinated alkyl substances by calcium compounds during waste thermal Perfluorooctanesulfonate (PFOS) during thermal treatment of lime-conditioned
treatment, Environ. Sci. Technol. 49 (2015) 5672–5680, https://doi.org/ sludge, Environ. Sci. Technol. 47 (2013) 2621–2627, https://doi.org/10.1021/
10.1021/es506234b. es305352p.
[143] D.A. Ellis, S.A. Mabury, J.W. Martin, D.C.G. Muir, Thermolysis of fluoropolymers [166] J.W. Washington, K. Rankin, E.L. Libelo, D.G. Lynch, M. Cyterski, Determining
as a potential source of halogenated organic acids in the environment, Nature 412 global background soil PFAS loads and the fluorotelomer-based polymer
(2001) 321–324, https://doi.org/10.1038/35085548. degradation rates that can account for these loads, Sci. Total Environ. 651 (2019)
[144] T. Yamada, P.H. Taylor, R.C. Buck, M.A. Kaiser, R.J. Giraud, Thermal degradation 2444–2449, https://doi.org/10.1016/j.scitotenv.2018.10.071.
of fluorotelomer treated articles and related materials, Chemosphere 61 (2005) [167] M.H. Russell, W.R. Berti, B. Szostek, N. Wang, R.C. Buck, Evaluation of PFO
974–984, https://doi.org/10.1016/j.chemosphere.2005.03.025. formation from the biodegradation of a fluorotelomer-based urethane polymer
[145] J.S.C.C. Liou, B. Szostek, C.M. DeRito, E.L. Madsen, Investigating the product in aerobic soils, Polym. Degrad. Stab. 95 (2010) 79–85, https://doi.org/
biodegradability of perfluorooctanoic acid, Chemosphere 80 (2010) 176–183, 10.1016/j.polymdegradstab.2009.10.004.
https://doi.org/10.1016/j.chemosphere.2010.03.009.

23
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

[168] M.J.A. Dinglasan, Y. Ye, E.A. Edwards, S.A. Mabury, Fluorotelomer alcohol Purif. Technol. 247 (2020), 116911, https://doi.org/10.1016/j.
biodegradation yields poly- and perfluorinated acids, Environ. Sci. Technol. 38 seppur.2020.116911.
(2004) 2857–2864, https://doi.org/10.1021/es0350177. [193] B. Xu, S. Liu, J.L. Zhou, C. Zheng, J. Weifeng, B. Chen, T. Zhang, W. Qiu, PFAS
[169] J.W. Washington, J.J. Ellington, T.M. Jenkins, J.J. Evans, H. Yoo, S.C. Hafner, and their substitutes in groundwater: occurrence, transformation and
Degradability of an acrylate-Linked, fluorotelomer polymer in soil, Environ. Sci. remediation, J. Hazard. Mater. (2021), https://doi.org/10.1016/j.
Technol. 43 (2009) 6617–6623, https://doi.org/10.1021/es9002668. jhazmat.2021.125159.
[170] M.H. Russell, W.R. Berti, B. Szostek, R.C. Buck, Investigation of the [194] P. McCleaf, S. Englund, A. Ostlund, K. Lindegren, K. Wiberg, L. Ahrens, Removal
biodegradation potential of a fluoroacrylate polymer product in aerobic soils, efficiency of multiple poly- and perfluoroalkyl substances (PFASs) in drinking
Environ. Sci. Technol. 42 (2008) 800–807, https://doi.org/10.1021/es0710499. water using granular activated carbon (GAC) and anion exchange (AE) column
[171] N. Gaur, K. Narasimhulu, P. Y, Recent advances in the bio-remediation of tests, Water Res. 120 (2017) 77–87, https://doi.org/10.1016/j.
persistent organic pollutants and its effect on environment, J. Clean. Prod. 198 watres.2017.04.057.
(2018) 1602–1631, https://doi.org/10.1016/j.jclepro.2018.07.076. [195] Z. Du, S. Deng, Y. Chen, B. Wang, J. Huang, Y. Wang, G. Yu, Removal of
[172] F. Chen, Z. Luo, G. Liu, Y. Yang, S. Zhang, J. Ma, Remediation of electronic waste perfluorinated carboxylates from washing wastewater of perfluorooctanesulfonyl
polluted soil using a combination of persulfate oxidation and chemical washing, fluoride using activated carbons and resins, J. Hazard. Mater. 286 (2015)
J. Environ. Manag. 204 (2017) 170–178, https://doi.org/10.1016/j. 136–143, https://doi.org/10.1016/j.jhazmat.2014.12.037.
jenvman.2017.08.050. [196] F. Dixit, B. Barbeau, S.G. Mostafavi, M. Mohseni, PFOA and PFOS removal by ion
[173] J.-R. Jeon, K. Murugesan, I.-H. Nam, Y.-S. Chang, Coupling microbial catabolic exchange for water reuse and drinking applications: Role of organic matter
actions with abiotic redox processes: a new recipe for persistent organic pollutant characteristics, Environ. Sci.: Water Res. Technol. 5 (2019) 1782–1795, https://
(POP) removal, Biotechnol. Adv. 31 (2013) 246–256, https://doi.org/10.1016/j. doi.org/10.1039/C9EW00409B.
biotechadv.2012.11.002. [197] F. Cao, L. Wang, Y. Tian, F. Wu, C. Deng, Q. Guo, H. Sun, S. Lu, Synthesis and
[174] D. Lu, S. Sha, J. Luo, Z. Huang, X. Zhang Jackie, Treatment train approaches for evaluation of molecularly imprinted polymers with binary functional monomers
the remediation of per- and polyfluoroalkyl substances (PFAS): a critical review, for the selective removal of perfluorooctanesulfonic acid and perfluorooctanoic
J. Hazard. Mater. 386 (2020), 121963, https://doi.org/10.1016/j. acid, J. Chromatogr. A 1516 (2017) 42–53, https://doi.org/10.1016/j.
jhazmat.2019.121963. chroma.2017.08.023.
[175] N. Watanabe, M. Takata, S. Takemine, K. Yamamoto, Thermal mineralization [198] L. Zhao, J. Bian, Y. Zhang, L. Zhu, Z. Liu, Comparison of the sorption behaviors
behavior of PFOA, PFHxA, and PFOS during reactivation of granular activated and mechanisms of perfluorosulfonates and perfluorocarboxylic acids on three
carbon (GAC) in nitrogen atmosphere, Environ. Sci. Pollut. Res. 25 (2018) kinds of clay minerals, Chemosphere 114 (2014) 51–58, https://doi.org/
7200–7205, https://doi.org/10.1007/s11356-015-5353-2. 10.1016/j.chemosphere.2014.03.098.
[176] L.-h Zhang, J.-h Cheng, X. You, X.-y Liang, Y.-y Hu, Photochemical defluorination [199] L. Qiang, M. Chen, L. Zhu, W. Wu, Q. Wang, Facilitated bioaccumulation of
of aqueous perfluorooctanoic acid (PFOA) by Fe0/GAC micro-electrolysis and perfluorooctanesulfonate in common carp (Cyprinus carpio) by graphene oxide
VUV-Fenton photolysis, Environ. Sci. Pollut. Res. 23 (2016) 13531–13542, and remission mechanism of fulvic acid, Environ. Sci. Technol. 50 (2016)
https://doi.org/10.1007/s11356-016-6539-y. 11627–11636, https://doi.org/10.1021/acs.est.6b02100.
[177] A. Ahmed, J. Wang, W. Wang, C.J. Okonkwo, N. Liu, A practical method to [200] H. Lin, Y. Wang, J. Niu, Z. Yue, Q. Huang, Efficient sorption and removal of
remove perfluorooctanoic acid from aqueous media using layer double hydride Perfluoroalkyl Acids (PFAAs) from aqueous solution by metal hydroxides
system: a prospect for environmental remediation, Environ. Technol. (2020) generated in-situ by electrocoagulation, Environ. Sci. Technol. 49 (2015)
1–12, https://doi.org/10.1080/09593330.2020.1812733. 10562–10569, https://doi.org/10.1021/acs.est.5b02092.
[178] W. Zhang, Y. Liang, Removal of eight perfluoroalkyl acids from aqueous solutions [201] M.J. Bentel, Y. Yu, L. Xu, Z. Li, B.M. Wong, Y. Men, J. Liu, Defluorination of per-
by aeration and duckweed, Sci. Total Environ. 724 (2020), 138357, https://doi. and polyfluoroalkyl substances (PFASs) with hydrated electrons: Structural
org/10.1016/j.scitotenv.2020.138357. dependence and implications to PFAS remediation and management, Environ.
[179] V.A. Arias Espana, M. Mallavarapu, R. Naidu, Treatment technologies for aqueous Sci. Technol. 53 (2019) 3718–3728, https://doi.org/10.1021/acs.est.8b06648.
perfluorooctanesulfonate (PFOS) and perfluorooctanoate (PFOA): a critical [202] M.-J. Chen, S.-L. Lo, Y.-C. Lee, C.-C. Huang, Photocatalytic decomposition of
review with an emphasis on field testing, Elsevier, 2015, pp. 168–181. htt perfluorooctanoic acid by transition-metal modified titanium dioxide, J. Hazard.
ps://doi.org/10.1016/j.eti.2015.06.001. Mater. 288 (2015) 168–175, https://doi.org/10.1016/j.jhazmat.2015.02.004.
[180] G. Van Gils, Oil/water emulsion and aqueous film forming foam (AFFF) treatment [203] M. Li, Z. Yu, Q. Liu, L. Sun, W. Huang, Photocatalytic decomposition of
using air-sparged hydrocyclone technology. https://apps.dtic.mil/dtic/tr/fulltext perfluorooctanoic acid by noble metallic nanoparticles modified TiO2, Chem. Eng.
/u2/a607005.pdf (Accessed 30 December 2020). J. 286 (2016) 232–238, https://doi.org/10.1016/j.cej.2015.10.037.
[181] M. Inyang, E. Dickenson, The potential role of biochar in the removal of organic [204] C.B. Ong, A.W. Mohammad, L.Y. Ng, E. Mahmoudi, S. Azizkhani, N.H.H. Hairom,
and microbial contaminants from potable and reuse water: a review, Solar photocatalytic and surface enhancement of ZnO/rGO nanocomposite:
Chemosphere 134 (2015) 232–240, https://doi.org/10.1016/j. Degradation of perfluorooctanoic acid and dye, Process Saf. Environ. Prot. 112
chemosphere.2015.03.072. (2017) 298–307, https://doi.org/10.1016/j.psep.2017.04.031.
[182] Y. Huang, Y. Yu, Y. Yu, B. Zhang, Oxygen vacancy engineering in photocatalysis, [205] E. Shang, Y. Li, J. Niu, S. Li, G. Zhang, X. Wang, Photocatalytic degradation of
Sol. RRL 4 (2020), 2000037, https://doi.org/10.1002/solr.202000037. perfluorooctanoic acid over Pb-BiFeO3/rGO catalyst: kinetics and mechanism,
[183] H. Tan, Z. Zhao, W.-b Zhu, E.N. Coker, B. Li, M. Zheng, W. Yu, H. Fan, Z. Sun, Chemosphere 211 (2018) 34–43, https://doi.org/10.1016/j.
Oxygen vacancy enhanced photocatalytic activity of pervoskite SrTiO3, ACS Appl. chemosphere.2018.07.130.
Mater. Interfaces 6 (2014) 19184–19190, https://doi.org/10.1021/am5051907. [206] C. Song, P. Chen, C. Wang, L. Zhu, Photodegradation of perfluorooctanoic acid by
[184] Y. Wu, Y. Li, C. Fang, C. Li, Highly efficient degradation of perfluorooctanoic acid synthesized TiO2-MWCNT composites under 365 nm UV irradiation,
over a MnOx-modified oxygen-vacancy-rich In2O3 photocatalyst, ChemCatChem Chemosphere 86 (2012) 853–859, https://doi.org/10.1016/j.
11 (2019) 2297–2303, https://doi.org/10.1002/cctc.201900273. chemosphere.2011.11.034.
[185] Y. Sun, G. Li, W. Wang, W. Gu, P.K. Wong, T. An, Photocatalytic defluorination of [207] Y. Wang, P. Zhang, Photocatalytic decomposition of perfluorooctanoic acid
perfluorooctanoic acid by surface defective BiOCl: fast microwave solvothermal (PFOA) by TiO2 in the presence of oxalic acid, J. Hazard. Mater. 192 (2011)
synthesis and photocatalytic mechanisms, J. Environ. Sci. 84 (2019) 69–79, 1869–1875, https://doi.org/10.1016/j.jhazmat.2011.07.026.
https://doi.org/10.1016/j.jes.2019.04.012. [208] X. You, L.-l Yu, F.-f Xiao, S.-c Wu, C. Yang, J.-h Cheng, Synthesis of
[186] Z. Song, X. Dong, J. Fang, C. Xiong, N. Wang, X. Tang, Improved photocatalytic phosphotungstic acid-supported bimodal mesoporous silica-based catalyst for
degradation of perfluorooctanoic acid on oxygen vacancies-tunable bismuth defluorination of aqueous perfluorooctanoic acid under vacuum UV irradiation,
oxychloride nanosheets prepared by a facile hydrolysis, J. Hazard. Mater. 377 Chem. Eng. J. 335 (2018) 812–821, https://doi.org/10.1016/j.cej.2017.10.123.
(2019) 371–380, https://doi.org/10.1016/j.jhazmat.2019.05.084. [209] H. Lin, J. Niu, S. Ding, L. Zhang, Electrochemical degradation of
[187] M. Wang, M. Shen, X. Jin, J. Tian, M. Li, Y. Zhou, L. Zhang, Y. Li, J. Shi, Oxygen perfluorooctanoic acid (PFOA) by Ti/SnO2-Sb, Ti/SnO2-Sb/PbO2 and Ti/SnO2-
vacancy generation and stabilization in CeO2–x by Cu introduction with improved Sb/MnO2 anodes, Water Res. 46 (2012) 2281–2289, https://doi.org/10.1016/j.
CO2 photocatalytic reduction activity, ACS Catal. 9 (2019) 4573–4581, https:// watres.2012.01.053.
doi.org/10.1021/acscatal.8b03975. [210] Y. Wang, R.D. Pierce, H. Shi, C. Li, Q. Huang, Electrochemical degradation of
[188] V. Kumaravel, S. Rhatigan, S. Mathew, M.C. Michel, J. Bartlett, M. Nolan, S. perfluoroalkyl acids by titanium suboxide anodes, Environ. Sci.: Water Res.
J. Hinder, A. Gascó, C. Ruiz-Palomar, D. Hermosilla, S.C. Pillai, Mo doped TiO2: Technol. 6 (2020) 144–152, https://doi.org/10.1039/C9EW00759H.
impact on oxygen vacancies, anatase phase stability and photocatalytic activity, [211] Z.S. Xu, Y.X. Yu, H. Liu, J.F. Niu, Highly efficient and stable Zr-doped
J. Phys.: Mater. 3 (2020), 025008, https://doi.org/10.1088/2515-7639/ab749c. nanocrystalline PbO2 electrode for mineralization of perfluorooctanoic acid in a
[189] D. Huang, L. Yin, J. Niu, Photoinduced hydrodefluorination mechanisms of sequential treatment system, Sci. Total Environ. 579 (2017) 1600–1607, https://
perfluorooctanoic acid by the SiC/graphene catalyst, Environ. Sci. Technol. 50 doi.org/10.1016/j.scitotenv.2016.11.180.
(2016) 5857–5863, https://doi.org/10.1021/acs.est.6b00652. [212] B. Yang, J. Wang, C. Jiang, J. Li, G. Yu, S. Deng, S. Lu, P. Zhang, C. Zhu, Q. Zhuo,
[190] S.C. Panchangam, C.S. Yellatur, J.-S. Yang, S.S. Loka, A.Y.C. Lin, V. Vemula, Electrochemical mineralization of perfluorooctane sulfonate by novel F and Sb co-
Facile fabrication of TiO2-graphene nanocomposites (TGNCs) for the efficient doped Ti/SnO2 electrode containing Sn-Sb interlayer, Chem. Eng. J. 316 (2017)
photocatalytic oxidation of perfluorooctanoic acid (PFOA), J. Environ. Chem. 296–304, https://doi.org/10.1016/j.cej.2017.01.105.
Eng. 6 (2018) 6359–6369, https://doi.org/10.1016/j.jece.2018.10.003. [213] Y. Liu, C.J. Ptacek, R.J. Baldwin, J.M. Cooper, D.W. Blowes, Application of zero-
[191] K. Park, I. Ali, J.-O. Kim, Photodegradation of perfluorooctanoic acid by graphene valent iron coupled with biochar for removal of perfluoroalkyl carboxylic and
oxide-deposited TiO2 nanotube arrays in aqueous phase, J. Environ. Manag. 218 sulfonic acids from water under ambient environmental conditions, Sci. Total
(2018) 333–339, https://doi.org/10.1016/j.jenvman.2018.04.016. Environ. 719 (2020), 137372, https://doi.org/10.1016/j.scitotenv.2020.137372.
[192] M.-K. Kim, T. Kim, T.-K. Kim, S.-W. Joo, K.-D. Zoh, Degradation mechanism of
perfluorooctanoic acid (PFOA) during electrocoagulation using Fe electrode, Sep.

24
S. Garg et al. Journal of Environmental Chemical Engineering 9 (2021) 105784

[214] Z. Zhang, J.-J. Chen, X.-J. Lyu, H. Yin, G.-P. Sheng, Complete mineralization of process, Water Res. 127 (2017) 50–58, https://doi.org/10.1016/j.
perfluorooctanoic acid (PFOA) by γ-irradiation in aqueous solution, Sci. Rep. 4 watres.2017.10.010.
(2014) 7418, https://doi.org/10.1038/srep07418. [221] V. Franke, P. McCleaf, K. Lindegren, L. Ahrens, Efficient removal of per- and
[215] R.K. Singh, S. Fernando, S.F. Baygi, N. Multari, S.M. Thagard, T.M. Holsen, polyfluoroalkyl substances (PFASs) in drinking water treatment: nanofiltration
Breakdown products from perfluorinated alkyl substances (PFAS) degradation in combined with active carbon or anion exchange, Environ. Sci.: Water Res.
a plasma-based water treatment process, Environ. Sci. Technol. 53 (2019) Technol. 5 (2019) 1836–1843, https://doi.org/10.1039/C9EW00286C.
2731–2738, https://doi.org/10.1021/acs.est.8b07031. [222] Y. Sun, B. Angelotti, M. Brooks, B. Dowbiggin, P.J. Evans, B. Devins, Z.-W. Wang,
[216] M. Sörengård, A.S. Lindh, L. Ahrens, Thermal desorption as a high removal A pilot-scale investigation of disinfection by-product precursors and trace organic
remediation technique for soils contaminated with per- and polyfluoroalkyl removal mechanisms in ozone-biologically activated carbon treatment for potable
substances (PFASs), PLOS ONE 15 (2020), 0234476, https://doi.org/10.1371/ reuse, Chemosphere 210 (2018) 539–549, https://doi.org/10.1016/j.
journal.pone.0234476. chemosphere.2018.06.162.
[217] F. Wang, X. Lu, K. Shih, C. Liu, Influence of calcium hydroxide on the fate of [223] C.C. Murray, H. Vatankhah, C.A. McDonough, A. Nickerson, T.T. Hedtke, T.
perfluorooctanesulfonate under thermal conditions, J. Hazard. Mater. 192 (2011) Y. Cath, C.P. Higgins, C.L. Bellona, Removal of per- and polyfluoroalkyl
1067–1071, https://doi.org/10.1016/j.jhazmat.2011.06.009. substances using super-fine powder activated carbon and ceramic membrane
[218] N. Tseng, Feasibility of biodegradation of polyfluoroalkyl and perfluoroalkyl filtration, J. Hazard. Mater. 366 (2019) 160–168, https://doi.org/10.1016/j.
substances, Department of Civil Engineering, University of California, 2012. htt jhazmat.2018.11.050.
ps://escholarship.org/uc/item/2x47296b (29 December 2020). [224] T. Xu, H. Ji, Y. Gu, T. Tong, Y. Xia, L. Zhang, D. Zhao, Enhanced adsorption and
[219] A. Presentato, S. Lampis, A. Vantini, F. Manea, F. Daprà, S. Zuccoli, G.J.M. Vallini, photocatalytic degradation of perfluorooctanoic acid in water using iron (hydr)
On the ability of perfluorohexane sulfonate (PFHxS) bioaccumulation by two oxides/carbon sphere composite, Chem. Eng. J. 388 (2020), 124230, https://doi.
Pseudomonas sp. strains isolated from PFAS-contaminated environmental org/10.1016/j.cej.2020.124230.
matrices, Microorganisms 8 (2020) 92, https://doi.org/10.3390/ [225] L. Wang, M. Nickelsen, S.-Y. Chiang, S. Woodard, Y. Wang, S. Liang, R. Mora,
microorganisms8010092. R. Fontanez, H. Anderson, Q. Huang, Repeated subarachnoid administrations of
[220] Z. Sun, C. Zhang, P. Chen, Q. Zhou, M.R. Hoffmann, Impact of humic acid on the allogeneic human umbilical cord mesenchymal stem cells for spinal cord injury: a
photoreductive degradation of perfluorooctane sulfonate (PFOS) by UV/Iodide phase 1/2 pilot study, Cytotherapy 23 (2021) 57–64, https://doi.org/10.1016/j.
ceja.2020.100078.

25

You might also like