Next Article in Journal
Exploring Key Factors and Driving Mechanisms of Construction Waste Recycling Development in China: Combination of PEST Model and Fuzzy-Set Qualitative Comparative Analysis
Next Article in Special Issue
A Numerical Investigation Concerning the Effect of Step-Feeding on Performance of Constructed Wetlands Operating under Mediterranean Conditions
Previous Article in Journal
Digital Twin Technology in Data Center Simulations: Evaluating the Feasibility of a Former Mine Site
Previous Article in Special Issue
Usage and Microbial Safety of Shared and Unshared Excreta Disposal Facilities in Developing Countries: The Case of a Ghanaian Rural District
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Critical Review on PFAS Removal from Water: Removal Mechanism and Future Challenges

1
Department of Sustainable Bioproducts, Mississippi State University, Mississippi State, MS 39762, USA
2
Department of Forestry, College of Forest Resources, Mississippi State University, Mississippi State, MS 39762, USA
*
Author to whom correspondence should be addressed.
Sustainability 2023, 15(23), 16173; https://doi.org/10.3390/su152316173
Submission received: 26 September 2023 / Revised: 15 November 2023 / Accepted: 18 November 2023 / Published: 21 November 2023
(This article belongs to the Special Issue Sustainable Wastewater: Solutions, Treatment & Opportunities)

Abstract

:
Per and polyfluoroalkyl substances (PFAS) have been extensively employed in a broad range of manufacturing and consumer goods due to their highly persistent nature. PFAS exposure is recognized to pose serious health hazards; therefore, addressing PFAS pollution in water has become a top priority for public health and environmental protection organizations. This review article focuses on the efficiency of different removal techniques (activated carbon, biochar, ion exchange resin, membrane filtration, reverse osmosis, metal-organic frameworks, foam fractionation, ozone fractionation, and destruction techniques) for eliminating different types of short- and long-chain PFAS from water. Hydrophobicity and electrostatic interactions are revealed to be the primary mechanisms for the elimination of PFAS. The efficiency of all techniques to eradicate short-chain PFAS is comparatively lower compared to long-chain PFAS. The destruction techniques are the most efficient but have some drawbacks, including the formation of PFAS precursors and high operational costs. According to the findings from the study, it is anticipated that combined methods will be required to effectively remediate PFAS-contaminated water.

1. Introduction

Per- and polyfluoroalkyl substances (PFAS) are man-made fluorine-containing organic compounds exhibiting distinctive physiochemical characteristics due to the strong carbon and fluorine bonding (C-F) [1,2]. PFAS are comprised of hydrophobic and oleophobic carbon chains in which the hydrogen can be completely or partly substituted with fluorine atoms [3,4]. Perfluoro carboxylic acids (PFCAs) and perfluoroalkane sulfonic acids (PFSAs) are the two well-studied major subclasses of PFAS [5]. They are also recognized as “forever chemicals” because of their persistent nature. Over 5000 different forms of PFAS exist in the world [6]. Due to the exceptional resistance properties of PFAS to heat, oil, and water, they are extensively used in different areas, including domestic (nonstick cookware), firefighting foams, and industrial food packaging applications [7]. It has been reported that PFAS concentrations in aqueous medium vary from pg/L to mg/L [8]. According to Hopkins et al. [9], the Cape Fear River, a water supply utilized by utilities in North Carolina, possesses high quantities of different PFAS [9]. In September 2015, PFAS contamination was identified by the New South Wales Government of Australia in Williamtown. Subsequently, in August 2016, Australia’s Department of Defense identified the key exposure route of PFAS contamination, causing health hazards by contaminating drinking groundwater [10].
Six illnesses have been linked to PFOA exposure in a study conducted involving more than 68,000 people living near a manufacturing factory in the United States, including testicular cancer, high cholesterol, pregnancy-hypertension, kidney cancer, thyroid disease, and ulcerative colitis [11]. PFAS have also been found in umbilical cord blood, breast milk, and human blood [12,13]. COVID-19 is known to be worsened by pulmonary buildup of PFAS, and the latest evidence found that increased plasma-PFAS concentrations in human blood are linked to this severity [14]. Over 95% of Americans have PFAS in their blood [5,15,16]. Human exposure to PFAS could happen through PFAS-contaminated food consumption (grown on PFAS-contaminated soil) [17]. PFAS may infiltrate soil through a variety of routes. PFAS concentration in soil ranges from nanograms to micrograms [18]. From the soil, the PFAS could leach down to contaminate the groundwater, which is another route of human exposure to PFAS [19]. The Stockholm Convention on Persistent Organic Pollutants (POPs) placed PFOA on its Annex A (elimination) list and PFOS on its Annex B (restricted) list due to the prolonged toxicity and biomagnification of both substances for a worldwide ban [20]. The Environmental Protection Agency (EPA) forbids corporations in the United States from producing, exporting, or importing items that include specific long-chain PFAS without first undergoing EPA inspection and clearance [21].
Short-chain PFAS (C2–C7) have been used as an alternative to PFOS and PFOA (C8) due to their decreased production [22]. In 2009, Dupont developed the PFOA alternative GenX (C6), also known as hexafluoropropylene oxide dimer acid (HFPO-DA), which was later commercialized by Chemours, a DuPont subsidiary. Due to its widespread use in industry as a less harmful replacement, GenX has gained greater prominence compared to other short-chain PFAS [22]. It has been detected in river water samples collected from different countries, including the United States, China, Germany, and the Netherlands. The GenX concentration in those samples was higher than the permissible limits of PFAS and PFOS set by the US EPA [23]. In recent investigations, GenX has been shown to be more hazardous and bioaccumulative compared to PFOA, and the efficiency of the drinking water plant to remove GenX was quite poor [24,25,26]. PFAS are very difficult to remove from the environment due to their persistent physiochemical and thermostable nature; however, owing to their increased solubility and mobility, the shorter ones are more challenging to eliminate [2,8]. Numerous techniques have been used to remove the PFAS and GenX from the water, including electrooxidation [24], activated carbon (AC) [27], membrane technologies [11,28], electrochemical degradation, clay minerals [8], ion-exchange technologies [7], nanoparticles [29,30], reverse osmosis [31], biochar [32,33], etc. However, research studies on GenX removal are still scarce. The efficacy of various methods for eradicating GenX and PFAS must be compared. Long- and short-chain PFAS, such as GenX, are the main focus of this assessment, along with their background, removal methods, removal processes, cost analysis, and comparison of efficiency across various technologies.

History and Sources of PFAS

Roy J. Plunkett accidentally discovered a fluoropolymer known as polytetrafluoroethylene (PTFE) on 6 April 1938, while doing an experimental study on DuPont Freon refrigerant gases. Ten years of study culminated in 1941 with the patenting and commercialization of PTFE under the brand name Teflon [34]. In 1947, another company named 3M invented its trademark PFOA with the invention of numerous chemical compounds having eight carbons tied to fluorine [35]. In 1963, scientists in the US Navy worked with 3M to manufacture aqueous film-forming foams and patented them in 1967 [36]. After that, different forms of PFAS were developed by numerous industries, e.g., packaging, paint, cleaning, and fabric industries, etc. A West Virginia farmer named Wilbur Tennant filed the first complaint against DuPont in 1999, accusing the company of releasing PFAS into local water supplies. Tennant saw several health problems, including black teeth, birth deformities, and cancers in his cattle. More than a hundred of his animals died after consuming water from a nearby creek that had been poisoned by PFAS sludge discharged by DuPont in the 1980s [35]. A summary of PFAS events history is displayed in Figure 1.
After that, PFAS were found to cause cancer, immune system suppression, disturbance of thyroid hormones, and low birth weight in infants [37,38]. As a result, efforts have been executed to mitigate PFAS environmental effects. The two most produced and utilized PFAS are PFOA and PFOS. Both are highly persistent and bioaccumulative [39,40]. Thus, the Stockholm Convention’s POPs list restricted their use. The PFOA Stewardship Program, launched by the US Environmental Protection Agency in January 2006, has accelerated the process of phasing out this chemical [41]. Despite having similar harmfulness and persistence, the short-chain substituents are thought to be less bioaccumulative and so represent less risk to human health. The chemical stability of the PFOA alternative GenX is low because of the presence of etheric bonds between the perfluorinated moieties [42]. According to the studies, GenX levels in the water systems were greater than the combined EPA recommendations for PFOS and PFOA. GenX, as shown in recent research, poses a significant challenge to water treatment facilities because of its high toxicity and bioaccumulation, which are similar to or superior to PFOA [24,25]. Table 1 demonstrates some physiochemical characteristics of the different types of PFAS reported in this study with their abbreviation given below:
There are two types of PFAS emission sources: direct emission sources and indirect emission sources. The direct emission sources include emissions at any time in the lifecycle when products contain the substance as a primary or residual impurity. The indirect emission sources refer to the formation of the product from the transformation of the product’s precursors in the environment. Primary PFAS producers, AFFF users, and downstream sectors (such as leather, metal plating, semiconductors, and textiles) are all direct contributors to PFAS contamination [43]. The two main indirect sources of PFAS contamination are water treatment plants and municipal landfills [43]. Some common sources of PFAS are exhibited in Figure 2.
AFFF (Airforce aqueous film-forming foam) formulations often include PFAS due to their performance as surfactants for Class B hazardous fuel fires [44]. The AFFF produced by 3M using electrochemical fluorination (ECF) technology is known as legacy PFOS AFFF [45]. The military, aviation industry, oil and gas industries, and the firefighting sector were among the first to use AFFF due to their superior efficacy and efficiency in combating fires [45,46].
Flame retardants in electronics may be either brominated, like polybrominated diphenyl ethers (PBDEs), or fluorinated, like PFAS [47]. As a flame retardant, polycarbonate resins often include the potassium salt of perfluorobutane sulfonate (PFBS). Lithium salts of PFAS are used as electrolytes in fuel cells and batteries [48]. Each year, we produce more than 50 million metric tons of e-waste, which includes all forms of electronic and electrical equipment as well as their components that have been discarded [49]. Electronic waste processing plants and landfills that accept abandoned electrical and electronic equipment discharge toxic substances, such as PFAS, into the environment. The leachate that is produced during landfill treatment and then treated locally or transported to wastewater treatment facilities contains PFAS compounds, and there is currently no efficient way to remediate or remove them [48,50].
New PFAS compounds, including fluorotelomers, require further study in wastewater because of their widespread use in industries like metal plating, oil production, fluorinated impregnations, and food packaging [51]. There is evidence from studies indicating shorter-chain PFAS compounds have increased in prevalence after the prohibition of longer-chain PFAS compounds [52]. Wastewater treatment plants (WWTPs) are used to treat wastewater from a variety of sources, including but not limited to landfill leachates, agricultural wastewater, and effluents produced from home and industrial sources. Even though many pollutants are degraded throughout the WWTP process, PFAS are resistant to the typical treatment procedures used in WWTPs [53]. Due to the transformation of large-chain precursor PFAS compounds into smaller-chain PFOS during treatment, PFAS concentrations in the effluent are frequently higher than those in the influent [51,54]. PFAS are less known for its use in cosmetics and other personal care items. The perfluoroalkyl moiety gives compounds interesting characteristics such as strong surface activity, resistance to water and oil, and chemical and thermal stability [43]. U.S. Environmental Protection Agency researchers have shown that PFAS may be found in a variety of common home items, including food packaging materials [55]. To improve their water- and oil-resistance, several types of packaging include a PFAS coating. In the end, these items are often thrown in landfills where they seep into different bodies of water.

2. Treatment Techniques for the Removal of PFAS

At present, water utilities have few viable methods for PFAS removal. Traditional filtration methods like GAC filters, reverse osmosis, and ion-exchange resins have proven to be effective in removing PFAS but they were not developed to precisely bind these emerging pollutants. The PFAS molecules might pass through these technologies without destruction or sorption. Researchers are currently building PFAS-specific adsorbents and treatment procedures for the complete destruction of PFAS molecules rather than just capturing and storing them [56]. The following are the recent technologies used by researchers for the treatment of PFAS-contaminated wastewater.

2.1. Sorption

2.1.1. Biochar

Biochar is a carbonaceous, highly stable, porous, and environment-friendly material. Biochar is primarily made from the abundant, low-cost, and widely available biowastes of the agricultural and food industries [57]. The surface of biochar is comprised of different types of functional groups, mainly heterocyclic and aromatic, including hydroxyl, carboxyl, carbonyl, etc. which assist in PFAS adsorption [58]. Hydrophobicity and electrostatic interactions are the main mechanisms of PFAS removal using biochar [59]. Biochar has been shown to be an effective method for detoxifying water of pharmaceuticals and heavy metals [60,61,62]. In addition, it is possible to tailor its exceptional physical features (like pore size distribution) to the specific needs of a variety of applications (like the removal of short-chain PFAS).
In order to remove PFAS from wastewater, Patel et al. produced a unique economical biochar for PFAS removal from the byproduct of anaerobic digestion of organic contaminants in the wastewater treatment process [59]. The anaerobic digestion of organic pollutants releases biogas, which can be converted into hydrogen and carbon nanomaterials (CNMs) by sequestering the carbon during catalytic decomposition. The carbon nanomaterials were loaded on the biochar (CNM-s loaded biochar). The CNMs-loaded biochar exhibited a higher removal efficiency (79%) for PFAS compared to ilmenite-modified biochar (54%). Softwood and hardwood shavings were used to prepare biochar by pyrolyzing at varied temperatures (300, 400, and 600 °C) for PFAS removal. These biochars were post-modified by thermal oxidation at 400 °C for half an hour which results in a high surface area and increased pore size. Once PFAS has been adsorbed on biochar, they are destroyed by heat activation at 500 °C in the presence of air or nitrogen. The biochar shows higher performance for PFAS removal in the next cycle after the thermal activation [63]. In another study conducted by Wang et al., softwood and hardwood shavings were used to fabricate the biochar at different pyrolysis temperatures of 300–600 °C [64]. After pyrolysis, the material is subjected to hot air oxidation, which boosts the material’s ability to absorb by increasing the surface area and pore size. This process improves the reactivity of biochar made from softwood for some PFAS by threefolds compared to untreated biochar. After being exposed to a solution containing PFCAs, PFSAs, and GenX, the materials were set aside to dry, followed by heating the dried materials at 500 °C in the air for 30 min to eliminate the PFAS. After being treated, the biochars and granular activated carbon (GAC) were put through further sorption tests, this time with a combination of PFBA, PFPeA, GenX, PFHpA, and PFOA. Reactivated biochar virtually always had a higher adsorption-desorption distribution coefficient (Kd) for all PFAS than raw charcoal.
Krahn et al. used an effective waste management practice for the fabrication of biochar by pyrolyzing it at 700 °C using sewage sludge (SSBCs) and wood chips (WCBC) waste for the elimination of short-chain per fluorinated carboxylic acids (PFPeA, PFHxA, and PFHpA) and long-chains (PFOA, PFNA, and PFDA) and the adsorption of all except PFPeA and PFHxA onto two biochar’s (SSBC and WCBC) followed a strong log-linear relationship (r2 > 0.9) [65]. These poor fits may be explained by the fact that ionic interactions are more crucial for the sorption of these short-chain PFCAs than hydrophobic ones are for the sorption of the longer-chain PFCAs.
Two-step pyrolysis or post-modification was utilized for producing the biochar in other research that used discarded coffee grounds. Two different approaches were investigated for functionalizing the biochar surface with nitrogen-containing groups on the surface, which would then allow for radical-initiated polymerization. The first approach was electrophilic aromatic substitution, carried out by mixing exact amounts of pristine biochar with nitric and sulfuric acid, following the reduction of the nitro group to -NHx using acetic acid, ammonium hydroxide, and sodium dithionite. In the second approach, they added melamine (BC-M) or ammonium chloride (BC-N) to the biochar water mixture and then heated it to initiate the nitrogen modification. To improve the adsorbent’s selectivity for PFAS, a layer of molecularly imprinted polymers (MIP) was added onto the surface of the nitrogen-modified biochar using thermally stimulated radical-initiated polymerization. The BC-M-MIP exhibited higher Kselectivity for PFBS (4.52) and PFOA (3.76) compared to unmodified biochar at 0.043 and 0.039 mg PFAA/g*g/m2, respectively [66]. In another experiment, tree bark and eucalyptus biochar were used for the removal of different chain lengths of PFAS (4 to 11 fluorinated carbons). The sorption increased with the chain length, which points out hydrophobic interactions as an important mechanism in the adsorption process [67]. Wu et al. used different feedstocks (switchgrass (SG), water oak leaves (WO), and biosolids (BS)) for the fabrication of biochar and modified them using FeCl3 and carbon nanotubes (CNTs), which increased their surface area and porosity [68]. The fastest adsorption equilibrium was achieved by the oak leaf-biochar composite. The biosolid-biochars in either pristine or modified form exhibited higher adsorption for PFOA compared to the other two biochars. The maximum adsorption capacity for PFOA was exhibited by BS-Fe (469.65 μmol/g), and the minimum adsorption potential of 39.54 μmol/g was exhibited by WO-CNT (Figure 3).
Gasification at 900–1000 °C was used to create biochar from Douglas fir in another investigation, using syngas as a byproduct. To modify the biochar with Fe2O3, iron (III) chloride (18 g), and iron (II) sulfate (36.6 g) were added to a solution (2 L) comprising BC (50 g). Next, after keeping the pH constant at 10 for 24 h while gently introducing 10 M NaOH, Fe3O4 particles were produced and deposited onto BC. The Fe2O3 modified biochar was ready after vacuum filtration, washing (3x ethanol), and drying (50 °C) step. The Fe2O3-modified biochar removed 75–85% of PFBS, GenX, and perfluorooctane sulfonamide (PFOSA) compared to PFOS (82%). The primary mechanisms for their adsorption onto Douglas fir Fe2O3-modified biochar were determined to be electrostatic and hydrophobicity interactions [69]. Modified or activated biochar exhibited higher adsorption potentials for both short- and long-chain PFAS compared to pristine biochar. The primary mechanism of PFAS removal by biochar hydrophobic and electrostatic attraction. The removal of short-chain PFAS is comparatively low; therefore, further investigation on sustainable removal methods specifically targeting short-chain PFAS during wastewater treatment is needed.

2.1.2. Activated Carbon

Granular activated carbon (GAC) adsorption has been shown to be an effective strategy for detoxifying water contaminated with long-chain PFAS. According to the Web of Science database, activated carbon (AC) is at the top of the list of adsorbents (40% of PFAS studies) for PFAS removal. The non-polar and hydrophobic surface of AC significantly interacts with hydrophobic PFAS, leading to its extensive usage in adsorption processes [70,71]. In recent years, GAC has been utilized in several trial experiments to investigate the impact of PFAS features, such as isomer structure (branched or linear), perfluorocarbon chain length, and functional groups (carboxylates or sulfonates), on the elimination of different PFAS. On aggregate, GAC was shown to be around 80% and 90% efficient in removing PFAS [72,73,74]. Activated carbon has become the most widely utilized sorbent for PFAS removal because it may function both as a single intervention and as part of a pre-treatment step.
Mohamed et al. used a cost-effective method to fabricate AC using sludge-based activated carbon (SBAC). Domestic wastewater treatment facility sludge was used in this study [75]. The manufactured sewage sludge (SS) was dried and pyrolyzed (at 500 °C for 2 h), then treated with 2.5 M ZnCl2 with agitation. The optimal SBAC was developed at activation time (0.5, 0.75, 1, 1.5, and 2 h), ZnCl2 concentration (1.5, 2, and 2.5 M), and pyrolysis temperature (400, 450, and 500 °C). To evaluate the viability of SBAC manufacturing, a life cycle cost analysis (LCCA) was also carried out. Most altered SBACs with a ZnCl2 impregnation ratio of 2.5 M removed PFAS and MB effectively (95.1–100%), except for the sample generated at a low temperature (400 °C). For short-chain PFAS chemicals, this sample had the least adsorption rate (35.3–57.8%). The high solubility and poor hydrophobicity of the SBAC made at 400 °C may explain why it has a low short-chain PFAS sorption potential. The optimal SBACs generated at 1.5 and 2 M ZnCl2 under varied pyrolysis durations were examined. Consequently, PFAS and MB were removed at 90.6–100%. The reduction of the ZnCl2 ratio from 2.5 to 1.5 reduced the possibility of mineral depletion and ecotoxicity of terrestrial, marine, and freshwater by 38%, 24%, 22%, and 49%, which are the most impacted factors. By comparison to natural adsorbents like kaolinite, whose surface area is measured in the tens of m2/g, AC’s is 100 m2/g. Furthermore, commercial AC is notably more costly than natural adsorbents (>4 USD per kilogram) [75,76]. Research by Mohammad et al. found that the price of producing SBAC may be brought down to 1.2 USD/kg [75]. After adsorbing short-chain PFAS, AC must be regenerated, which substantially increases the PFAS treatment price [77]. Meanwhile, Zhong et al. found that the biologically activated carbon (BAC) with different age periods of 1, 4, 7, and 13 years were analyzed for the sorption of 14 PFAS types from drinking water treatment plants (DWTPs) [78]. As a result of their surface and microporous nature, bacteria may cling to GAC and transform it into BAC via the formation of biofilms. Neither the carbon maturity nor the depth had any effect on the microbial species present in the BAC biofilm; rather, the major difference was population. Bacteria of the proteobacterial family have a significant impact on BAC’s ability to remove PFAS. Raw water had a total PFAS content of 127.4 ng/L. The total PFAS concentration decreased to 101.9 ng/L following drinking water treatment plant remediation. In terms of the relative effectiveness of the various methods of removal, pre-ozone treatment led to a rise in PFAS content, from 127.4 ng/L to 142.7 ng/L. The occurrence of its intermediates changed to PFHxA by ozonation likely accounts for most of the rise in PFAS content [79]. Treatment with ozone biological activated carbon (O3-BAC) was found to be promising in removing PFAS from water (20.74%), indicating that O3-BAC plays a pivotal role in the treatment of water to eliminate PFAS [78]. Riegel et al. conducted lab and field experiments to examine the adsorption potential of five different GACs (lignite-based GAC, Coconut shell GAC, bituminous coal-based GAC, and some unknown raw material) for short- and long-chain PFAS removal from drinking water [80]. In the laboratory experiment, demineralized water (contaminated with four PFAS having an initial concentration of 6 μg/L: PFOA, PFBA, PFHxA, and PFPeA) was used, whereas PFAS-contaminated groundwater was used for the field column trial in the Southern Germany waterworks, which was contaminated due to paper sludge. As the polarity of PFAS decreases with increasing chain length, GAC is more effective in adsorbing long-chain PFAS versus short-chain PFAS [81]. The adsorption of PFSA was higher compared to PFCA [82]. The results of the AC sorption experiments indicated that PFOA was the most readily adsorbed, followed by PFHxA, PFPeA, and finally PFBA. In the field column experiment, the long-chain PFOA, as well as the short-chain PFHpA and PFHxA, were both efficiently removed by GAC treatment. Nevertheless, adsorption by GAC is either ineffective for PFBA and PFPeA or leads to short operating durations of GAC adsorbers, causing expensive treatment costs [80]. In a water utility, Burkhardt et al. performed pilot research consisting of two parts: in the first, nine PFAS compounds were investigated; in the second, sixteen PFAS compounds were analyzed; and in both parts, five different GACs were used and pore and surface diffusion models were used for the prediction of outcomes [27]. The GAC treatment was shown to be effective in removing all the PFAS tested in this investigation. Nevertheless, the remediation effectiveness differed significantly among the two pilot phases, which were carried out at various periods of the year, demonstrating that the PFAS had considerably varying adsorption properties that also fluctuated periodically. Granular activated carbon (GAC) was shown to be an effective remediation approach for PFAS removal in this pilot study, meeting the treatment objectives of the utility that conducted the research [27]. Mini-column studies at the research lab level were used to assess the effectiveness of GAC in removing PFOA and PFOS, two typical persistent PFAS, from drinkable water at three separate treatment facilities in Ontario, Canada [83]. Column studies were performed utilizing both sterilized and unsterilized GAC to differentiate between the adsorption and the possible degradability of PFOA and PFOS over the GAC. GAC service duration, temperature, influent concentration, and empty bed contact time (EBCT) were studied for their impact on PFOA and PFOS removal. The GAC samples were found to eradicate 40–55% of PFOS and 20–40% of PFOA, apart from Plant B, where removal of PFOA was insignificant due to 9-year-old GAC, which might be due to natural fouling of organic matter (OM) on the GAC surface. These outcomes are of great importance as they exhibited that the old GAC in drinking water treatment plants with continuous operation for taste and smell regulation were able to remove PFAS without any modification or further action [83].
To sum up, activated carbon is an efficient and common approach for treating water contaminated with PFAS (per- and polyfluoroalkyl compounds). Many variables influence how effectively AC filters remove PFAS from water. These include the type of AC employed, the characteristics of PFAS, and the water quality criteria. The possibility of secondary contamination makes it crucial to give careful thought to how AC laden with PFAS will be disposed of. The downsides of AC, such as the necessity to renew the carbon medium and the possibility of PFAS desorption, are outweighed by its benefits.

2.1.3. Metal-Organic Frameworks

Metal-organic frameworks (MOFs) are crystalline materials with a porous structure made up of inorganic nodes and organic linkers that self-assemble into three-dimensional lattices via coordination bonding [84]. These materials attractive properties include their amenability to modification at multiple levels (nodes and linkers), their tendency toward crystalline structure, and the existence of empty spaces, which have all led to numerous applications that follow the concepts of sorption processes, such as catalysis, separation, and filtration. The elimination of PFAS from water is a promising new use for MOFs [85]. The pore volume and surface area of MOFs are very high, up to 3.9 cm3/g and up to 7130 m2/g, respectively. MOFs are promising options for adsorption-based applications requiring rapid diffusion and kinetics due to their adaptable pore diameters (pore sizes up to 10 nm) and functionalities [86,87,88]. Further study is warranted to enhance the stability of this class of adsorbents because of the low water stability of MOFs, which might be a concern when employing these materials for water treatment. Zeolitic imidazolate framework (ZIF) and chromium (III) terephthalate metal-organic framework (MIL-101) are two examples of water-stabilized porous MOFs that have recently emerged and are showing tremendous promise for use in water treatment operations [89,90]. The adsorption process with MOFs is highly dependent on their physicochemical and geometric parameters, such as morphology, pore size, interactions, shape, and surface area, all of which are influenced by the MOF synthesis technique [91,92].
Chang et al. fabricated a water-stable MOF called PCN-222 using zirconium tetroxide (7.3 mg) as a metal source and tetrakis (4-carboxyphenyl) porphyrim (4.5 mg) as a ligand [90]. The PCN-22 exhibited a maximum adsorption potential of 2257 mg/g for PFOA removal and fitted well to the Langmuir isotherm, which favors monolayer adsorption. The pH of PCN-22 was maintained around 5 to maintain the positive charge on PCN-222. Further increases in pH cause a decrease in the adsorption potential of PCN-222 for PFAS due to the deprotonation of PFOA. The adsorption of PFOA increased with temperature, suggesting the reaction nature is endothermic. The PFOA were adsorbed on PCN-222 via hydrogen bond formation, electrostatic attraction, and hydrophobic attraction.
In another study, a MOF-808 was fabricated using zirconium chloride for nodes and H3BTC ligands for the removal of perfluorobutanesulfonate (PFBS) by senior-level students in a capstone course. Students were able to swiftly synthesize MOF-808 with a high yield in the chemical lab and characterize it using X-ray diffraction (XRD), Thermogravimetric analysis (TGA), and Diffuse Reflectance Infrared Fourier Transforms Spectroscopy (DRIFTS). Adsorption isotherms for PFBS on MOF-808 samples generated by students range from 2.8(8) to 10−4 M PFBS. When the dilution factor is included, the adsorbate concentration of PFBS rises to 3.0(8) 10−4 M [93]. The 2 g of Cr(NO3)3.9H2O, 0.83 g of H2BDC, and 0.15 mL of HF were mixed together in 70 mL of distilled water to create MIL-101 (Cr). The MIL-101 (Cr)@AC was fabricated by mixing 50 mg of AC in the precursor solution. To make MIL-101(Cr)-NH2, 0.4 g of NaOH, 0.72 g of 2-Amino terephthalic acid, and 1.6 g of Cr(NO3)3.9H2O were dissolved in DI water (30 mL) and mixed for 15 min at 160 °C for 16 h. In FTIR outcomes, the C-H bond at 1104 cm−1, C=O bond at 1650 cm−1, and 1620 cm−1 (C=C bond) displayed greater peak intensity compared to MIL-101 (Cr), indicating contact between ligand and AC, but no new functional group was introduced by AC’s interaction with MIL-101 (Cr). All the adsorbents showed an increase in adsorption from pH 2 to 4, followed by a reduction from pH 4 to 10. A comparison of AC’s adsorption effectiveness across a pH scale of 2–10 revealed no discernible differences. This might be because AC maintains a constant surface charge across the specified pH range. The highest adsorption was exhibited by AC (28.248 mg/g), followed by MIL-101(Cr)@AC (25.706 mg/g) (Table 2). The most prominent adsorption interactions for AC and MOFs were hydrophobic-hydrophobic and electrostatic, respectively [91].
Molecular simulations were used to examine the efficacy of various fluorine-functionalization techniques on MOFs for the removal of PFAS from water. The first class consists of fluorinated anion-functionalized MOFs, with AF6−2 anions (A = Si or Ti) serving as the connecting ligands. The second group consists of substitutions made using trifluoromethyl or fluorine. Finally, perfluoroalkane graft-functionalized MOFs are the third kind of fluorine-functionalized MOFs in this study. Of all the materials studied in this work, two of them, TIFSIX-1-Cu and Zn(4,4′-bpy)2(SiF6), had the greatest Henry’s Law Coefficients (HLCs) for PFOA. The results of this investigation show that fluorination is an effective method for removing PFAS from water resources because it improves the MOF-PFOA interactions [102]. GenX has lesser adsorption on AC than other legacy PFAS like PFOA (Wang et al., 2019). Commercialization of carbon-based adsorption technologies is hindered by issues with adsorbent regeneration, tuning adsorbents with favorable characteristics, and OM competition. Commercialization of carbon-based adsorption technologies is hindered by issues with OM competition, adsorbent renewal, and the design of adsorbents with the necessary surface chemistry/properties [103,104].
The large surface area, customizable pore widths, and metal coordination sites of MOFs make them a desirable option for removing PFAS from water sources that have been polluted. The short-chain PFAS have received less attention from researchers due to their less absorbable and more persistent nature. To remove PFAS more effectively, further study is required to enhance MOF features, such as enhancing selectivity for particular PFAS molecules and strengthening stability under different environmental conditions.

2.1.4. Ion Exchange Resins

Polymers’ strong adsorption capabilities make them attractive for PFAS elimination. Investigation into PFAS remediation utilizing ion exchange resin (IXR) has risen dramatically in the past few years in consideration of the material’s established adsorption potential for PFAS [105]. Research on PFAS adsorption using anion exchange polymers or resins (AER) has been conducted most often. The adsorption mechanism of PFAS by AER is based on hydrophobic and selective ion-exchange interactions [106]. Chow et al. conducted a pilot study in which they examined the adsorption behaviors of PFAS by GAC and two strong-base gel anion exchange resin (AER) columns that were operating in parallel to clean up contaminated groundwater over the course of 441 days. In terms of mass, GAC was able to hold onto 85 μg PFAS/g adsorbent, with 71% PFSA and 28% PFCA. The AER1 and AER2 adsorbed 441 and 460 μg/g of PFAS. The PFCA adsorption percentage was 11–12%, while PFSA adsorption potential was 88–89% by AER, which is seven times more than PFSA. The amount of PFAS absorbed by AER was more than five times that of GAC. Gel AERs performed slightly better than GAC for long-chain PFCA and significantly better for all PFSA in terms of mass adsorption [107]. In another experiment, Pluronic® F-127 and Abberlite® IRA910 beads in quartz sands were used as an ion-exchange resin for the adsorption of six PFAS (PFBS, PFOS, PFHpA, PFOA, PFHxS, and PFNA) in a low concentration range up to 100 μg/L and a high concentration range from 50 to 500 μg/L in a single and multi-solute system. The Kf value for PFOS in a multi-solute system is 404 L/g. At the start, the adsorption of PFHxS and PFNA increased at aqueous concentrations less than 5 mg/L and remained constant till 40 mg/L (3.89% PFNA and 23.6% PFHxS), then decreased at a higher equilibrium concentration (100 mg/L). Studies in both batch reactors and columns showed that at higher concentrations (60 mg/L total), longer-chain length PFAS, notably PFOS, were predominantly adsorbing. Adsorbed levels of all six PFAS were lower in the multi-solute system compared to those obtained from trials conducted with a single solute system [105].
To remove GenX, and two additional perfluorinated ether acids (PFEAs), researchers utilized a highly basic anion exchange resin (Purolite® A860) in both surface and recycled water. Isothermal tests included dissolving 10–1000 mg of resins in 1 L of water sample with 0.1–100 g GenX and 5 mg C/L starting concentration of natural fluids and letting the mixture stand at room temperature for 24 h (equilibrium). Following 24 h of treatment, the resins were removed from the water with 0.45 m pre-rinsed syringe filters (Fisher Scientific, SLHV033RS, Nepean, ON, Canada) before the final PFAS concentrations were measured. Uptake is represented as a linear function of compound charge density, with the largest uptake occurring for the utmost charged portion (PFMOPrA > PFMOBA > GenX (p 0.05)). This is consistent with earlier research on IX, which found that organic molecules with larger charge densities were selectively eliminated. With a trend of 1.02–0.04 (R2 > 0.98), they find that the intake of all PFEAS is linearly related to the analogous outflow of chloride ions. This data suggests that extremely basic macroporous IX resin (A860) absorbs GenX and other perfluorinated ether acids by a mechanism similar to that of ion exchange (IX), with just a little contribution from hydrophobic interactions and other physical variables [104]. Reversible addition-fragmentation chain transfer (RAFT) polymerization was used to create a wide variety of PFPE-containing polymers, both ionic and cationic, and then bonded onto magnetic iron oxide nanoparticles (IONPs, Fe3O4). The quaternized ammonium monomer of 2 and 9 and the degree of polymerization (DP) of oligo (ethylene glycol) methyl ether acrylate (OEGA), respectively, were found to provide the outstanding outcome in the separation of numerous PFAS, and P2-9+@IONPs, the grafted polymer sorbent consisting of the greatest percentage of cationic groups (39 wt%), was the most effective (Figure 4).
The addition of the quaternized ammonium groups dramatically increased the PFAS removals from all three quaternized polymer-grafted IONPs (P2-9+@IONPs, P4-6+@IONPs, and P6-3+@IONPs). It is believed that electrostatic attractions between the polymer’s quaternized ammonium group and the PFAS’s charged head group are responsible for this phenomenon, which is notably evident in the elimination of GenX and other short-chain PFAS. P2-9+@IONPs demonstrate quick capture of GenX with an outstanding removal rate, capturing >99% in 30 s and increasing to 99.5% in less than three minutes [108]. IX’s performance is highly impacted by the presence of co-contaminants in the feed solution, and the process generates secondary waste during the regeneration phase. Furthermore, the resin is sensitive to oxidation. Overall, ion exchange resin is a viable option for the efficient treatment of water contaminated with PFAS. However, choosing the right resin is essential, since long-chain PFAS have a higher affinity for the resin and need the use of specific resins for their removal. Furthermore, the resin has a limited capacity and must be renewed whenever it reaches saturation.

2.2. Membrane Separation Technology

Nanofiltration (NF) and reverse osmosis (RO) are two membrane-based treatment methods that depend on the elimination of PFAS by charge and size exclusion, or sorption onto the membrane polymers. PFAS rejection via membranes may be as high as 99% in certain cases, but their use is limited by energy requirements and a limitation of molecular selectivity. The summary of membrane filtration and reverse osmosis for PFAS removal is represented in Table 3.
The current treatments used combined approaches and hybrid membrane technologies for the elimination and destruction of PFAS via adsorption, ion exchange, and membrane separation. In membrane separation, nanofiltration and reverse osmosis are efficient techniques for PFAS removal from wastewater. Novel membranes consisting of nanomaterials were the most efficient in removing PFAS. The cost of membrane treatment and ion exchange has been found to be very effective. Irrespective of the chain length or molecular structure of PFAS, Ross et al. reported that RO and NF are very efficient at eliminating several of them, particularly PFAA intermediates [120]. They highlighted the significant challenges inherent in these methods, such as the high cost of large-scale implementation, the difficulty of dealing with suspended particles and water geochemistry, and the difficulty of keeping membranes healthy, especially in groundwater applications. Figure 5 displays the NF/RO membranes that have been experimentally indicated to be effective in the removal of PFAS from water.
Nanofiltration and reverse osmosis have been proven to be effective in the removal of per- and polyfluoroalkyl compounds from water, and their efficacy has been evaluated by [121]. The NF membrane used for the treatment of PFAS had the molecular weight cut-off (MWCO) and pore size in the range of 90–1000 Da and 0.3–2.1 nm, while the RO membrane was in the range of ≤100 Da and ~0.1 nm. The highest removal rates were shown by the most commercial NF membranes, about 90.0% to 99.0% for common PFOS and PFOA. On the other hand, removal by RO membrane showed 96% to 100% removal rates which is greater as compared to NF membrane. It was concluded that overall, the eradication of PFAS via RO and NF membrane depends upon size exclusion and electrostatic and hydrophobic interactions. The process of PFAS elimination is also influenced by the membrane characteristics and physiochemical properties of PFAS. Liu et al. reviewed the effectiveness of nanofiltration and reverse osmosis technologies for the removal of per- and polyfluoroalkyl substances from water. The NF membrane used for the treatment of PFAS had a molecular weight cut-off and pore size in the range of 90–1000 Da and 0.3–2.1 nm, while the RO membrane was in the range of ≤100 Da and ~0.1 nm. The highest removal rates were shown by the most commercial NF membranes about 90.0% to 99.0% for common PFOS and PFOA. On the other hand, removal by RO membrane showed 96% to ~100% removal rates which is higher than NF membrane. It was concluded that overall, the removal of PFAS by RO and NF membrane depends upon the size exclusion and electrostatic and hydrophobic interactions. The process of PFAS removal is also affected by the membrane characteristics and physiochemical properties of PFAS. In another study, Zhi et al. used five different types of commercial membranes to analyze their removal efficiency for three perfluoroalkyl ether acids (PFEA), seven PFAS, and two fluorotelomer sulfonates (FTS) [118]. Rejection rates for GenX (C5) were 11.6% greater compared to PFHxA (C5) on average (p < 0.05). When evaluating the effectiveness of the investigated membranes in removing PFOA and GenX, PFOA removal was shown to be more consistent. In one study, it was shown that powdered activated carbon (PAC) made from wood using 60 mg/L of dosage was able to successfully adsorb 80% of PFOA, >95% of PFOS, and 30% of GenX. Even at a dosage of 100 mg/L carbon, short-chain PFEA like perfluoro-3,5-dioxahexanoic acid (PFO2HxA) were almost completely nonabsorbable. Short-chain chemicals (perfluoroalkyl-chain lengths of 3 and 4) including PFMOBA, PFBS, 4:2FTSA, PFMOPrA, and PFBA were effectively removed by DK and NF90 membranes. To treat certain types of PFAS, the NF method emerges as a more viable approach over GAC adsorption.
While the RO and NF are effective in removing PFAS, they do have certain drawbacks. Energy input is substantial, and the process may be costly to run. To minimize performance degradation due to fouling, the membranes need to be cleaned and maintained often. It is important to consider the high energy consumption and maintenance expenses connected with the technology. Instead of relying only on RO or NF technologies, it may be preferable to use these technologies as a pretreatment process in the treatment system for fluids with complex mixtures like groundwater and wastewater.

2.3. Fractionation Techniques

In the physiochemical process known as foam fractionation (FF), pollutants that are surface-active compounds may be extracted from a growing foam and purified (Figure 6). By compressing the accumulated foam, a waste stream rich in surface-active contaminants may be created. Foam fractionation seems ideally adapted for the isolation of these chemicals from surface and groundwater systems because of the considerable surface activity of many distinct kinds of PFAS molecules, with the compressed foam resulting in a low proportion of the remaining fluid waste stream. Recently, foam fractionation has received more attention, which has led to the discovery of novel uses for this technique [122]. Surface-active foam fractionation (SAFF) was used in a field study by Burns et al. (2021) to clean up PFAS in groundwater at an Australian location with a treatment capacity of up to 250 m3/day. The SAFF process was more effective at removing PFAS with relatively long fluorocarbon chains and greater air/water adsorption ratios because these PFAS had a stronger affinity for the bubble surface [123].
In another study, a bubble fractionation technique was used to analyze the influence of different environmental parameters including temperature, salinity, and aeration on PFAS removal in the absence of a co-surfactant. Before subjecting contaminated waters to high-energy PFAS destructive procedures, aeration removal may be a viable option as a non-destructive means of separation/concentration of amphiphilic (i.e., surface-active) chemicals. Aeration causes gas bubbles to develop in a liquid, creating a gas-liquid interface where the desired surface-active molecules may bind.
As the bubbles float to the top of the liquid column, they are easily isolated from the fluid being processed [124,125]. The FF technique has benefits over more traditional preconcentration methods like adsorption on AC or membrane filtration because of its minimal consumables use and resistance to complicated and changing water matrices [123,126]. Short-chain or non-amphiphilic PFAS are more difficult to remove since this method mainly effective for surface-active PFAS [127]. PFDS, PFHpS, PFHpA, and PFNS were likewise nearly completely eradicated after 60 min of aeration, whereas PFHxA, PFBS, and PFPeS were removed to varying degrees. At the conclusion of the 60-min aeration treatment, the total PFAS species were removed with efficiencies of 95.7% in the room-temperature condition, 97% in the low-temperature condition, and 93% in the high-temperature condition. Although aeration proved to be an effective method of PFAS treatment, the concentrated PFAS foamate that is generated must be further treated in order to degrade the PFAS compounds and remove them from waste streams [125]. Buckley et al. conducted a study wherein sodium dodecyl sulfate was utilized as a hydrocarbon surfactant to investigate the impact of salts on foam fractionation of PFAS species. It was discovered that the addition of salts improved the efficiency of PFAS removal due to an increase in surfactant adsorption at the air-water interface. However, it should be noted that the inclusion of extra chemicals may restrict the locations where the treated product water may be dumped [122]. Foam fractionation was utilized by Smith et al. to concentrate PFAS in groundwater, and then electrochemical oxidation (ECO) was employed to destroy the PFAS [127].
There was a 10 L min−1 air flow rate, 20-min residence duration, and a 10% collected foam percentage. The system ran constantly for 9 h, and samples of influent, effluent, and foam (at 2, 4, 6, and 8 h) were taken for PFAS measurement. In a 20 L circulation tank with a combined total stainless-steel cathodic surface area of 9200 cm2 and BDD anodic, 9 h batch tests were performed in triplicate for electrochemical oxidation. Ninety percent of PFAS were removed from groundwater and leachate during foam fractionation. The electrochemical degradation technique degraded PFAS by 84% after 540 min long-chain PFAS was eliminated more efficiently than short-chain PFAS (mean 22%) due to their greater air–water interface adsorption ratios [127].
Foam fractionation works well on its own, but when ozone is added, the process becomes much more efficient (Ozofractionation). To create ozofractionation, the ozofractionation catalyzed reagent addition (OCRA) procedure oxidizes organic pollutants chemically while simultaneously producing intense foam fractions that can be easily isolated from the final treated water product. Foam fractionation can efficiently remove PFAS in the liquid media because the perfluoroalkyl groups shift to the gas-fluid interface, and ozone oxidizes and degrades organic molecules. Although ozone (O3) has only been explored in small-scale experiments for the cleanup of PFAS in water, it has been shown to achieve removal rates > 97% in a sensible procedure time [128]. Full-scale assessment employing ozofractionation for PFAS elimination proved to be successful, as reported by Horst et al. (2018). In order to meet stricter emission regulations, polishing treatment steps, including adsorption and filtration, were included [129]. O’Connor et al. utilized ozofractionation for the treatment of PFAS-contaminated marine fish hatcheries. OCRA is a revolutionary technique designed by the University of Newcastle and its industrial partner, Evocra Pty Ltd., for the removal of PFAS, acid mine drainage, and hydrocarbons from water. The OCRA technique is a successful fractionation method that controls the pH and oxidation-reduction potential of a solvent by using high quantities of ozone. Experiments conducted at both sites showed that commonly existing foam fractionation systems treating polluted saltwater with air or ozone could eliminate more than 90% of the PFOS and PFOA present. Ozone is widely employed in aquaculture systems today to disinfect saltwater, and research indicates that it is effective in removing PFAS. Nonetheless, ozone usage requires caution, and future studies are necessary to properly optimize its use in fish production [130].

2.4. Destruction Techniques

Destructive technologies are a kind of remediation that may, in most cases, decompose pollutants into safer byproducts. There are several destructive technologies that have been shown to be effective in degrading PFAS, with removal efficiencies ranging from 60% to 100%, including electric field-nanofiltration, electrochemical water treatment technology, direct current plasma, plasma-based technologies, catalytic ozonation, etc. A vast variety of PFAS compounds have been found to be destroyed and defluorinated to a great degree using hydrothermal alkaline treatment (HALT), mainly with NaOH. Compressed water with a high pH (>14), high pressure (25 MPa), and high temperature (350 °C) may be used to convert PFAS into innocuous fluoride salts if the kinetics are severe enough [131]. Although perfluorocarboxylic acids (PFCAs) are destroyed more quickly, perfluoroalkyl sulfonic acids (PFSAs) need lengthy residence periods (30 min) and large hydroxide loadings (e.g., 5 M-NaOH) in batch HALT reactors to attain greater percentages (>90%) of degradation and defluorination [132,133,134]. Pinkard et al. used three different concentrations of NaOH: 5 M-NaOH (200 g/L), 0.1 M-NaOH (4 g/L), and 1 M-NaOH (40 g/L) during the HALT. The inner temperature of the continuous reactor is maintained at 350 °C throughout operation. Changing the pump’s flow rate at a constant internal pressure of 25 Mpa causes a change in the reaction’s residence time from 1.6 to 10 min. This work seems to show that the residence periods needed to accomplish >99% annihilation of parent PFAS species may be reduced using the continuous flow HALT reactor setup. Residence periods of 10 min. resulted in >99% annihilation of the most resistant PFSA species, including PFHxS, PFPeS, and PFBS, whilst PFOS levels were reduced by >99.99% after only 1.6 min of processing. Based on these findings, it seems that continuous flow HALT technology may be expanded to practical throughputs [134].
Novel treatment techniques that can be effective against PFAS at room temperature and pressure are desperately needed. Yang et al. used a nonthermal ball mining technique for the destruction of PFAS. In a jar (100 mL) containing zirconium and stainless-steel balls, the PFAS-contaminated sediment samples collected from Schriever Air Force Base (Colorado Springs, CO, USA) were combined with co-milling reagents (KOH and boron nitride) and spun at a rate of 580 revolutions per minute for different time intervals. After that, the samples were analyzed using LC-MS/MS. After 6 h of treatment, almost 80% of the 21 PFAS were eliminated. It was found that the reaction pathways included both fluorination of boron nitride and piezo-electrochemical oxidation of PFAS [135].
Austin et al. contrasted the efficacy of continuous flow supercritical water oxidation (SCWO) for PFSAs and PFCAs [136]. The destruction seemed directly proportional to the rising temperature. At 610–650 °C, the destruction and removal efficiencies (DRE) were found to be more than 99.999%. Trace amounts of PFOS (1.66 g/L) were the only PFAS found in the effluent at T = 650 °C, which is equivalent to a DRE of >99.999%. Keep in mind that at lower working temperatures, PFHpS seems to have a lesser DRE compared to PFOS. It was discovered that both PFSAs seem to be more persistent than PFCAs. This is consistent with the comparison of PFOS and PFOA performed by [137]. Li et al. have described the decomposition and mineralization of PFOS and PFOA across a wide temperature spectrum (T = 410–650 °C). The PFOS DRE was found to be more than 99.999% at a residence duration <30 s and T = 650 °C. In the wastewater discharges from gaseous channels, PFOS are converted into a pool of PFCAs, including 1H perfluoroalkane, at lower temperatures (T = 420–600 °C). It is important to deal with the toxic byproducts formed during and after the treatment. Negative public health and environmental repercussions occur if toxic liquid and gaseous wastes are not contained.
The possible drawbacks of traditional membrane filtration may be overcome by incorporating electrical activity into the separation process. As PFOA has a strongly electronegative carbon-fluorine bond, its calculated pKa shows that it is negatively charged in water across common pH ranges. This makes it fair to hypothesize that PFOA would be impacted by an electric field. The literature reports that putting a direct-current electrical field over a microfiltration membrane dramatically increases PFOA and PFOS elimination from 0% to 70% [138,139]. The elimination of PFOA was investigated using a crossflow membrane cell constructed from a porous SnO2-Sb anode and a porous titanium cathode in an electric field nanofiltration [140]. A cross-sectional element mapping overlay of Sn and Ti, proving that the SnO2 layer was effectively deposited on the Ti substrate and demonstrating its coherence and density.
Researchers used a voltage gradient between 0 and 90 V while maintaining a constant transmembrane pressure of 4.14 bar (60 psi). From 0 to 30 V, the electrical voltage increased the PFOA rejection from 45% to 97%. Compared to non-saline feed, the saline feed has a reduced PFOA removal rate with inconsistent filtration effectiveness, ranging from 55% to 80%, which might be due to the bubble formation during water electrolysis, deteriorating the electric field force. Expanding the applicability of electrochemical water treatment technology, electric field-assisted nanofiltration offers a revolutionary separation strategy for efficiently eliminating ionized organic contamination [140]. Microwave power supply, direct and alternate current, radiofrequency sources, and the pulsed discharge technique are only a few of the ways that non-thermal plasma may be generated for use in water and wastewater purification. The removal of PFOA and PFOS by direct current plasma has been stated to be 100%. Hence, plasma-based technologies are highly efficient in filtering PFAS out of both natural and artificial water supplies. Its efficiency, though, suffers when it encounters inorganic and organic contaminants from outside sources. In addition, the plasma technique produces harmful by-products. Notwithstanding its limitations, this technology requires more investigation on a large-scale, real-world scale [141].
Lashuk et al. evaluated the catalytic ozonation removal potential for five different PFAS including GenX in the presence of four catalysts fabricated using titanium in different ratios of butaoxide, mixed with t-butanol (WO3/TiO2 catalyst) [142]. Photocatalytic ozonation removed 4–24% of PFAS, whereas photolysis destroyed 4–13%, ozone photolysis eliminated 7–16%, and photolysis removed 2–4%. In general, after 4 h, 6:2 Fluorotelomer sulfonic acid (6:2 FTSA) and PFOA were removed at higher rates than PFHxS, PFBS, and GenX. When comparing the short-chain PFAS, photocatalytic treatments removed approximately twice as much PFHxS (6-carbon chain) as PFBS (4-carbon chain). A developing issue due to their increased usage in industry and presence in the environment, poorer removal percentages for shorter-chain PFAS have been reported by researchers. GenX, the new PFOA replacement material, proved very resistant to all treatments. This is in line with the latest results of [42], who found that the CF3 group on the alpha carbon of GenX acts as a protective group, making it resistant to advanced oxidative processes [142].
The most extensively researched approaches for PFAS destruction are electrochemical advanced oxidation processes (EAOPs), which produce highly susceptible hydroxyl radicals (OH•) with a high redox potential (E = 2.8 V versus SHE). They may generate immiscible alcohol compounds to break C–F bonds. Due to its efficacy and simplicity in current water treatment facilities, anodic oxidation is a popular EAOP [143]. Using ultraviolet light and persulfate, [144] investigated the oxidation of hexafluoropropylene oxide tetramer acid (HFPO-TeA), reporting that HFPO-TeA depolymerizes to Hexafluoropropylene oxide trimer acid (HFPO-TA), which in turn depolymerizes to HFPO-DA.
The effectiveness of boron-doped diamond (BDD) anodes in the breakdown and defluorination of HFPO-DA was studied by [20]. Investigations involving ECO were conducted in a polypropylene electrochemical cell with a single chamber. The operating anode was a boron-doped diamond electrode with an efficient surface area of 22.6 cm2, and the cathode was a platinum mesh set up adjacent to the anode on each side, with a distance of 10 mm between the electrodes. In order to determine the fluoride content, an ion-chromatograph system (DX-120, DIONEX, Lane Cove West, NSW, Australia) was used. This system included a separation column (IonPak AS12A, 200 mm, and 4 mm) and a suppressed conductivity detector. The flow rate was adjusted to 1.3 mL/min and the mobile phase was an aqueous solution of 1 mM Na2CO3 and 1 mM NaHCO3. At 4 h, the removal rate was nearly comparable for the two electrolytes, at 91% (Na2SO4) and 93% (NaClO3). The observed reaction rates (kapp) in the Na2SO4 and NaClO4 electrolytes were 0.0106 and 0.0113 mL/min, respectively, indicating that the process followed pseudo-first-order kinetics.
After 4 h, defluorination efficiency was 80% in Na2SO4 electrolytes and 84% in NaClO4 electrolytes. Breakdown of HFPO-DA after 4 h was 96%, 92%, and 81% at 0.005, 0.01, and 0.05 M electrolyte concentrations, respectively (Figure 7, Table 4). The rate of HFPO-DA breakdown was shown to be determined by direct electron transfer. It is suggested that HFPO-DA degrades via the creation of three intermediates before finally mineralizing to CO2 and F (Figure 7) [20]. Unfortunately, the high upfront expenditures and ongoing operational and maintenance expenses of these technologies have prevented their widespread implementation on a practical scale.

3. Removal Mechanism

Understanding the mechanism of PFAS adsorption and destruction helps with both the assessment of removal techniques for prospective applications and the development of novel, more efficient removal techniques. According to Smaili and Ng, 34 investigations found PFAS sorption to be dominated by hydrophobic contacts, 19 by electrostatic interactions, and 48 by a mix of both [150] (Figure 8).
Other interactions, including hydrogen bonding, fluorine-fluorine interactions, and ligand exchange (although weakly), were shown to influence the adsorption of PFAS in eight more investigations. Research was conducted by Ma et al. to assess the influence of iron minerals in laboratory-scale parallel-constructed wetlands (CWs) for the elimination of PFOA and PFOS [151]. To remove PFOS and PFOA from CWs, iron minerals of different shapes and sizes were used as substrates, with the goal of optimizing microbial degradation and substrate efficiency. The maximum removal of PFOA and PFOS (57.2% and 63.9%) was exhibited by MC2 (Gravel:Magnetite, 1:2). The concentration of PFOA and PFOS in the effluents were decreased when iron minerals were added to CWs, relative to the control group. Their elimination is aided by electron transfer and denitrification, both of which are favorably linked with iron content (Figure 9) [151,152]. Moreover, because of the increased translocation factor of PFOA in CW, its removal was more successful than that of PFOS.
Two main pathways exist for emerging contaminants biodegradation by microalgae: metabolic degradation, in which the emerging contaminants act as a carbon source or electron giver/acceptor, and co-metabolism, in which the emerging contaminants are destroyed by enzymes activating other substrates [153].
Adsorption of PFAS onto carbon nanotubes (CNTs) is a complex process that has not yet been fully elucidated. Du et al. concluded in their review that hydrogen bonding interaction is not likely to be involved in the adsorption process of PFCAs on CNTs [70]. Recent research reveals that charge-assisted hydrogen bonding may react with the carboxyl group of the PFCAs and the functional group on the CNT surface to adsorb ionizable molecules on CNTs.
RCOO +H ++ -O-CNTs (RCOO…H…O-CNTs)
The carboxyl, hydroxyl, and amine functional groups on the CNTs surface also attract water molecules which form water molecule clusters on the CNTs surface, which could be the reason for negligible involvement of hydrogen bonding in PFCAs adsorption on CNTs due to the competitive adsorption. The hydrophobicity of the PFAS alkyl chain is proportional to the amount of PFAS adsorption on CNTs. PFOA, PFBA, and PFHxA all contain CF chains with different numbers of carbons (eight, four, and six, respectively). Adsorption studies using CNTs reveal that after 15 h, over 95% of PFOA had been absorbed using single-wall carbon nanotubes (SWCNTs), but PFBA (7.5%) was hardly eliminated [154,155,156]. Adsorption on carbon nanotubes increases with hydrophobicity and PFAS chain length [70]. In water, PFAS mostly exists in anionic forms and CNTs have anionic functional surface groups causing electrostatic repulsion and resisting PFAS adsorption. However, the attractive hydrophobicity of the CNTs surface is thought to outweigh the repelling electrostatic interactions [157]. The lack of surface hydrogen in most CNTs prevents them from forming H-bonds with the fluorine atoms in PFAS [156].
KD for perfluorocarbons of the PFCA and PFSA series increases with chain length because of the larger dispersion forces and solvophobicity that the perfluorinated chain may exert at the interface [158]. According to past research of PFAS adsorption on AC, the sulfonates adsorption is stronger than the carboxylates [63]. The different group characteristics of the charged CO2 and SO3 units explain why sulfonate is more attractive than carboxylate [110]. As compared to the carboxylate ion, the sulfonate ion has lower solvation because its charge is more delocalized. The sulfonate is preferentially favored in its surface-dispersion interactions, as dispersion is related to polarizability and the total of the polarizabilities of the atoms in their ground state [63]. In a study, Min et al. fabricated functionalized periodic mesoporous nanosilica (PMO) with different amine contents using a condensation method for the removal of a short-chain PFAS, perflourobutanoic acid (PFBA). The functionalized PMO material with the maximum fluoroalkyl group loading of 1.05 mmol/g (PMO-NHCF-3, removed PFBA with little interference from coexisting anions. Adding a fluoroalkyl group to an aminated organic silica may enhance nonelectrostatic interactions with PFBA, which can then increase PFBA adsorption onto bifunctional PMOs while decreasing inorganic anion (IA) competition. The wide contact angle of n-hexadecane on PMO-NHCF-3 (128°) further demonstrated the material’s significant oleophobicity, which is a unique feature of the fluoroalkyl group. As PFBA has a distinctive fluorous chemistry and a stiff structural C-F skeleton, fluorophilic interactions could further improve the attraction among PFBA and the bifunctional PMOs [159,160,161].
Mantripragada et al. produced and compared the removal efficiency of two types of nanofibrous membranes polyacrylonitrile (ESPAN) and amidoxime surface functionalized ESPAN (ASFPAN) for the removal of GenX from water [117]. The water contact angle of the ESPAN is 11 ± 52, making it hydrophobic. GenX adsorption on ESPAN showed hydrophobic contact regardless of the negative surface charges. Because of the high electronegativity of F, the CF chain of GenX may include strong dipoles of (δ+)C−F(δ−), whereas polyacrylonitrile (PAN) molecules may have strong dipoles of (δ+)C−N(δ−). These strong dipole-dipole forces might be the cause of GenX adsorption on ESPAN. More concentrations of amidoxime functional groups are present in ASFPAN molecules than in PAN, indicating that they may be positively charged in acidic environments. Adsorption of GenX onto ASFPAN molecules may be facilitated by electrostatic contact (Coulomb forces) between positively charged sites >C−N+< on ASFPAN molecules and the negatively charged COO groups on GenX. The surface hydrophilicity and electrostatic interaction contribute to a nearly twofold improvement in GenX eradication rate by the ASFPAN versus ESPAN [117].

4. Cost Analysis

It is estimated that for every 1 g of PFAS removed from a polluted site, the Global warming potential (GWP) increases by 83–122 kg of carbon dioxide equivalents (eq), and the cost increases by >75 United States dollars each gram of PFAS treated [75,162]. Murray et al. did the cost analysis of two PFAS treatment techniques, AC and IX. Treatment costs were compared to 10% and 50% breakthrough based on media choice and treatment goals by conducting a sensitivity analysis of operation and maintenance (O&M) expenses for GAC and IX projects [163]. Calculations show that compared to average-performing GAC, which needs a changeout around every other month, PFAA treatment by IX is at least 20% less costly (given by levelized media cost) throughout all unit cost points studied due to 16 times longer changeout period.
To reach a point of equity between GAC and IX, effectiveness must improve by 25% in GAC to reduce changeouts or unit cost must fall to less than $2.75/kg, while performance must decline in IX, leading to an increase in unit cost of more than $20.35/kg. There is still needed to develop efficient cost-effective disposal techniques for the used adsorbents for PFAS removal, for now, the AC and IX are piled up in the landfills. However, the value of porous carbon-based materials is far cheaper than that of many other absorbents. Chemically activated carbon made from nutshells costs roughly USD 1.82 per kilogram to produce, according to recent research [164]. Another study reported the chemical and physical activation of AC was between $0.96–4.22 and $0.96–1.54 [165,166]. A study has been conducted on the cost analysis of AER and GAC for GenX and PFAS. To treat water polluted with GenX or PFOA using GAC, the yearly cost per home is $224 or $85. There is no difference between GenX and PFOA in terms of direct or total capital expenses related to AE resin (Type I strong base polystyrenic gel). Operating and maintenance expenses for treating GenX and PFOA with AE resin are similar (US$156,594 and US$157,482, respectively) per year. Costs of operation and maintenance fluctuate somewhat from year to year owing to differences in GenX and PFOA removal rate of AE resin [22].
The GAC filters were installed in a Ridgewood, New Jersey car treatment plant which costs $3.5 million for installation and treats 1 million gallons/Day [167]. In comparison to synthetic materials (US$1500 ton−1 for activated carbon), utilizing biomaterials (US$246 ton−1 for activated biochar) could aid in cost savings [168]. The Sweeney water treatment plant was upgraded by adding 8 GAC filters in Wilmington, North Carolina having an installation cost of $46 million and an operational cost of $2.9 million for the treatment of 44 million gallons/Day [169]. As compared to other water treatment methods, RO is the most affordable ($2 m−3), even when the average cost of power is factored in (8.14 cents KW h−1) [170]. A new RO plant was installed in West Morgan-East Lawrence Water Authority, Decatur, Alabama for $30 million and costs $1 to clean 10 million gallons/Day [167,171].
Kanchanapiya and Tantisattayaku worked to analyze the excessive cost of removal of PFAS by using reverse osmosis from landfill leachate in Thailand [172]. Pretreatment and evaporation ponds were part of the two-pass reverse osmosis (RO) system used in this research. Using the water balance approach, they calculated the total volume of 111 sanitary landfills and analyzed the project’s financial viability using the net present cost (NPC) method. The findings indicated that the annual LL volume was about equal to 17.5 cubic meters. The RO system’s NPCs were calculated to be $577.9 million USD for a setup that included an evaporation pond, and $391.9 million USD for one that did without. The systems with an evaporation pond had a treatment unit cost ranging between $1.72 and $2.71 per cubic meter, whereas those without had a cost of between $1.06 and $2.09 per cubic meter.
The EPA presents the treatment technologies (GAC, IX, and RO) and costs for PFAS removal from water in an Annual Conference 2020. The total annual cost for PFAS treatment using GAC, RO, and IX is approximately $48,000,000, $49,000,000, and $50,000,000, respectively for more than 2000 households. Estimated laboratory testing expenses for PFAS are $250 per sample. A tank having 30 lbs of GAC (1 cu ft) for whole house water costs $539 whereas two large tanks for whole house water filter having 240 lbs of GAC costs $3990. A small GAC system weighs 200 lbs and costs $1200 before installation, and RO modification for Point-of-Entry use weighs 150 lbs and costs $2000 before installation [173].
Cost-benefit analyses of oxidation-reduction, sonochemistry, and biological technologies are extremely scarce, and further study is needed to make these approaches scalable. The processes of conversion and the viability of these approaches for actual wastewater need to be demonstrated experimentally in more depth. It is well-documented that the use of multimodal approaches, such as the use of dual-frequency systems for sonochemistry, may reduce expenses by 23%, from $10 m−3 to $7.5 m−3 [166,170].

5. Conclusions and Future Recommendations

An in-depth analysis of the current developments in PFAS removal and degradation technologies is reported in this review article. Since the ban on longer-chain PFAS chemicals, investigations show that the shorter-chain varieties have become more prevalent in the environment. The increased amount of PFAS have been detected in the effluents of WWTPs compared to influents which is possibly due to the breakdown of PFOA and other long-chain precursors into shorter-chain derivatives. Our analysis of available treatment options suggests that short-chain PFAS are more difficult to eliminate than long-chain PFAS. This disparity may be attributable to the fact that ionic interactions are more important for the sorption of short-chain PFAS compared to hydrophobic ones for long-chain PFAS. The MOFs exhibited higher adsorption capacity compared to biochar and AC. Reverse osmosis and nanofiltration are both found to be effective techniques for PFAS removal, however, the need to clean or change the filter and degradation of the membrane is costly. The destruction technique is the most effective one, but its use is limited due to the formation of PFAS precursors during treatment techniques. Therefore, it is recommended to use combined treatment techniques for PFAS removal. Current studies have yielded several suggestions on how to proceed with PFAS removal in the future.
No research has been conducted on the use of MOF for the elimination of GenX. Larger surface areas for adsorbents may be achieved in metal-organic frameworks due to the presence of positively charged metals in their crystallized framework and the uniformity of their pore structure. Due to their electrostatic and hydrophobic interactions, they may be excellent fits for eliminating GenX.
Short-chain PFAS were less likely to be adsorbed by activated carbon than long-chain PFAS. The AC must be tailored via various methods and examined to remove short-chain PFAS and GenX.
New PFAS compounds, including fluorotelomers, require further study in wastewater because of their widespread use in industries like fluorinated impregnations, metal plating, oil production, and food packaging.
Future research might focus on optimizing electrode design and flow-cell arrangement to boost electrosorption kinetics and increase mass transfer in electrochemical reactors. Further mechanistic research is required to fully understand the mechanisms involved in short-chain degradation using a variety of redox-electrode materials.
Byproducts from several PFAS cleanup methods may be hazardous to human and environmental health. There should be more investigation into the hazards of various PFAS-removal technology.
It is challenging to create analytical methods that might identify the kinds and amounts of PFAS since the molecular identities of several PFAS on the worldwide market are still uncertain and PFAS are found in the environment at extremely low quantities.
Future research should include evaluating a combination of PFAS instead of single compounds due to the heterogeneity in degradation efficiency amongst the different PFAS studied.
The destruction techniques with combined treatment technique targeting the elimination of long and short-chain PFAS is suggested after reviewing the relevant technologies. The implementation expenses of such a strategy, though, continue to be a barrier to its widespread use. Consequently, it is essential to reduce the overall cost of the adsorption process, which includes the cost of regeneration.

Author Contributions

Conceptualization, R.A. and E.B.H.; methodology, A.I.; writing—original draft preparation, R.A. and W.S.; writing—review and editing, R.A. and A.I.; supervision, E.B.H. All authors have read and agreed to the published version of the manuscript.

Funding

This publication is based upon work supported by the McIntire Stennis project under accession number 70011735.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No data was used for the research described in the article.

Acknowledgments

This manuscript is publication #SB1115 of the Sustainable Bioproducts, Mississippi State University. This publication is a contribution of the Forest and Wildlife Research Center, Mississippi State University.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AbbreviationsExplanation
6:2 FTSA6:2 Fluorotelomer sulfonic acid
ACActivated carbon
AERAnion exchange resins
AFFFAirforce aqueous film forming foam
BACBiological activated carbon
BDDBoron-doped diamond
BSBiosolids
CNMsCarbon nanomaterials
CNTsCarbon Nanotubes
CWsConstructed wetlands
DPDegree of polymerization
DREDestruction and removal efficiencies
DRIFTSDiffuse Reflectance Infrared Fourier Transforms Spectroscopy
DWTPsDrinking water treatment plant
EAOPsElectrochemical advanced oxidation processes
EBCTEmpty bed contact time
ECFElectrochemical fluorination
ECOElectrochemical oxidation
EPAEnvironmental Protection Agency
FFFoam fractionation
GACGranular activated carbon
GWPGlobal warming potential
HALTHydrothermal alkaline treatment
HFPO-DAHexafluoropropylene oxide dimer acid
HFPO-TAHexafluoropropylene oxide trimer acid
HFPO-TeAhexafluoropropylene oxide tetramer acid
HLCsHenry’s Law Coefficients
IAInorganic anion
IXIon exchange
IXRIon exchange resin
KdAdsorption-desorption distribution coefficient
MIL-101Chromium (III) terephthalate metal-organic framework
MOFsMetal-organic frameworks
MWCOMolecular weight cut-off
NFNanofiltration
NPCNet present cost
O3-BACOzone biological activated carbon
OCRAOzofractionation catalyzed reagent addition
OEGAOligo (ethylene glycol)methyl ether acrylate
OMOrganic matter
PACPowdered activated carbon
PANPolyacrylonitrile nanofibrous membrane
PBDEsPolybrominated diphenyl ethers
PFASPer and polyfluoroalkyl substances
PFBAPerfluorobutanoic acid
PFBSPerfluorobutane sulfonate
PFCAsPerfluorocarboxylic acids
PFDAPerfluorodecanoic acid
PFDSPerfluorodecanesulfonic acid
PFEA3 perfluoroalkyl ether acids
PFEA3 Perfluoroalkyl ether acids
PFECAsPer- or polyfluoroalkyl ether carboxylic acid
PFHpAPerfluoroheptanoic acid
PFHpSPerfluoroheptanesulfonic acid
PFHxAPerfluorohexanoic acid
PFHxSPerfluorohexanesulfonic acid
PFMOBAPerfluoro-(4-methoxybutanoic) acid
PFMOPrAPerfluoro-2-methoxypropanoic acid
PFNAPerfluorononaoic acid
PFNSPerfluorononanesulfonic acid
PFO2HxAperfluoro-3,5-dioxahexanoic acid
PFOAPerfluorooctanoic acid
PFOSPerfluorooctanesulfonic acid
PFOSAPerfluorooctane sulfonamide
PFPeAPerfluoropentanoic acid
PFPeSPerfluoropentanesulfonic acid
PFSAsPerfluoroalkanesulfonic acids
PMOPeriodic mesoporous nanosilica
POPsPersistent Organic Pollutants
PTFEPolytetrafluoroethylene
RAFTReversible addition-fragmentation chain transfer
ROReverse osmosis
SAFFSurface-active foam fractionation
SBACSludge-based activated carbon
SCWOSupercritical water oxidation
SGSwitchgrass
SSSewage sludge
SSBCsSewage sludge biochar
SWCNTsSingle-wall carbon nanotubes
TGAThermogravimetric analysis
TOFTotal organic fluorine
WCBCWood chips biochar
WOWater oak leaves
WWTPWastewater treatment plants
XRDX-ray diffraction
ZIFZeolitic imidazolate framework

References

  1. Ordonez, D.; Valencia, A.; Sadmani, A.A.; Chang, N.-B. Green sorption media for the removal of perfluorooctanesulfonic acid (PFOS) and perfluorooctanoic acid (PFOA) from water. Sci. Total Environ. 2022, 819, 152886. [Google Scholar] [CrossRef]
  2. Bolan, N.; Sarkar, B.; Yan, Y.; Li, Q.; Wijesekara, H.; Kannan, K.; Tsang, D.C.; Schauerte, M.; Bosch, J.; Noll, H. Remediation of poly-and perfluoroalkyl substances (PFAS) contaminated soils–To mobilize or to immobilize or to degrade? J. Hazard. Mater. 2021, 401, 123892. [Google Scholar] [CrossRef]
  3. OECD. Toward a New Comprehensive Global Database of Per-and Polyfluoroalkyl Substances (PFAS): Summary Report on Updating the OECD 2007 List of Per-and Polyfluoroalkyl Substances (PFAS); Organisation for Economic Cooperation and Development (OECD): Paris, France, 2018. [Google Scholar]
  4. Kucharzyk, K.H.; Darlington, R.; Benotti, M.; Deeb, R.; Hawley, E. Novel treatment technologies for PFAS compounds: A critical review. J. Environ. Manag. 2017, 204, 757–764. [Google Scholar] [CrossRef]
  5. Santiago, A.R.; Medina, P.B.; Su, X. Electrochemical remediation of perfluoroalkyl substances from water. Electrochim. Acta 2022, 403, 139635. [Google Scholar] [CrossRef]
  6. ITRC. Per- and Polyfluoroalkyl Substances (PFAS). 2020. Available online: https://pfas-1.itrcweb.org/ (accessed on 25 September 2023).
  7. Liu, Y.-L.; Sun, M. Ion exchange removal and resin regeneration to treat per- and polyfluoroalkyl ether acids and other emerging PFAS in drinking water. Water Res. 2021, 207, 117781. [Google Scholar] [CrossRef]
  8. Mukhopadhyay, R.; Sarkar, B.; Palansooriya, K.N.; Dar, J.Y.; Bolan, N.S.; Parikh, S.J.; Sonne, C.; Ok, Y.S. Natural and engineered clays and clay minerals for the removal of poly-and perfluoroalkyl substances from water: State-of-the-art and future perspectives. Adv. Colloid Interface Sci. 2021, 297, 102537. [Google Scholar] [CrossRef]
  9. Hopkins, Z.R.; Sun, M.; DeWitt, J.C.; Knappe, D.R. Recently detected drinking water contaminants: GenX and other per- and polyfluoroalkyl ether acids. J.-Am. Water Work. Assoc. 2018, 110, 13–28. [Google Scholar] [CrossRef]
  10. Xu, B.; Liu, S.; Zhou, J.L.; Zheng, C.; Weifeng, J.; Chen, B.; Zhang, T.; Qiu, W. PFAS and their substitutes in groundwater: Occurrence, transformation and remediation. J. Hazard. Mater. 2021, 412, 125159. [Google Scholar] [CrossRef]
  11. Jin, T.; Peydayesh, M.; Joerss, H.; Zhou, J.; Bolisetty, S.; Mezzenga, R. Amyloid fibril-based membranes for PFAS removal from water. Environ. Sci. Water Res. Technol. 2021, 7, 1873–1884. [Google Scholar] [CrossRef]
  12. Fàbrega, F.; Kumar, V.; Benfenati, E.; Schuhmacher, M.; Domingo, J.L.; Nadal, M. Physiologically based pharmacokinetic modeling of perfluoroalkyl substances in the human body. Toxicol. Environ. Chem. 2015, 97, 814–827. [Google Scholar] [CrossRef]
  13. Müller, M.H.B.; Polder, A.; Brynildsrud, O.; Grønnestad, R.; Karimi, M.; Lie, E.; Manyilizu, W.B.; Mdegela, R.; Mokiti, F.; Murtadha, M. Prenatal exposure to persistent organic pollutants in Northern Tanzania and their distribution between breast milk, maternal blood, placenta and cord blood. Environ. Res. 2019, 170, 433–442. [Google Scholar] [CrossRef]
  14. Grandjean, P.; Timmermann, C.A.G.; Kruse, M.; Nielsen, F.; Vinholt, P.J.; Boding, L.; Heilmann, C.; Mølbak, K. Severity of COVID-19 at elevated exposure to perfluorinated alkylates. PLoS ONE 2020, 15, e0244815. [Google Scholar] [CrossRef]
  15. Calafat, A.M.; Wong, L.-Y.; Kuklenyik, Z.; Reidy, J.A.; Needham, L.L. Polyfluoroalkyl chemicals in the US population: Data from the National Health and Nutrition Examination Survey (NHANES) 2003–2004 and comparisons with NHANES 1999–2000. Environ. Health Perspect. 2007, 115, 1596–1602. [Google Scholar] [CrossRef]
  16. Hearon, S.E.; Orr, A.A.; Moyer, H.; Wang, M.; Tamamis, P.; Phillips, T.D. Montmorillonite clay-based sorbents decrease the bioavailability of per- and polyfluoroalkyl substances (PFAS) from soil and their translocation to plants. Environ. Res. 2022, 205, 112433. [Google Scholar] [CrossRef]
  17. Herzke, D.; Huber, S.; Bervoets, L.; D’Hollander, W.; Hajslova, J.; Pulkrabova, J.; Brambilla, G.; De Filippis, S.P.; Klenow, S.; Heinemeyer, G. Perfluorinated alkylated substances in vegetables collected in four European countries; occurrence and human exposure estimations. Environ. Sci. Pollut. Res. 2013, 20, 7930–7939. [Google Scholar] [CrossRef]
  18. Luo, Q.; Ren, T.; Lei, Z.; Huang, Y.; Huang, Y.; Xu, D.; Wan, C.; Guo, X.; Wu, Y. Non-toxic chitosan-based hydrogel with strong adsorption and sensitive detection abilities for tetracycline. Chem. Eng. J. 2022, 427, 131738. [Google Scholar] [CrossRef]
  19. Brusseau, M.; Guo, B. PFAS concentrations in soil versus soil porewater: Mass distributions and the impact of adsorption at air-water interfaces. Chemosphere 2022, 302, 134938. [Google Scholar] [CrossRef]
  20. Babu, D.S.; Mol, J.M.; Buijnsters, J.G. Experimental insights into anodic oxidation of hexafluoropropylene oxide dimer acid (GenX) on boron-doped diamond anodes. Chemosphere 2022, 288, 132417. [Google Scholar] [CrossRef]
  21. Kancharla, S.; Choudhary, A.; Davis, R.T.; Dong, D.; Bedrov, D.; Tsianou, M.; Alexandridis, P. GenX in Water: Interactions and Self-Assembly. J. Hazard. Mater. 2022, 428, 128137. [Google Scholar] [CrossRef]
  22. Heidari, H.; Abbas, T.; Ok, Y.S.; Tsang, D.C.; Bhatnagar, A.; Khan, E. GenX is not always a better fluorinated organic compound than PFOA: A critical review on aqueous phase treatability by adsorption and its associated cost. Water Res. 2021, 205, 117683. [Google Scholar] [CrossRef]
  23. Vakili, M.; Bao, Y.; Gholami, F.; Gholami, Z.; Deng, S.; Wang, W.; Awasthi, A.K.; Rafatullah, M.; Cagnetta, G.; Yu, G. Removal of HFPO-DA (GenX) from aqueous solutions: A mini-review. Chem. Eng. J. 2021, 424, 130266. [Google Scholar] [CrossRef]
  24. Yang, L.-H.; Yang, W.-J.; Lv, S.-H.; Zhu, T.-T.; Sharif, H.M.A.; Yang, C.; Du, J.; Lin, H. Is HFPO-DA (GenX) a suitable substitute for PFOA? A comprehensive degradation comparison of PFOA and GenX via electrooxidation. Environ. Res. 2022, 204, 111995. [Google Scholar] [CrossRef]
  25. Gebreab, K.Y.; Eeza, M.N.; Bai, T.; Zuberi, Z.; Matysik, J.; O’Shea, K.E.; Alia, A.; Berry, J.P. Comparative toxicometabolomics of perfluorooctanoic acid (PFOA) and next-generation perfluoroalkyl substances. Environ. Pollut. 2020, 265, 114928. [Google Scholar] [CrossRef]
  26. Bao, Y.; Huang, J.; Cagnetta, G.; Yu, G. Removal of F–53B as PFOS alternative in chrome plating wastewater by UV/Sulfite reduction. Water Res. 2019, 163, 114907. [Google Scholar] [CrossRef]
  27. Burkhardt, J.B.; Burns, N.; Mobley, D.; Pressman, J.G.; Magnuson, M.L.; Speth, T.F. Modeling PFAS removal using granular activated carbon for full-scale system design. J. Environ. Eng. 2022, 148, 04021086. [Google Scholar] [CrossRef]
  28. Wang, W.; Zhou, Z.; Shao, H.; Zhou, S.; Yu, G.; Deng, S. Cationic covalent organic framework for efficient removal of PFOA substitutes from aqueous solution. Chem. Eng. J. 2021, 412, 127509. [Google Scholar] [CrossRef]
  29. Zhang, J.; Pang, H.; Gray, S.; Ma, S.; Xie, Z.; Gao, L. PFAS removal from wastewater by in-situ formed ferric nanoparticles: Solid phase loading and removal efficiency. J. Environ. Chem. Eng. 2021, 9, 105452. [Google Scholar] [CrossRef]
  30. Dickman, R.A.; Aga, D.S. A Review of Recent Studies on Toxicity, Sequestration, and Degradation of Per-and Polyfluoroalkyl Substances (PFAS). J. Hazard. Mater. 2022, 436, 129120. [Google Scholar] [CrossRef]
  31. Appleman, T.D.; Dickenson, E.R.; Bellona, C.; Higgins, C.P. Nanofiltration and granular activated carbon treatment of perfluoroalkyl acids. J. Hazard. Mater. 2013, 260, 740–746. [Google Scholar] [CrossRef]
  32. Kundu, S.; Patel, S.; Halder, P.; Patel, T.; Marzbali, M.H.; Pramanik, B.K.; Paz-Ferreiro, J.; de Figueiredo, C.C.; Bergmann, D.; Surapaneni, A. Removal of PFAS from biosolids using a semi-pilot scale pyrolysis reactor and the application of biosolids derived biochar for the removal of PFAS from contaminated water. Environ. Sci. Water Res. Technol. 2021, 7, 638–649. [Google Scholar] [CrossRef]
  33. Krahn, K.M. Biochars from Contaminated Sewage Sludge as Sustainable PFAS-Sorbents for Water and Soil Clean-Up. Master’s Thesis, Norwegian University of Life Sciences, Ås, Norway, 2022. [Google Scholar]
  34. Plunkett, R.J. Discoverer of Fluoropolymers. Available online: https://web.archive.org/web/20030709071637/http://www.fluoropolymers.org/news/PlunkArt94.pdf (accessed on 25 September 2023).
  35. Prevedouros, K.; Cousins, I.T.; Buck, R.C.; Korzeniowski, S.H. Sources, fate and transport of perfluorocarboxylates. Environ. Sci. Technol. 2006, 40, 32–44. [Google Scholar] [CrossRef]
  36. Place, B.J.; Field, J.A. Identification of novel fluorochemicals in aqueous film-forming foams used by the US military. Environ. Sci. Technol. 2012, 46, 7120–7127. [Google Scholar] [CrossRef]
  37. Abunada, Z.; Alazaiza, M.Y.; Bashir, M.J. An overview of per- and polyfluoroalkyl substances (PFAS) in the environment: Source, fate, risk and regulations. Water 2020, 12, 3590. [Google Scholar] [CrossRef]
  38. Fan, X.; Tang, S.; Wang, Y.; Fan, W.; Ben, Y.; Naidu, R.; Dong, Z. Global exposure to per- and polyfluoroalkyl substances and associated burden of low birthweight. Environ. Sci. Technol. 2022, 56, 4282–4294. [Google Scholar] [CrossRef]
  39. Florentin, A.; Deblonde, T.; Diguio, N.; Hautemaniere, A.; Hartemann, P. Impacts of two perfluorinated compounds (PFOS and PFOA) on human hepatoma cells: Cytotoxicity but no genotoxicity? Int. J. Hyg. Environ. Health 2011, 214, 493–499. [Google Scholar] [CrossRef]
  40. Espartero, L.J.L.; Yamada, M.; Ford, J.; Owens, G.; Prow, T.; Juhasz, A. Health-related toxicity of emerging per- and polyfluoroalkyl substances: Comparison to legacy PFOS and PFOA. Environ. Res. 2022, 212, 113431. [Google Scholar] [CrossRef]
  41. Ahrens, L.; Bundschuh, M. Fate and effects of poly- and perfluoroalkyl substances in the aquatic environment: A review. Environ. Toxicol. Chem. 2014, 33, 1921–1929. [Google Scholar] [CrossRef]
  42. Bao, Y.; Deng, S.; Jiang, X.; Qu, Y.; He, Y.; Liu, L.; Chai, Q.; Mumtaz, M.; Huang, J.; Cagnetta, G. Degradation of PFOA substitute: GenX (HFPO–DA ammonium salt): Oxidation with UV/persulfate or reduction with UV/sulfite? Environ. Sci. Technol. 2018, 52, 11728–11734. [Google Scholar] [CrossRef]
  43. Dasu, K.; Xia, X.; Siriwardena, D.; Klupinski, T.P.; Seay, B. Concentration profiles of per- and polyfluoroalkyl substances in major sources to the environment. J. Environ. Manag. 2022, 301, 113879. [Google Scholar] [CrossRef]
  44. Hall, S.M.; Patton, S.; Petreas, M.; Zhang, S.; Phillips, A.L.; Hoffman, K.; Stapleton, H.M. Per-and polyfluoroalkyl substances in dust collected from residential homes and fire stations in North America. Environ. Sci. Technol. 2020, 54, 14558–14567. [Google Scholar] [CrossRef]
  45. Darwin, R. Estimated Quantities of Aqueous Film Forming Foam (AFFF) in the United States; Hughes Associates, Inc.: Baltimore, MD, USA, 2004. [Google Scholar]
  46. Rosenfeld, P.E.; Spaeth, K.R.; Remy, L.L.; Byers, V.; Muerth, S.A.; Hallman, R.C.; Summers-Evans, J.; Barker, S. Perfluoroalkyl substances exposure in firefighters: Sources and implications. Environ. Res. 2023, 220, 115164. [Google Scholar] [CrossRef]
  47. Fan, J.-L.; Zhang, Y.-J.; Wang, B. The impact of urbanization on residential energy consumption in China: An aggregated and disaggregated analysis. Renew. Sustain. Energy Rev. 2017, 75, 220–233. [Google Scholar] [CrossRef]
  48. Garg, S.; Kumar, P.; Mishra, V.; Guijt, R.; Singh, P.; Dumée, L.F.; Sharma, R.S. A review on the sources, occurrence and health risks of per-/poly-fluoroalkyl substances (PFAS) arising from the manufacture and disposal of electric and electronic products. J. Water Process Eng. 2020, 38, 101683. [Google Scholar] [CrossRef]
  49. Tansel, B. From electronic consumer products to e-wastes: Global outlook, waste quantities, recycling challenges. Environ. Int. 2017, 98, 35–45. [Google Scholar] [CrossRef]
  50. Hamid, H.; Li, L.Y.; Grace, J.R. Review of the fate and transformation of per- and polyfluoroalkyl substances (PFAS) in landfills. Environ. Pollut. 2018, 235, 74–84. [Google Scholar] [CrossRef]
  51. O’Connor, J.; Bolan, N.S.; Kumar, M.; Nitai, A.S.; Ahmed, M.B.; Bolan, S.S.; Vithanage, M.; Rinklebe, J.; Mukhopadhyay, R.; Srivastava, P. Distribution, transformation and remediation of poly-and per-fluoroalkyl substances (PFAS) in wastewater sources. Process Saf. Environ. Prot. 2022, 164, 91–108. [Google Scholar] [CrossRef]
  52. Göckener, B.; Fliedner, A.; Rüdel, H.; Fettig, I.; Koschorreck, J. Exploring unknown per- and polyfluoroalkyl substances in the German environment–the total oxidizable precursor assay as helpful tool in research and regulation. Sci. Total Environ. 2021, 782, 146825. [Google Scholar] [CrossRef]
  53. Filipovic, M.; Berger, U. Are perfluoroalkyl acids in waste water treatment plant effluents the result of primary emissions from the technosphere or of environmental recirculation? Chemosphere 2015, 129, 74–80. [Google Scholar] [CrossRef]
  54. Coggan, T.L.; Moodie, D.; Kolobaric, A.; Szabo, D.; Shimeta, J.; Crosbie, N.D.; Lee, E.; Fernandes, M.; Clarke, B.O. An investigation into per- and polyfluoroalkyl substances (PFAS) in nineteen Australian wastewater treatment plants (WWTPs). Heliyon 2019, 5, e02316. [Google Scholar] [CrossRef]
  55. Kotthoff, M.; Müller, J.; Jürling, H.; Schlummer, M.; Fiedler, D. Perfluoroalkyl and polyfluoroalkyl substances in consumer products. Environ. Sci. Pollut. Res. 2015, 22, 14546–14559. [Google Scholar] [CrossRef]
  56. Jansen, K. Forever Chemicals’ No More? These Technologies Aim to Destroy PFAS in Water. Available online: https://cen.acs.org/environment/persistent-pollutants/Forever-chemicals-technologies-aim-destroy/97/i12 (accessed on 25 September 2023).
  57. Amen, R.; Bashir, H.; Bibi, I.; Shaheen, S.M.; Niazi, N.K.; Shahid, M.; Hussain, M.M.; Antoniadis, V.; Shakoor, M.B.; Al-Solaimani, S.G. A critical review on arsenic removal from water using biochar-based sorbents: The significance of modification and redox reactions. Chem. Eng. J. 2020, 396, 125195. [Google Scholar] [CrossRef]
  58. Li, X.; Shen, Q.; Zhang, D.; Mei, X.; Ran, W.; Xu, Y.; Yu, G. Functional groups determine biochar properties (pH and EC) as studied by two-dimensional 13C NMR correlation spectroscopy. PLoS ONE 2013, 8, e65949. [Google Scholar] [CrossRef]
  59. Patel, S.; Marzbali, M.H.; Hakeem, I.G.; Veluswamy, G.; Rathnayake, N.; Nahar, K.; Agnihotri, S.; Bergmann, D.; Surapaneni, A.; Gupta, R. Production of H2 and CNM from biogas decomposition using biosolids-derived biochar and the application of the CNM-coated biochar for PFAS adsorption. Waste Manag. 2023, 159, 146–153. [Google Scholar] [CrossRef]
  60. Ndoun, M.C.; Elliott, H.A.; Preisendanz, H.E.; Williams, C.F.; Knopf, A.; Watson, J.E. Adsorption of pharmaceuticals from aqueous solutions using biochar derived from cotton gin waste and guayule bagasse. Biochar 2021, 3, 89–104. [Google Scholar] [CrossRef]
  61. Inyang, M.; Gao, B.; Yao, Y.; Xue, Y.; Zimmerman, A.R.; Pullammanappallil, P.; Cao, X. Removal of heavy metals from aqueous solution by biochars derived from anaerobically digested biomass. Bioresour. Technol. 2012, 110, 50–56. [Google Scholar] [CrossRef]
  62. Militao, I.M.; Roddick, F.; Fan, L.; Zepeda, L.C.; Parthasarathy, R.; Bergamasco, R. PFAS removal from water by adsorption with alginate-encapsulated plant albumin and rice straw-derived biochar. J. Water Process Eng. 2023, 53, 103616. [Google Scholar] [CrossRef]
  63. Wang, Z.; Alinezhad, A.; Nason, S.; Xiao, F.; Pignatello, J.J. Enhancement of per- and polyfluoroalkyl substances removal from water by pyrogenic carbons: Tailoring carbon surface chemistry and pore properties. Water Res. 2023, 229, 119467. [Google Scholar] [CrossRef]
  64. Wang, Z.; Alinezhad, A.; Sun, R.; Xiao, F.; Pignatello, J.J. Pre-and Postapplication Thermal Treatment Strategies for Sorption Enhancement and Reactivation of Biochars for Removal of Per-and Polyfluoroalkyl Substances from Water. ACS ES&T Eng. 2023, 3, 193–200. [Google Scholar]
  65. Krahn, K.M.; Cornelissen, G.; Castro, G.; Arp, H.P.H.; Asimakopoulos, A.G.; Wolf, R.; Holmstad, R.; Zimmerman, A.R.; Sørmo, E. Sewage sludge biochars as effective PFAS-sorbents. J. Hazard. Mater. 2023, 445, 130449. [Google Scholar] [CrossRef]
  66. Steigerwald, J.M.; Peng, S.; Ray, J.R. Novel Perfluorooctanesulfonate-Imprinted Polymer Immobilized on Spent Coffee Grounds Biochar for Selective Removal of Perfluoroalkyl Acids in Synthetic Wastewater. ACS ES&T Eng. 2023, 3, 520–532. [Google Scholar]
  67. Fabregat-Palau, J.; Vidal, M.; Rigol, A. Examining sorption of perfluoroalkyl substances (PFAS) in biochars and other carbon-rich materials. Chemosphere 2022, 302, 134733. [Google Scholar] [CrossRef]
  68. Wu, Y.; Qi, L.; Chen, G. A mechanical investigation of perfluorooctane acid adsorption by engineered biochar. J. Clean. Prod. 2022, 340, 130742. [Google Scholar] [CrossRef]
  69. Rodrigo, P.M.; Navarathna, C.; Pham, M.T.; McClain, S.J.; Stokes, S.; Zhang, X.; Perez, F.; Gunatilake, S.R.; Karunanayake, A.G.; Anderson, R. Batch and fixed bed sorption of low to moderate concentrations of aqueous per- and poly-fluoroalkyl substances (PFAS) on Douglas fir biochar and its Fe3O4 hybrids. Chemosphere 2022, 308, 136155. [Google Scholar] [CrossRef]
  70. Du, Z.; Deng, S.; Bei, Y.; Huang, Q.; Wang, B.; Huang, J.; Yu, G. Adsorption behavior and mechanism of perfluorinated compounds on various adsorbents—A review. J. Hazard. Mater. 2014, 274, 443–454. [Google Scholar] [CrossRef]
  71. Pauletto, P.S.; Bandosz, T.J. Activated carbon versus metal-organic frameworks: A review of their PFAS adsorption performance. J. Hazard. Mater. 2022, 425, 127810. [Google Scholar] [CrossRef]
  72. Kim, K.Y.; Ekpe, O.D.; Lee, H.-J.; Oh, J.-E. Perfluoroalkyl substances and pharmaceuticals removal in full-scale drinking water treatment plants. J. Hazard. Mater. 2020, 400, 123235. [Google Scholar] [CrossRef]
  73. Rodowa, A.E.; Knappe, D.R.; Chiang, S.-Y.D.; Pohlmann, D.; Varley, C.; Bodour, A.; Field, J.A. Pilot scale removal of per- and polyfluoroalkyl substances and precursors from AFFF-impacted groundwater by granular activated carbon. Environ. Sci. Water Res. Technol. 2020, 6, 1083–1094. [Google Scholar] [CrossRef]
  74. Elika, K.; Medha, K.; Sweta, M.; Jasneet, P.; Mohammad, K.; Ali, A.; Rabbani, E.M.; Amir, R. A juxtaposed review on adsorptive removal of PFAS by metal-organic frameworks (MOFs) with carbon-based materials, ion exchange resins, and polymer adsorbents. Chemosphere 2022, 311, 136933. [Google Scholar]
  75. Mohamed, B.A.; Li, L.Y.; Hamid, H.; Jeronimo, M. Sludge-based activated carbon and its application in the removal of perfluoroalkyl substances: A feasible approach towards a circular economy. Chemosphere 2022, 294, 133707. [Google Scholar] [CrossRef]
  76. Fingolo, A.C.; Klein, B.C.; Rezende, M.C.; Silva e Souza, C.A.; Yuan, J.; Yin, G.; Bonomi, A.; Martinez, D.S.; Strauss, M. Techno-economic assessment and critical properties tuning of activated carbons from pyrolyzed sugarcane bagasse. Waste Biomass Valorization 2020, 11, 1–13. [Google Scholar] [CrossRef]
  77. Gagliano, E.; Sgroi, M.; Falciglia, P.P.; Vagliasindi, F.G.; Roccaro, P. Removal of poly-and perfluoroalkyl substances (PFAS) from water by adsorption: Role of PFAS chain length, effect of organic matter and challenges in adsorbent regeneration. Water Res. 2020, 171, 115381. [Google Scholar] [CrossRef]
  78. Zhong, T.; Lin, T.; Zhang, X.; Jiang, F.; Chen, H. Impact of biological activated carbon filtration and backwashing on the behaviour of PFAS in drinking water treatment plants. J. Hazard. Mater. 2023, 446, 130641. [Google Scholar] [CrossRef]
  79. Rahman, M.F.; Peldszus, S.; Anderson, W.B. Behaviour and fate of perfluoroalkyl and polyfluoroalkyl substances (PFAS) in drinking water treatment: A review. Water Res. 2014, 50, 318–340. [Google Scholar] [CrossRef]
  80. Riegel, M.; Haist-Gulde, B.; Sacher, F. Sorptive removal of short-chain perfluoroalkyl substances (PFAS) during drinking water treatment using activated carbon and anion exchanger. Environ. Sci. Eur. 2023, 35, 12. [Google Scholar] [CrossRef]
  81. Medina, R.; Pannu, M.W.; Grieco, S.A.; Hwang, M.; Pham, C.; Plumlee, M.H. Pilot-scale comparison of granular activated carbons, ion exchange, and alternative adsorbents for per- and polyfluoroalkyl substances removal. AWWA Water Sci. 2022, 4, e1308. [Google Scholar] [CrossRef]
  82. McCleaf, P.; Englund, S.; Östlund, A.; Lindegren, K.; Wiberg, K.; Ahrens, L. Removal efficiency of multiple poly-and perfluoroalkyl substances (PFAS) in drinking water using granular activated carbon (GAC) and anion exchange (AE) column tests. Water Res. 2017, 120, 77–87. [Google Scholar] [CrossRef]
  83. Yuan, J.; Mortazavian, S.; Passeport, E.; Hofmann, R. Evaluating perfluorooctanoic acid (PFOA) and perfluorooctanesulfonic acid (PFOS) removal across granular activated carbon (GAC) filter-adsorbers in drinking water treatment plants. Sci. Total Environ. 2022, 838, 156406. [Google Scholar] [CrossRef]
  84. Li, R.; Alomari, S.; Islamoglu, T.; Farha, O.K.; Fernando, S.; Thagard, S.M.; Holsen, T.M.; Wriedt, M. Systematic study on the removal of per- and polyfluoroalkyl substances from contaminated groundwater using metal–organic frameworks. Environ. Sci. Technol. 2021, 55, 15162–15171. [Google Scholar] [CrossRef]
  85. Clark, C.A.; Heck, K.N.; Powell, C.D.; Wong, M.S. Highly defective UiO-66 materials for the adsorptive removal of perfluorooctanesulfonate. ACS Sustain. Chem. Eng. 2019, 7, 6619–6628. [Google Scholar] [CrossRef]
  86. Li, R.; Alomari, S.; Stanton, R.; Wasson, M.C.; Islamoglu, T.; Farha, O.K.; Holsen, T.M.; Thagard, S.M.; Trivedi, D.J.; Wriedt, M. Efficient removal of per- and polyfluoroalkyl substances from water with zirconium-based metal–organic frameworks. Chem. Mater. 2021, 33, 3276–3285. [Google Scholar] [CrossRef]
  87. Chen, Z.; Li, P.; Anderson, R.; Wang, X.; Zhang, X.; Robison, L.; Redfern, L.R.; Moribe, S.; Islamoglu, T.; Gómez-Gualdrón, D.A. Balancing volumetric and gravimetric uptake in highly porous materials for clean energy. Science 2020, 368, 297–303. [Google Scholar] [CrossRef]
  88. Li, P.; Vermeulen, N.A.; Malliakas, C.D.; Gómez-Gualdrón, D.A.; Howarth, A.J.; Mehdi, B.L.; Dohnalkova, A.; Browning, N.D.; O’Keeffe, M.; Farha, O.K. Bottom-up construction of a superstructure in a porous uranium-organic crystal. Science 2017, 356, 624–627. [Google Scholar] [CrossRef]
  89. Sharma, V.; Orejon, D.; Takata, Y.; Krishnan, V.; Harish, S. Gladiolus dalenii based bioinspired structured surface via soft lithography and its application in water vapor condensation and fog harvesting. ACS Sustain. Chem. Eng. 2018, 6, 6981–6993. [Google Scholar] [CrossRef]
  90. Chang, P.-H.; Mukhopadhyay, R.; Zhong, B.; Yang, Q.-Y.; Zhou, S.; Tzou, Y.-M.; Sarkar, B. Synthesis and characterization of PCN-222 metal organic framework and its application for removing perfluorooctane sulfonate from water. J. Colloid Interface Sci. 2023, 636, 459–469. [Google Scholar] [CrossRef]
  91. Pala, J.; Le, T.; Kasula, M.; Esfahani, M.R. Systematic investigation of PFOS adsorption from water by Metal Organic Frameworks, Activated Carbon, Metal Organic Framework@ Activated Carbon, and functionalized Metal Organic Frameworks. Sep. Purif. Technol. 2023, 309, 123025. [Google Scholar] [CrossRef]
  92. Kasula, M.; Le, T.; Thomsen, A.; Esfahani, M.R. Silver metal organic frameworks and copper metal organic frameworks immobilized on graphene oxide for enhanced adsorption in water treatment. Chem. Eng. J. 2022, 439, 135542. [Google Scholar] [CrossRef]
  93. VanOursouw, T.M.; Rottiger, T.; Wadzinski, K.A.; VanderWaal, B.E.; Snyder, M.J.; Bittner, R.T.; Farha, O.K.; Riha, S.C.; Mondloch, J.E. Adsorption of a PFAS Utilizing MOF-808: Development of an Undergraduate Laboratory Experiment in a Capstone Course. J. Chem. Educ. 2023, 100, 861–868. [Google Scholar] [CrossRef]
  94. Zhang, Y.; Tan, X.; Lu, R.; Tang, Y.; Qie, H.; Huang, Z.; Zhao, J.; Cui, J.; Yang, W.; Lin, A. Enhanced Removal of Polyfluoroalkyl Substances by Simple Modified Biochar: Adsorption Performance and Theoretical Calculation. ACS ES&T Water 2023, 3, 817–826. [Google Scholar]
  95. Mohamed, B.A.; Nicomel, N.R.; Hamid, H.; Li, L.Y. Using circular economy principles in the optimisation of sludge-based activated carbon production for the removal of perfluoroalkyl substances. Sci. Total Environ. 2023, 874, 162392. [Google Scholar] [CrossRef]
  96. Saeidi, N.; Kopinke, F.-D.; Georgi, A. What is specific in adsorption of perfluoroalkyl acids on carbon materials? Chemosphere 2021, 273, 128520. [Google Scholar] [CrossRef]
  97. Zhi, Y.; Liu, J. Adsorption of perfluoroalkyl acids by carbonaceous adsorbents: Effect of carbon surface chemistry. Environ. Pollut. 2015, 202, 168–176. [Google Scholar] [CrossRef]
  98. Saeidi, N.; Kopinke, F.-D.; Georgi, A. Understanding the effect of carbon surface chemistry on adsorption of perfluorinated alkyl substances. Chem. Eng. J. 2020, 381, 122689. [Google Scholar] [CrossRef]
  99. Azmi, L.H.M.; Williams, D.R.; Ladewig, B.P. Polymer-assisted modification of metal-organic framework MIL-96 (Al): Influence of HPAM concentration on particle size, crystal morphology and removal of harmful environmental pollutant PFOA. Chemosphere 2021, 262, 128072. [Google Scholar] [CrossRef]
  100. Hu, Y.; Guo, M.; Zhang, S.; Jiang, W.; Xiu, T.; Yang, S.; Kang, M.; Dongye, Z.; Li, Z.; Wang, L. Microwave synthesis of metal-organic frameworks absorbents (DUT-5-2) for the removal of PFOS and PFOA from aqueous solutions. Microporous Mesoporous Mater. 2022, 333, 111740. [Google Scholar] [CrossRef]
  101. Yang, Y.; Zheng, Z.; Ji, W.; Xu, J.; Zhang, X. Insights to perfluorooctanoic acid adsorption micro-mechanism over Fe-based metal organic frameworks: Combining computational calculation with response surface methodology. J. Hazard. Mater. 2020, 395, 122686. [Google Scholar] [CrossRef]
  102. Erkal, T.S.; Shamsuddin, N.; Kirmizialtin, S.; Yazaydin, A.O. Computational Investigation of Structure–Function Relationship in Fluorine-Functionalized MOFs for PFOA Capture from Water. J. Phys. Chem. C 2023, 127, 3204–3216. [Google Scholar] [CrossRef]
  103. Wang, W.; Maimaiti, A.; Shi, H.; Wu, R.; Wang, R.; Li, Z.; Qi, D.; Yu, G.; Deng, S. Adsorption behavior and mechanism of emerging perfluoro-2-propoxypropanoic acid (GenX) on activated carbons and resins. Chem. Eng. J. 2019, 364, 132–138. [Google Scholar] [CrossRef]
  104. Dixit, F.; Barbeau, B.; Mostafavi, S.G.; Mohseni, M. Efficient removal of GenX (HFPO-DA) and other perfluorinated ether acids from drinking and recycled waters using anion exchange resins. J. Hazard. Mater. 2020, 384, 121261. [Google Scholar] [CrossRef]
  105. Liu, C.; Chu, J.; Cápiro, N.L.; Fortner, J.D.; Pennell, K.D. In-situ sequestration of perfluoroalkyl substances using polymer-stabilized ion exchange resin. J. Hazard. Mater. 2022, 422, 126960. [Google Scholar] [CrossRef]
  106. Boyer, T.H.; Fang, Y.; Ellis, A.; Dietz, R.; Choi, Y.J.; Schaefer, C.E.; Higgins, C.P.; Strathmann, T.J. Anion exchange resin removal of per- and polyfluoroalkyl substances (PFAS) from impacted water: A critical review. Water Res. 2021, 200, 117244. [Google Scholar] [CrossRef]
  107. Chow, S.J.; Croll, H.C.; Ojeda, N.; Klamerus, J.; Capelle, R.; Oppenheimer, J.; Jacangelo, J.G.; Schwab, K.J.; Prasse, C. Comparative investigation of PFAS adsorption onto activated carbon and anion exchange resins during long-term operation of a pilot treatment plant. Water Res. 2022, 226, 119198. [Google Scholar] [CrossRef]
  108. Tan, X.; Dewapriya, P.; Prasad, P.; Chang, Y.; Huang, X.; Wang, Y.; Gong, X.; Hopkins, T.E.; Fu, C.; Thomas, K.V. Efficient Removal of Perfluorinated Chemicals from Contaminated Water Sources Using Magnetic Fluorinated Polymer Sorbents. Angew. Chem. Int. Ed. 2022, 61, e202213071. [Google Scholar] [CrossRef]
  109. Soriano, A.; Schaefer, C.; Urtiaga, A. Enhanced treatment of perfluoroalkyl acids in groundwater by membrane separation and electrochemical oxidation. Chem. Eng. J. Adv. 2020, 4, 100042. [Google Scholar] [CrossRef]
  110. Herkert, N.J.; Merrill, J.; Peters, C.; Bollinger, D.; Zhang, S.; Hoffman, K.; Ferguson, P.L.; Knappe, D.R.; Stapleton, H.M. Assessing the effectiveness of point-of-use residential drinking water filters for perfluoroalkyl substances (PFAS). Environ. Sci. Technol. Lett. 2020, 7, 178–184. [Google Scholar] [CrossRef]
  111. Pica, N.E.; Funkhouser, J.; Yin, Y.; Zhang, Z.; Ceres, D.M.; Tong, T.; Blotevogel, J. Electrochemical oxidation of hexafluoropropylene oxide dimer acid (GenX): Mechanistic insights and efficient treatment train with nanofiltration. Environ. Sci. Technol. 2019, 53, 12602–12609. [Google Scholar] [CrossRef]
  112. Li, X.; Li, A.; Li, Z.; Sun, H.; Shi, P.; Zhou, Q.; Shuang, C. Organic micropollutants and disinfection byproducts removal from drinking water using concurrent anion exchange and chlorination process. Sci. Total Environ. 2021, 752, 141470. [Google Scholar] [CrossRef]
  113. Li, M.; Sun, F.; Shang, W.; Zhang, X.; Dong, W.; Dong, Z.; Zhao, S. Removal mechanisms of perfluorinated compounds (PFCs) by nanofiltration: Roles of membrane-contaminant interactions. Chem. Eng. J. 2021, 406, 126814. [Google Scholar] [CrossRef]
  114. Soriano, A.; Gorri, D.; Urtiaga, A. Selection of high flux membrane for the effective removal of short-chain perfluorocarboxylic acids. Ind. Eng. Chem. Res. 2019, 58, 3329–3338. [Google Scholar] [CrossRef]
  115. Liu, C.J.; Strathmann, T.J.; Bellona, C. Rejection of per- and polyfluoroalkyl substances (PFAS) in aqueous film-forming foam by high-pressure membranes. Water Res. 2021, 188, 116546. [Google Scholar] [CrossRef]
  116. Guo, H.; Zhang, J.; Peng, L.E.; Li, X.; Chen, Y.; Yao, Z.; Fan, Y.; Shih, K.; Tang, C.Y. High-efficiency capture and recovery of anionic perfluoroalkyl substances from water using PVA/PDDA nanofibrous membranes with near-zero energy consumption. Environ. Sci. Technol. Lett. 2021, 8, 350–355. [Google Scholar] [CrossRef]
  117. Mantripragada, S.; Obare, S.O.; Zhang, L. Addressing Short-Chain PFAS Contamination in Water with Nanofibrous Adsorbent/Filter Material from Electrospinning. Acc. Chem. Res. 2023, 131, 235–29170. [Google Scholar] [CrossRef]
  118. Zhi, Y.; Zhao, X.; Qian, S.; Faria, A.F.; Lu, X.; Wang, X.; Li, W.; Han, L.; Tao, Z.; He, Q. Removing emerging perfluoroalkyl ether acids and fluorotelomer sulfonates from water by nanofiltration membranes: Insights into performance and underlying mechanisms. Sep. Purif. Technol. 2022, 298, 121648. [Google Scholar] [CrossRef]
  119. Wang, Y.; Zucker, I.; Boo, C.; Elimelech, M. Removal of emerging wastewater organic contaminants by polyelectrolyte multilayer nanofiltration membranes with tailored selectivity. ACS ES&T Eng. 2020, 1, 404–414. [Google Scholar]
  120. Ross, I.; McDonough, J.; Miles, J.; Storch, P.; Thelakkat Kochunarayanan, P.; Kalve, E.; Hurst, J.; Dasgupta, S.; Burdick, J. A review of emerging technologies for remediation of PFAS. Remediat. J. 2018, 28, 101–126. [Google Scholar] [CrossRef]
  121. Liu, C.; Zhao, X.; Faria, A.F.; Quiñones, K.Y.D.; Zhang, C.; He, Q.; Ma, J.; Shen, Y.; Zhi, Y. Evaluating the Efficiency of Nanofiltration and Reverse Osmosis Membrane Processes for the Removal of Per-and Polyfluoroalkyl Substances from Water: A Critical Review. Sep. Purif. Technol. 2022, 2022, 122161. [Google Scholar] [CrossRef]
  122. Buckley, T.; Karanam, K.; Xu, X.; Shukla, P.; Firouzi, M.; Rudolph, V. Effect of mono-and di-valent cations on PFAS removal from water using foam fractionation–A modelling and experimental study. Sep. Purif. Technol. 2022, 286, 120508. [Google Scholar] [CrossRef]
  123. Burns, D.J.; Stevenson, P.; Murphy, P.J. PFAS removal from groundwaters using Surface-Active Foam Fractionation. Remediat. J. 2021, 31, 19–33. [Google Scholar] [CrossRef]
  124. Stevenson, P.; Li, X. Foam Fractionation: Principles and Process Design; CRC Press: Boca Raton, FL, USA, 2014. [Google Scholar]
  125. Morrison, A.L.; Strezov, V.; Niven, R.K.; Taylor, M.P.; Wilson, S.P.; Wang, J.; Burns, D.J.; Murphy, P.J. Impact of Salinity and Temperature on Removal of PFAS Species from Water by Aeration in the Absence of Additional Surfactants: A Novel Application of Green Chemistry Using Adsorptive Bubble Fractionation. Ind. Eng. Chem. Res. 2023, 62, 5635–5645. [Google Scholar] [CrossRef]
  126. Burns, D.J.; Hinrichsen, H.M.; Stevenson, P.; Murphy, P.J. Commercial-scale remediation of per- and polyfluoroalkyl substances from a landfill leachate catchment using Surface-Active Foam Fractionation (SAFF®). Remediat. J. 2022, 32, 139–150. [Google Scholar] [CrossRef]
  127. Smith, S.J.; Lauria, M.; Ahrens, L.; McCleaf, P.; Hollman, P.; Bjälkefur Seroka, S.; Hamers, T.; Arp, H.P.H.; Wiberg, K. Electrochemical Oxidation for Treatment of PFAS in Contaminated Water and Fractionated Foam—A Pilot-Scale Study. ACS ES&T Water 2023, 3, 1201–1211. [Google Scholar]
  128. Marquínez, A.N.M.; Molina, N.S.L.; Quiroz-Fernández, L.S.; Maddela, N.R.; Luque, R.; Díaz, J.M.R. Recent advances in the remediation of perfluoroalkylated and polyfluoroalkylated contaminated sites. Environ. Res. 2022, 219, 115152. [Google Scholar] [CrossRef]
  129. Horst, J.; McDonough, J.; Ross, I.; Dickson, M.; Miles, J.; Hurst, J.; Storch, P. Water treatment technologies for PFAS: The next generation. Groundw. Monit. Remediat. 2018, 38, 13–23. [Google Scholar] [CrossRef]
  130. O’Connor, W.; Partridge, G.; Fielder, S.; Woolley, L.; Palanisami, T. Evaluation of Practical Technologies for Per-and Polyfluoroalkyl Substance (PFAS) Remediation in Marine Fish Hatcheries. 2020. Available online: https://www.frdc.com.au/project/2018-125 (accessed on 25 September 2023).
  131. Li, J.; Pinkard, B.R.; Wang, S.; Novosselov, I.V. Hydrothermal treatment of per- and polyfluoroalkyl substances (PFAS). Chemosphere 2022, 307, 135888. [Google Scholar] [CrossRef]
  132. Hao, S.; Choi, Y.J.; Deeb, R.A.; Strathmann, T.J.; Higgins, C.P. Application of hydrothermal alkaline treatment for destruction of per- and polyfluoroalkyl substances in contaminated groundwater and soil. Environ. Sci. Technol. 2022, 56, 6647–6657. [Google Scholar] [CrossRef]
  133. Pinkard, B.R.; Shetty, S.; Stritzinger, D.; Bellona, C.; Novosselov, I.V. Destruction of perfluorooctanesulfonate (PFOS) in a batch supercritical water oxidation reactor. Chemosphere 2021, 279, 130834. [Google Scholar] [CrossRef]
  134. Pinkard, B.R.; Austin, C.; Purohit, A.L.; Li, J.; Novosselov, I.V. Destruction of PFAS in AFFF-impacted fire training pit water, with a continuous hydrothermal alkaline treatment reactor. Chemosphere 2023, 314, 137681. [Google Scholar] [CrossRef]
  135. Yang, N.; Yang, S.; Ma, Q.; Beltran, C.; Guan, Y.; Morsey, M.; Brown, E.; Fernando, S.; Holsen, T.M.; Zhang, W. Solvent-Free Nonthermal Destruction of PFAS Chemicals and PFAS in Sediment by Piezoelectric Ball Milling. Environ. Sci. Technol. Lett. 2023, 10, 198–203. [Google Scholar] [CrossRef]
  136. Austin, C.; Li, J.; Moore, S.; Purohit, A.; Pinkard, B.R.; Novosselov, I.V. Destruction and defluorination of PFAS matrix in continuous-flow supercritical water oxidation reactor: Effect of operating temperature. Chemosphere 2023, 327, 138358. [Google Scholar] [CrossRef]
  137. Li, J.; Austin, C.; Moore, S.; Pinkard, B.R.; Novosselov, I.V. PFOS destruction in a continuous supercritical water oxidation reactor. Chem. Eng. J. 2023, 451, 139063. [Google Scholar] [CrossRef]
  138. Van Hoomissen, D.J.; Vyas, S. Early events in the reductive dehalogenation of linear perfluoroalkyl substances. Environ. Sci. Technol. Lett. 2019, 6, 365–371. [Google Scholar] [CrossRef]
  139. Tsai, Y.-T.; Yu-Chen Lin, A.; Weng, Y.-H.; Li, K.-C. Treatment of perfluorinated chemicals by electro-microfiltration. Environ. Sci. Technol. 2010, 44, 7914–7920. [Google Scholar] [CrossRef]
  140. Ji, Y.; Choi, Y.J.; Fang, Y.; Pham, H.S.; Nou, A.T.; Lee, L.S.; Niu, J.; Warsinger, D.M. Electric Field-Assisted Nanofiltration for PFOA Removal with Exceptional Flux, Selectivity, and Destruction. Environ. Sci. Technol. 2023. [Google Scholar] [CrossRef]
  141. Hayashi, R.; Obo, H.; Takeuchi, N.; Yasuoka, K. Decomposition of perfluorinated compounds in water by DC plasma within oxygen bubbles. Electr. Eng. Jpn. 2015, 190, 9–16. [Google Scholar] [CrossRef]
  142. Lashuk, B.; Pineda, M.; AbuBakr, S.; Boffito, D.; Yargeau, V. Application of photocatalytic ozonation with a WO3/TiO2 catalyst for PFAS removal under UVA/visible light. Sci. Total Environ. 2022, 843, 157006. [Google Scholar] [CrossRef]
  143. Amor, C.; Marchão, L.; Lucas, M.S.; Peres, J.A. Application of advanced oxidation processes for the treatment of recalcitrant agro-industrial wastewater: A review. Water 2019, 11, 205. [Google Scholar] [CrossRef]
  144. Bao, Y.; Cagnetta, G.; Huang, J.; Yu, G. Degradation of hexafluoropropylene oxide oligomer acids as PFOA alternatives in simulated nanofiltration concentrate: Effect of molecular structure. Chem. Eng. J. 2020, 382, 122866. [Google Scholar] [CrossRef]
  145. Masud, A.; Guardian, M.G.E.; Travis, S.C.; Soria, N.G.C.; Jarin, M.; Aga, D.S.; Aich, N. Redox-active rGO-nZVI nanohybrid-catalyzed chain shortening of perfluorooctanoic acid (PFOA) and perfluorooctane sulfonic acid (PFOS). J. Hazard. Mater. Lett. 2021, 2, 100007. [Google Scholar] [CrossRef]
  146. Zhang, W.; Cao, H.; Liang, Y. Degradation by hydrothermal liquefaction of fluoroalkylether compounds accumulated in cattails (Typha latifolia). J. Environ. Chem. Eng. 2021, 9, 105363. [Google Scholar] [CrossRef]
  147. Wu, B.; Hao, S.; Choi, Y.; Higgins, C.P.; Deeb, R.; Strathmann, T.J. Rapid destruction and defluorination of perfluorooctanesulfonate by alkaline hydrothermal reaction. Environ. Sci. Technol. Lett. 2019, 6, 630–636. [Google Scholar] [CrossRef]
  148. Krause, M.J.; Thoma, E.; Sahle-Damesessie, E.; Crone, B.; Whitehill, A.; Shields, E.; Gullett, B. Supercritical water oxidation as an innovative technology for PFAS destruction. J. Environ. Eng. 2022, 148, 05021006. [Google Scholar] [CrossRef]
  149. Zhu, Y.; Ji, H.; He, K.; Blaney, L.; Xu, T.; Zhao, D. Photocatalytic degradation of GenX in water using a new adsorptive photocatalyst. Water Res. 2022, 220, 118650. [Google Scholar] [CrossRef]
  150. Smaili, H.; Ng, C. Adsorption as a Remediation Technology for Short-Chain Per-and Polyfluoroalkyl Substances (PFAS) from Water-A Critical Review. Environ. Sci. Water Res. Technol. 2023, 9, 344–362. [Google Scholar] [CrossRef]
  151. Ma, H.; Kang, Y.; Li, M.; Dong, J.; Wang, Y.; Xiao, J.; Guo, Z. Enhancement of perfluorooctanoic acid and perfluorooctane sulfonic acid removal in constructed wetland using iron mineral: Performance and mechanisms. J. Hazard. Mater. 2023, 447, 130819. [Google Scholar] [CrossRef]
  152. Shi, H.-T.; Feng, X.-C.; Xiao, Z.-J.; Wang, W.-Q.; Wang, Y.-M.; Zhang, X.; Xu, Y.-J.; Ren, N.-Q. Analysis of the β-cyclodextrin enhancing bio-denitrification from the perspective of substrate metabolism, electron transfer, and iron acquisition. Chem. Eng. J. 2022, 446, 137358. [Google Scholar] [CrossRef]
  153. Douna, B.K.; Yousefi, H. Removal of PFAS by Biological Methods. Asian Pac. J. Environ. Cancer 2023, 6, 53–68. [Google Scholar] [CrossRef]
  154. Wanninayake, D.M. Comparison of currently available PFAS remediation technologies in water: A review. J. Environ. Manag. 2021, 283, 111977. [Google Scholar] [CrossRef]
  155. Deng, S.; Zhang, Q.; Nie, Y.; Wei, H.; Wang, B.; Huang, J.; Yu, G.; Xing, B. Sorption mechanisms of perfluorinated compounds on carbon nanotubes. Environ. Pollut. 2012, 168, 138–144. [Google Scholar] [CrossRef]
  156. Yin, S.; Villagrán, D. Design of nanomaterials for the removal of per- and poly-fluoroalkyl substances (PFAS) in water: Strategies, mechanism, challenges, and opportunities. Sci. Total Environ. 2022, 831, 154939. [Google Scholar] [CrossRef]
  157. Bei, Y.; Deng, S.; Du, Z.; Wang, B.; Huang, J.; Yu, G. Adsorption of perfluorooctane sulfonate on carbon nanotubes: Influence of pH and competitive ions. Water Sci. Technol. 2014, 69, 1489–1495. [Google Scholar] [CrossRef]
  158. Park, M.; Wu, S.; Lopez, I.J.; Chang, J.Y.; Karanfil, T.; Snyder, S.A. Adsorption of perfluoroalkyl substances (PFAS) in groundwater by granular activated carbons: Roles of hydrophobicity of PFAS and carbon characteristics. Water Res. 2020, 170, 115364. [Google Scholar] [CrossRef]
  159. Fang, C.; Sobhani, Z.; Megharaj, M.; Naidu, R. Electrochemical Proof of Fluorophilic Interaction among Fluoro-Carbon Chains. Electroanalysis 2018, 30, 2349–2355. [Google Scholar] [CrossRef]
  160. Min, X.; Huo, J.; Dong, Q.; Xu, S.; Wang, Y. Enhanced sorption of perfluorooctanoic acid with organically functionalized layered double hydroxide. Chem. Eng. J. 2022, 446, 137019. [Google Scholar] [CrossRef]
  161. Min, X.; Wang, Y. Enhanced adsorption of short-chain perfluorobutanoic acid by functionalized periodic mesoporous organosilica: Performance and mechanisms. J. Hazard. Mater. 2023, 449, 131047. [Google Scholar] [CrossRef]
  162. Feng, D.; Song, C.; Mo, W. Environmental, human health, and economic implications of landfill leachate treatment for per- and polyfluoroalkyl substance removal. J. Environ. Manag. 2021, 289, 112558. [Google Scholar] [CrossRef]
  163. Murray, C.C.; Marshall, R.E.; Liu, C.J.; Vatankhah, H.; Bellona, C.L. PFAS treatment with granular activated carbon and ion exchange resin: Comparing chain length, empty bed contact time, and cost. J. Water Process Eng. 2021, 44, 102342. [Google Scholar] [CrossRef]
  164. León, M.; Silva, J.; Carrasco, S.; Barrientos, N. Design, cost estimation and sensitivity analysis for a production process of activated carbon from waste nutshells by physical activation. Processes 2020, 8, 945. [Google Scholar] [CrossRef]
  165. Stavropoulos, G.; Zabaniotou, A. Minimizing activated carbons production cost. Fuel Process. Technol. 2009, 90, 952–957. [Google Scholar] [CrossRef]
  166. Saha, D.; Khan, S.; Van Bramer, S.E. Can porous carbons be a remedy for PFAS pollution in water? A perspective. J. Environ. Chem. Eng. 2021, 9, 106665. [Google Scholar] [CrossRef]
  167. Stoiber, T.; Evans, S.; Temkin, A.M.; Andrews, D.Q.; Naidenko, O.V. PFAS in drinking water: An emergent water quality threat. Water Solut. 2020, 1, e49. [Google Scholar]
  168. Inyang, M.; Dickenson, E. The potential role of biochar in the removal of organic and microbial contaminants from potable and reuse water: A review. Chemosphere 2015, 134, 232–240. [Google Scholar] [CrossRef]
  169. Calafat, A.M.; Kato, K.; Hubbard, K.; Jia, T.; Botelho, J.C.; Wong, L.-Y. Legacy and alternative per- and polyfluoroalkyl substances in the US general population: Paired serum-urine data from the 2013–2014 National Health and Nutrition Examination Survey. Environ. Int. 2019, 131, 105048. [Google Scholar] [CrossRef]
  170. Garg, S.; Wang, J.; Kumar, P.; Mishra, V.; Arafat, H.; Sharma, R.S.; Dumée, L.F. Remediation of water from per-/poly-fluoroalkyl substances (PFAS)–Challenges and perspectives. J. Environ. Chem. Eng. 2021, 9, 105784. [Google Scholar] [CrossRef]
  171. Grandjean, P.; Clapp, R. Perfluorinated alkyl substances: Emerging insights into health risks. New Solut. A J. Environ. Occup. Health Policy 2015, 25, 147–163. [Google Scholar] [CrossRef]
  172. Kanchanapiya, P.; Tantisattayakul, T. Analysis of the additional cost of addressing per- and polyfluoroalkyl substance contamination from landfill leachate by reverse osmosis membranes in Thailand. J. Water Process Eng. 2022, 45, 102520. [Google Scholar] [CrossRef]
  173. Speth, T.; Khera, R.; Patterson, C.; Ransom, P. Cost of POU vs Centralized Treatment. In Proceedings of the WQA Virtual Annual, Virtual Meeting, OH, USA, 16–21 April 2020. [Google Scholar]
Figure 1. Timeline of the PFAS transformation with the passage of time from 1938 to 2024.
Figure 1. Timeline of the PFAS transformation with the passage of time from 1938 to 2024.
Sustainability 15 16173 g001
Figure 2. Sources of PFAS.
Figure 2. Sources of PFAS.
Sustainability 15 16173 g002
Figure 3. The isotherm models for PFOA adsorption on different types of biochar at varied temperatures (323.15 K, 303.15 K, and 313.15 K) (reproduced with permission from the publisher; Wu, Qi [68]).
Figure 3. The isotherm models for PFOA adsorption on different types of biochar at varied temperatures (323.15 K, 303.15 K, and 313.15 K) (reproduced with permission from the publisher; Wu, Qi [68]).
Sustainability 15 16173 g003
Figure 4. Synthetic route for featured fluorinated polymer sorbents with magnetic properties (reproduced with permission from publisher; Tan et al. [108]).
Figure 4. Synthetic route for featured fluorinated polymer sorbents with magnetic properties (reproduced with permission from publisher; Tan et al. [108]).
Sustainability 15 16173 g004
Figure 5. Identified NF/RO membranes used to remove a wide range of PFAS from water. “*” in the figure represent the data was established from the Web of Science (reproduced with permission from publisher; Liu et al. [121]).
Figure 5. Identified NF/RO membranes used to remove a wide range of PFAS from water. “*” in the figure represent the data was established from the Web of Science (reproduced with permission from publisher; Liu et al. [121]).
Sustainability 15 16173 g005
Figure 6. The schematic diagram exhibiting foam fractionation and electrochemical oxidation.
Figure 6. The schematic diagram exhibiting foam fractionation and electrochemical oxidation.
Sustainability 15 16173 g006
Figure 7. The intermediates in the oxidative degradation of HFPO-DA (reproduced with permission from publisher; Babu et al. [20]).
Figure 7. The intermediates in the oxidative degradation of HFPO-DA (reproduced with permission from publisher; Babu et al. [20]).
Sustainability 15 16173 g007
Figure 8. Mechanisms for the removal of PFAs via adsorption.
Figure 8. Mechanisms for the removal of PFAs via adsorption.
Sustainability 15 16173 g008
Figure 9. Mechanism of PFAS removal in iron mineral-based wetlands (Reproduced with permission from publisher; Ma et al. [151].
Figure 9. Mechanism of PFAS removal in iron mineral-based wetlands (Reproduced with permission from publisher; Ma et al. [151].
Sustainability 15 16173 g009
Table 1. Physiochemical characteristics of PFAS reported in this study.
Table 1. Physiochemical characteristics of PFAS reported in this study.
TypeAbbreviationMolecular FormulaIUPAC NameMolecular WeightStructure
PFECAsPer- or polyfluoroalkyl ether carboxylic acid
PFMOPrAPerfluoro-2-methoxypropanoic acidC4HF7O3Perfluoro-2-(perfluoromethoxy)propanoic acid, 2,3,3,3-Tetrafluoro-2-(trifluoromethoxy)propanoic acid230.04Sustainability 15 16173 i001
PFO2HxAPerfluoro(3,5-dioxahexanoic) acidC4HF7O4Perfluoro-3,5-dioxahexanoic acid246.04Sustainability 15 16173 i002
PFMOBAPerfluoro-4-methoxybutanoic acidC5HF9O3Perfluoro-4-methoxybutanoic acid, 2,2,3,3,4,4-Hexafluoro-4-(trifluoromethoxy)butanoic acid280.04Sustainability 15 16173 i003
HFPO-DAHexafluoropropylene oxide dimer acidC6HF11O3Perfluoro-2-methyl-3-oxahexanoic acid330.05Sustainability 15 16173 i004
PFSAsPerfluorosulfonic Acid
PFBSperfluorobutane sulfonateC4HF9O3S1,1,2,2,3,3,4,4,4-Nonafluorobutane-1-sulfonic acid, Nonafluorobutanesulfonic acid300.1Sustainability 15 16173 i005
PFPeSPerfluoropentanesulfonic acidC5HF11O3Sperfluoropentanesulfonic acid, 1,1,2,2,3,3,4,4,5,5,5-undecafluoropentane-1-sulfonic acid350.11Sustainability 15 16173 i006
PFHxSPerfluorohexanesulfonic acidC6HF13O3SPerfluorohexane-1-sulphonic acid, 1,1,2,2,3,3,4,4,5,5,6,6,6-Tridecafluorohexane-1-sulfonic acid400.12Sustainability 15 16173 i007
PFHpSPerfluoroheptanesulfonic acidC7HF15O3Sperfluoroheptane sulfonic acid, 1,1,2,2,3,3,4,4,5,5,6,6,7,7,7-pentadecafluoroheptane-1-sulfonic acid450.12Sustainability 15 16173 i008
PFOSPerfluorooctanesulfonic acidC8HF17O3SHeptadecafluorooctanesulfonic acid, 1,1,2,2,3,3,4,4,5,5,6,6,7,7,8,8,8-Heptadecafluorooctane-1-sulfonic acid.500.13Sustainability 15 16173 i009
PFNSPerfluorononanesulfonic acidC9HF19O3SNonadecafluoro-1-nonanesulfonic acid, 1-Nonanesulfonic acid, 1,1,2,2,3,3,4,4,5,5,6,6,7,7,8,8,9,9,9-nonadecafluoro-550.14Sustainability 15 16173 i010
PFDSPerfluorodecanesulfonic acidC10HF21O3SHenicosafluorodecanesulphonic acid, 1,1,2,2,3,3,4,4,5,5,6,6,7,7,8,8,9,9,10,10,10-henicosafluorodecane-1-sulfonic acid600.15Sustainability 15 16173 i011
PFCAsPerfluoroalkyl carboxylic acids
PFBAPerfluorobutanoic acidC3F7COOHHeptafluorobutanoic acid214.04Sustainability 15 16173 i012
PFPeAPerfluoropentanoic acidC4F9COOHPerfluorovaleric acid, nonafluoropentanoic acid264.05Sustainability 15 16173 i013
PFHxAPerfluorohexanoic acidC5F11COOHUndecafluorohexanoic acid314.05Sustainability 15 16173 i014
PFHpAPerfluoroheptanoic acidC6F13COOHTridecafluoroheptanoic acid364.06Sustainability 15 16173 i015
PFOAPerfluorooctanoic acidC8HF15O2Pentadecafluorooctanoic acid, 2,2,3,3,4,4,5,5,6,6,7,7,8,8,8-Pentadecafluorooctanoic acid414.07Sustainability 15 16173 i016
PFNAPerfluorononanoic acidC9F17O2heptadecafluorononanoic acid, Perfluoro-n-nonanoic acid464.08Sustainability 15 16173 i017
PFDAPerfluorodecanoic acidC9F19COOHNonadecafluorodecanoic acid, Perfluoro-N-decanoic acid514.08Sustainability 15 16173 i018
Table 2. Removal of PFAS via adsorption on biochar, metal-organic frameworks, and activated carbon.
Table 2. Removal of PFAS via adsorption on biochar, metal-organic frameworks, and activated carbon.
Removal Technique or AdsorbentPFAS TypepHPZCDosageSurface Area (m2/g)Initial
Conc.
Adsorption Capacity (mg g−1)Regeneration MethodRemoval
Rate (%)
Ref
Carbon nanomaterial coated biochar (CNM-Biochar)PFOS--1 g40 ± 2.1310 µg/L--95[59]
PFOA----25 µg/L--71
PFHxS----140 µg/L--72
Biochar alginate compositePFOS84.370.5–1.5 g/L4.7100 μg L−11572 µg g−1-99[62]
PFBS--1.0–3.0 g/L--2895 µg g−1-40
SW600-PPAO (Soft wood-post pyrolysis air oxidation)PFBA7.4-400 mg8320.0687–4.63 μM PFAS1.75--[64]
PFPeA-----1.75--
GenX-----1.45--
PFDA-----4.76--
PFBS-----2.6--
PFHxS-----3.79--
Acid-modified biochar (H-SL)PFOA7-1 g3.62650–100 mg/L49.02--[94]
PFOS-----79.8--
Sewage sludge-activated carbon (ZnCl2-impregnated)PFBA--5 mg72350 μg/L--87[95]
PFPeA-------95
PFHxA-------96
PFHpA-------97
PFOA-------100
PFNA-------100
PFBS-------99
PFHxS-------98
PFOS-------100
Classical powder-activated carbonPFOS4-0.004 g1232.382 ppm28.24850:50 ethanol and water mixture93[91]
Thermal defunctionalized CACF (DeCACF)PFOA79.3-10008 (µg/L)80methanol extraction-[96]
PFOS 103 -
PFBA 5.1 -
Commercial PAC AquaNuchar (AquaNC)PFOA74.7510 mg18600.5 mg2.46--[97]
PFOS 2.48 -
Commercial-activated carbon felt ACTITEX WK L20 (CACF)PFOS75.9 1100-33.3--[96]
PFOA 2.08 -
Commercial-activated carbon felt ACTITEX FC 1501 (FC15)PFOS77.1-1600-156--[98]
PFOA 35.7 -
Commercial-activated carbon felt ACTITEX FC 1001 (FC10)PFOS76.5-1400-108--
PFOA 10.2
MIL-101 (Cr)@AC MOFPFOS4-100 ppm69552 ppm25.70650:50 ethanol and water mixture80[91]
PCN-222 MOFPFOS4.1–5.37.50.003 g1948500 mg/L2257--[90]
UiO-66PFOS56.40.5 g/L1423500 mg/L1.24 mmol--[85]
MIL-96-Al modified with hydrolyzed polyacrylamidePFOA--1.0 g/L751000 mg/L340methanol/water-[99]
DUT-5-2PFOA3.0-5 mg184030 mg/L98.2methanol60.9[100]
PFOS-----145.4--
MIL-100-FePFOA3.33.6450–150 mg123750–1000 mg/L426.6--[101]
Table 3. Removal of PFAS via nanofiltration and reverse osmosis.
Table 3. Removal of PFAS via nanofiltration and reverse osmosis.
MembraneWater FluxFeed SolutionpHRemoval EfficiencyDuration (h)MechanismReferences
BW30-RO PFOA @ 10 µg/L-92%6-[109]
RO GenX @ < 0.01 >99% -[110]
NF GenX @ 1 mg/L 99.50%6-[111]
NF90 PFOS @ 100 ppb 96.40%--[112]
PFOA @ 100 ppb 97.40%--
NFG1.4 × 10−5 m/sPFOS @ 100 ppb7.427.90%24-[112]
PFOA @ 100 ppb 55.60%24-
BW30 PFOS @ 5.0 × 102–1.6 × 1064>99.1--[113]
BW30 PFHxA @ 1.0 × 1053.5–7.196.0 to >99.9--[114]
ESPA-2540 PFPeA, PFHxA, PFHpA, PFOA, PFPrS, PFBS, PFPeS, PFHxS, PFHpS, PFOS @ 0.3–58.47.0 ± 0.4>99.0 (ALL)--[115]
ESPA-2540 PFPeA, PFHxA, PFHpA, PFOA, PFPrS, PFBS, PFPeS, PFHxS, PFHpS, PFOS)7.0 ± 0.4>99.0 (ALL)--
XLE PFHxA @ 1.0 × 1053.5–7.1>97.5--[114]
SW30XLE PFHxA @ 1.0 × 1053.5–7.1>96.0--
NF90 PFOA @ 10 µg/L 72%5-[109]
GenX @ 200 μg/L-97%-NaCl/methanol solution[116]
Porous amine-functionalized membrane PFOS @ ≤ 1.2 × 106 μg/m2 >99%--
PFOA @ ≤ 1.2 × 106 μg/m2 >99%--
ASFPAN nanofibrous membrane GenX @ 100 mg/L435%-superhydrophilicity and the Coulomb force[117]
ESPAN nanofibrous membrane 88%-hydrophobic interaction and dipole–dipole interaction
Five commercial membranes DK > NF90 > XN45 > NF270 > DL≤0.1 L⋅m−2·h−1PFCA @ 100 µg L−15.5 ± 0.0667.6% to 95.8%-Hydrophobic interactions[118]
79.2% to >99.9%-
PFSA @ 100 µg L−1 79.2% to >99.9%-
FTS @ 100 µg L−1 69.0% to 97.2%-
PFEA @ 100 µg L−1 66.0% to >99.9%-
PEMe PFOS @ 1.0 × 103 μg⋅L−1765.0–90.0- [119]
PFOA @ 1.0 × 103 μg⋅L−135.0–90.0-
Table 4. Efficiencies of different destructive technologies for PFAS removal.
Table 4. Efficiencies of different destructive technologies for PFAS removal.
Removal MethodPFAS TypeInitial ConcentrationAdditivesRemoval Efficiency (%)Reaction DurationDefluorination EfficiencyMechanismReferences
Nanoscale Zero-Valent Iron (nZVI) on 2D reduced graphene oxide (rGO) nanosheetsPFOA200 μg/L-3910 min--[145]
PFOS -85
HALTGenX20 μg/L-1002 h--[146]
F–53B20 μg/L-12.9 ± 3.42 h-
ADONA20 μg/L-1002 h-
HALTAFFF3200 mg/LNaOH99.952 h--[134]
HALTAFFF-NaOH−1001.5 h98-[132]
HALTPFOS50 mg/LNaOH−1001.5 h>80-[147]
HALTAFFF (PFAS mixture)38,714.5 μg/Lair99.3710 s--[148]
Boron-doped diamond anodesGenX15 mg/L-57%1 h>80%Oxidative degradation, direct electron transfer[20]
PFOA38%
Electric field-assisted nanofiltrationPFOA0.5 mg/LNa2SO497.5%48 h90%1-electromigration of PFOA away from the membrane, 44 2-electro-absorption of PFOA on the electrodes, 45 and 3-membrane fouling enhancement[140]
Photocatalytic degradationGenX100 µg/L-99%1 h--[149]
piezoelectric-material-assisted ball millingPFOS-->80%2 h100%piezoelectric oxidation pathway[135]
>100
PFOA
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Amen, R.; Ibrahim, A.; Shafqat, W.; Hassan, E.B. A Critical Review on PFAS Removal from Water: Removal Mechanism and Future Challenges. Sustainability 2023, 15, 16173. https://doi.org/10.3390/su152316173

AMA Style

Amen R, Ibrahim A, Shafqat W, Hassan EB. A Critical Review on PFAS Removal from Water: Removal Mechanism and Future Challenges. Sustainability. 2023; 15(23):16173. https://doi.org/10.3390/su152316173

Chicago/Turabian Style

Amen, Rabia, Alhassan Ibrahim, Waqar Shafqat, and El Barbary Hassan. 2023. "A Critical Review on PFAS Removal from Water: Removal Mechanism and Future Challenges" Sustainability 15, no. 23: 16173. https://doi.org/10.3390/su152316173

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop