ACS Publications. Most Trusted. Most Cited. Most Read
My Activity
CONTENT TYPES

Figure 1Loading Img
RETURN TO ISSUEPREVTreatment and Resour...Treatment and Resource RecoveryNEXT

Accelerated Degradation of Perfluorosulfonates and Perfluorocarboxylates by UV/Sulfite + Iodide: Reaction Mechanisms and System Efficiencies

  • Zekun Liu
    Zekun Liu
    Department of Chemical & Environmental Engineering, University of California, Riverside, California 92521, United States
    More by Zekun Liu
  • Zhanghao Chen
    Zhanghao Chen
    State Key Laboratory of Pollution Control and Resource Reuse, School of the Environment, Nanjing University, Nanjing 210023, Jiangsu, China
    More by Zhanghao Chen
  • Jinyu Gao
    Jinyu Gao
    Department of Chemical & Environmental Engineering, University of California, Riverside, California 92521, United States
    More by Jinyu Gao
  • Yaochun Yu
    Yaochun Yu
    Department of Chemical & Environmental Engineering, University of California, Riverside, California 92521, United States
    Department of Civil & Environmental Engineering, University of Illinois at Urbana-Champaign, Urbana, Illinois 61801, United States
    More by Yaochun Yu
  • Yujie Men
    Yujie Men
    Department of Chemical & Environmental Engineering, University of California, Riverside, California 92521, United States
    Department of Civil & Environmental Engineering, University of Illinois at Urbana-Champaign, Urbana, Illinois 61801, United States
    More by Yujie Men
  • Cheng Gu
    Cheng Gu
    State Key Laboratory of Pollution Control and Resource Reuse, School of the Environment, Nanjing University, Nanjing 210023, Jiangsu, China
    More by Cheng Gu
  • , and 
  • Jinyong Liu*
    Jinyong Liu
    Department of Chemical & Environmental Engineering, University of California, Riverside, California 92521, United States
    *Email: [email protected], [email protected]
    More by Jinyong Liu
Cite this: Environ. Sci. Technol. 2022, 56, 6, 3699–3709
Publication Date (Web):February 28, 2022
https://doi.org/10.1021/acs.est.1c07608

Copyright © 2022 American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0.
  • Open Access

Article Views

11657

Altmetric

-

Citations

24
LEARN ABOUT THESE METRICS
PDF (4 MB)
Supporting Info (1)»

Abstract

The addition of iodide (I) in the UV/sulfite system (UV/S) significantly accelerated the reductive degradation of perfluorosulfonates (PFSAs, CnF2n+1SO3) and perfluorocarboxylates (PFCAs, CnF2n+1COO). Using the highly recalcitrant perfluorobutane sulfonate (C4F9SO3) as a probe, we optimized the UV/sulfite + iodide system (UV/S + I) to degrade n = 1–7 PFCAs and n = 4, 6, 8 PFSAs. In general, the kinetics of per- and polyfluoroalkyl substance (PFAS) decay, defluorination, and transformation product formations in UV/S + I were up to three times faster than those in UV/S. Both systems achieve a similar maximum defluorination. The enhanced reaction rates and optimized photoreactor settings lowered the EE/O for PFCA degradation below 1.5 kW h m–3. The relatively high quantum yield of eaq from I made the availability of hydrated electrons (eaq) in UV/S + I and UV/I two times greater than that in UV/S. Meanwhile, the rapid scavenging of reactive iodine species by SO32– made the lifetime of eaq in UV/S + I eight times longer than that in UV/I. The addition of I also substantially enhanced SO32– utilization in treating concentrated PFAS. The optimized UV/S + I system achieved >99.7% removal of most PFSAs and PFCAs and >90% overall defluorination in a synthetic solution of concentrated PFAS mixtures and NaCl. We extended the discussion over molecular transformation mechanisms, development of PFAS degradation technologies, and the fate of iodine species.

This publication is licensed under

CC-BY-NC-ND 4.0.
  • cc licence
  • by licence
  • nc licence
  • nd licence

Synopsis

Adding iodide in the UV/sulfite system significantly enhanced the reaction rate, energy efficiency, and chemical utilization for reductive defluorination of various per- and polyfluoroalkyl substance pollutants.

Introduction

ARTICLE SECTIONS
Jump To

Per- and polyfluoroalkyl substances (PFASs) have become global pollutants (1) due to their (i) wide applications since the 1940s, (ii) high recalcitrance in natural environments, and (iii) diverse toxicities to animals (2,3) and health concerns for humans. (4−7) While carbon sorption, (8,9) ion exchange, (9−11) and membrane filtration (12,13) can capture PFAS from polluted water, the subsequent treatment of the concentrated PFAS remains challenging. (14−16)
Among the rapidly developing technologies for PFAS degradation, (17−23) photochemical treatment with UV-generated electron (eaq) has demonstrated great promise for environmental remediation. (24−34) The C–F bond can be reductively converted into C–H by eaq (eq 1), (25,29) which can be generated from UV irradiation of SO32– (eq 2) with a quantum yield (QY) at 0.12–0.14 mol/einstein (35,36)
(1)
(2)
Although the 254 nm irradiation used in common water treatment applications is not the maximum absorption wavelength of SO32–, (36) the cost-effectiveness of PFAS degradation by UV/sulfite (UV/S) has been greatly enhanced by tuning reaction conditions, such as pH adjustment. (37) At pH 12.0, the defluorination has reached 73–93% for n = 2–8 perfluorocarboxylates (CnF2n+1COO, PFCAs) and 79–84% for n = 4, 6, 8 perfluorosulfonates (CnF2n+1SO3, PFSAs). (38) The C–H bonds in the UV/S treatment residue allow for further defluorination by oxidation with HO, resulting in near-complete defluorination. (38) However, PFSAs are intrinsically more recalcitrant than PFCAs for UV/S treatment. The short-chain n = 4 perfluorobutane sulfonate (PFBS, C4F9SO3) is even more challenging than longer-chain n = 6, 8 PFSAs. (25,29) Multiple spikes of sulfite and >24 h are required to degrade all parent PFBS, (38) resulting in high chemical dosage and energy consumption for an effective treatment.
Before the application of UV/S for pollutant degradation, (26,36) UV/iodide (UV/I) had been developed earlier for PFAS degradation. (24,25,39) The generation of eaq from iodide (eq 3) has a QY at 0.22–0.29 mol/einstein (40,41)
(3)
A previous study reported >95% defluorination of perfluorooctanoate (C7F15COO, PFOA, an extensively studied “representative” PFAS) by UV/I at pH 9.2 and ambient temperature. (42) Although a higher concentration of eaq could be expected from UV/I than UV/S, our use of UV/I only achieved PFOA defluorination of up to 60% (see Table A2 of the Supporting Information of ref (29)). The low efficiency could be attributed to eaq scavenging by reactive iodine species (RIS, e.g., I2, I2•–, and I3) and dissolved oxygen (DO). (25,39) While DO removal by N2 was necessary for UV/I, N2 purge is not necessary for UV/S (29) because the initial DO can be rapidly scavenged by sulfite. (43) Therefore, we added sulfite to simplify the DO removal for UV/I. To our surprise, the degradation and defluorination of the highly recalcitrant PFBS were significantly accelerated even at a 10-fold higher concentration than our previous studies (i.e., 250 vs 25 μM). To date, the UV/sulfite + iodide (UV/S + I) system has only been examined for reductive degradation of monochloroacetate (44) and bromate (45) in 2018. The first report on dechlorination by UV/S + I (44) has revealed key mechanistic insights to understand the enhanced degradation of highly recalcitrant PFAS. Herein, we systematically optimized the UV/S + I system using the PFBS as the challenging probe, elucidated the reaction mechanisms using various model fluorinated compounds, and showcased the significantly enhanced system efficiencies in both chemical dosing and energy consumption for the degradation of legacy PFSAs and PFCAs.

Materials and Methods

ARTICLE SECTIONS
Jump To

Chemicals

Sodium sulfite (Na2SO3), potassium iodide (KI), sodium hydroxide (NaOH), and sodium bicarbonate (NaHCO3) were purchased from Fisher Chemical. PFSAs (n = 4, 6, 8 CnF2n+1SO3, PFSAs) and PFCAs (n = 1–8 CnF2n+1COO) were purchased in bulk quantities (i.e., 0.1–5 g) and used as received. The information on CAS numbers, purities, and vendors are listed in Supporting Information.

Photochemical Reaction

The optimization of UV/S + I system parameters used the previously reported photochemical reactor configuration (Ace Glass parts #7864-10, #7874-38, and #7506-14, wrapped with aluminum foil). (29,46) A 600 mL solution containing 25 μM PFBS, 5 mM NaHCO3, and predetermined concentrations of KI and Na2SO3 and pH (adjusted with NaOH) was irradiated by an 18 W low-pressure mercury lamp (GPH212T5L/HO) in the closed photoreactor with jacketed water cooling at 20 °C. For the reduction of energy consumption, the optimized solution condition was further applied in a new reactor configuration, where a 2000 mL solution was irradiated by a 10 W low-pressure mercury lamp (GPH212T5L, with the same geometry as the 18 W lamp) in a quartz sleeve placed in the center of a 2 L tall-form beaker (wrapped with aluminum foil and covered by a plastic cap with parafilm sealing). Because the 2 L beaker did not have a water cooling jacket, the solution was gradually heated from 20 to 36 °C in the first 4 h by the UV lamp and then maintained around 36 °C due to the balance with heat dissipation.

Water Sample Analysis

The parent PFAS and transformation products (TPs) were analyzed by liquid chromatography–high-resolution quadrupole Orbitrap mass spectrometer (LC–HRMS/MS). Both suspect and non-target screening were conducted to identify TPs. Short-chain (n = 1 and 2) PFCAs and their TPs were analyzed by ion chromatography (IC). The degradation kinetics were fit with C/C0 > 0.2 data using the pseudo-first-order kinetic model. Rationales for this fitting method are described in Text S1. The released fluoride ion (F) was measured by an ion-selective electrode (ISE) and validated by IC. (29,38) Detailed operation parameters of LC–HRMS/MS, IC, and ISE have been reported previously (29,32,37,38) and also included in the Supporting Information. The defluorination percentage (deF%) is defined as the molar ratio between the released F in solution and the total F in the parent PFAS compounds.

Laser Flash Photolysis

A transient absorption spectrometer (LP980, Edinburgh Instruments) equipped with 266 nm laser pulse (Nd:YAG) (34) was used to measure the yield and lifetime of eaq produced from UV/S, UV/I, and UV/S + I. The sample solution (2.5 mL) containing KI and Na2SO3 at predetermined dosage was filled in a 1 cm × 1 cm × 3 cm quartz cell and then purged with N2 for 30 min. The solution was then excited by the laser beam along with the 1 cm optical path. The absorption by eaq and I3 was measured at 700 (34) and 350 nm, (47) respectively. The lifetime of eaq and I3 was calculated by eq 4
(4)
where Rt is the signal response at time t, B is the amplitude of the absorption curves at time zero, τ represents the time required for the response to decay to 37% of the original signal strength, and the fitted τ is defined as the lifetime. (34)

Results and Discussion

ARTICLE SECTIONS
Jump To

Significantly Accelerated PFBS Degradation by UV/S + I

Previously, we have used PFOA as the probe to optimize the UV/S system (pH 12.0 and 10 mM Na2SO3). (37) However, when the target pollutant was switched to PFBS, the optimized UV/S only achieved 48% decay of the parent PFBS (Figure 1a) and 42% overall defluorination after 24 h (Figure 1b). (38) To enhance the degradation, two more spikes of 10 mM Na2SO3 at 8 and 16 h (i.e., 30 mM Na2SO3 in total) boosted the parent PFBS decay and overall defluorination to 86 and 69%, respectively. (38) In sharp contrast, the addition of 2 mM KI to 10 mM Na2SO3 in the UV/S system substantially accelerated the reaction by reaching >99% decay of the parent PFBS and 78% overall defluorination after 24 h, without additional Na2SO3 spikes during the reaction. Thus, I not only accelerated the reaction but also significantly enhanced the utilization efficiency of SO32–. Notably, for most samples from different reaction settings and sampling times, the deF% from the reacted portion of the PFBS was consistent (i.e., 83 ± 2%). Therefore, the decay of the parent PFBS is the rate-limiting step and the following defluorination from TPs is fast. (29)

Figure 1

Figure 1. (a) Parent compound decay and (b) defluorination of PFBS by UV/S (10 mM), UV/I (2 mM), and UV/S (10 mM) + I (2 mM); PFBS degradation by UV/S + I at varied (c,d) iodide concentrations, (e) sulfite concentrations, and (f) solution pH. Default reaction conditions: PFBS (0.025 mM), carbonate (5 mM), pH 12.0, 254 nm irradiation (an 18 W low-pressure Hg lamp in 600 mL solution), and 20 °C.

We continued using the highly recalcitrant PFBS to further optimize the UV/S + I system. The examination on varying concentrations of I, SO32–, and pH (Figure 1c–f) confirmed that the initially used 2 mM I, 10 mM SO32–, and pH 12.0 constitute the optimal reaction conditions for 0.025 mM PFBS. The use of ≤5 mM SO32– led to early termination of defluorination (Figure 1e) probably due to the early depletion of sulfite. Further elevated concentrations of either SO32– (e.g., 20 mM) or I (e.g., 5 mM) did not result in a better performance. In our previous study on the UV/S system, 10 mM SO32– was found to be necessary to achieve the maximum defluorination of 0.025 mM PFOA. (37)

Energy Efficiency of the UV/S + I System

We further confirmed the high energy efficiency of UV/S + I from the rapid degradation of n = 6, 8 PFSAs and n = 1–7 PFCAs (Table 1). The first-order rate constants for the parent compound decay were 2.4–4.1-fold of those by UV/S, where all conditions were the same except for the absence of iodide. The term EE/O, which is defined as the electrical energy consumed for lowering the pollutant concentration by one order-of-magnitude (i.e., C/C0 = 0.1 or 90% removal), has been frequently used for water technology evaluation (48)
where P is the power of the UV lamp (kW), k is the first-order rate constant (h–1), and V is the volume of water (m3). To probe the possibility of further lowering the EE/O by UV/S + I, we decreased the power of the UV lamp from 18 to 10 W and increased the volume of water from 600 to 2000 mL. The decay of the parent PFAS in the new configuration of 10 W: 2000 mL was 20–50% slower than in the previous configuration of 18 W: 600 mL. Hence, the EE/O values were reduced by 3–5-fold from modifying the reactor configuration (see Text S2 for an extended discussion). Together with the acceleration by iodide, the EE/O for n = 2–7 PFCAs by UV/S + I have become <1.5 kW h/m3 (Table 1). Interestingly, the maximum deF% of PFCAs by UV/S in the 10 W: 2000 mL configuration was significantly lower than that in the 18 W: 600 mL configuration (Figure S1). In stark comparison, defluorination of PFCAs by UV/S + I in the 10 W: 2000 mL configuration only showed a lower initial rate, but the maximum deF% after 4 h was the same as in the 18 W: 600 mL configuration.
Table 1. DeF%, Rate Constant, EE/O, and EE/Max.deF of PFCAs and PFSAs by UV/S and UV/S + I Treatmenta
  defluorination after 1 h (%) parent compound decay k (h–1)b EE/O (kW h m–3)b EE/Max.deF (kW h m–3)b
PFCA(CnF2n+1COO) UV/S UV/S + I UV/S UV/S + I UV/S UV/S + I UV/S UV/S + I
n = 1 TFA 84 ± 1.2 100 ± 1.3 3.7 (2.4) 8.6 (6.8) 18.6 (4.8) 7.8 (1.7) <60 (10) <22.8 (3.8)
n = 2 PFPrA 47 ± 0.4 75 ± 0.9 6.7 (4.9) 17.2 (8.9) 10.2 (2.3) 4.2 (1.3) <240 (40) <60 (20)
n = 3 PFBA 52 ± 2.1 77 ± 0.8 5.8 (3.5) 19.2 (9.9) 12.0 (3.3) 3.6 (1.2) <240 (40) <120 (20)
n = 4 PFPeA 42 ± 3.9 73 ± 1.5 5.5 (3.4) 16.9 (8.7) 12.6 (3.4) 4.2 (1.3) <240 (40) <120 (20)
n = 5 PFHxA 51 ± 3.1 79 ± 1.3 6.1 (4.1) 19.7 (10.1) 11.4 (2.8) 3.6 (1.1) <240 (40) <60 (20)
n = 6 PFHpA 41 ± 2.7 77 ± 3.6 5.4 (3.7) 17.5 (9.4) 12.6 (3.1) 4.2 (1.2) <240 (60) <60 (20)
n = 7 PFOA 39 ± 1.5 88 ± 0.3 5.2 (3.1) 15.4 (8.1) 13.2 (3.7) 4.2 (1.4) <240 (60) <60 (20)
  defluorination after 24 h (%) parent compound decay k (h–1) EE/O (kW h m–3) EE/Max.deF (kW h m–3)
PFSA (CnF2n+1SO3) UV/S UV/S + I UV/S UV/S + I UV/S UV/S + I UV/S UV/S + I
n = 4 PFBS 43 ± 1.8 78 ± 1.2 0.05 (0.02) 0.19 (0.05) 1320 (576) 366 (230) ∼1320 (576) ∼366 (230)
n = 6 PFHxS 60 ± 3.1 84 ± 1.3 0.38 (0.29) 1.56 (0.61) 181 (39.7) 44.4 (18.9) <360 (60) <120 (40)
n = 8 PFOS 84 ± 0.9 92 ± 0.7 0.78 (0.65) 1.90 (1.01) 88.8 (17.7) 36.6 (11.5) <240 (40) <60 (40)
a

Reaction conditions were the following: individual PFCA/PFSA (0.025 mM), Na2SO3 (10 mM), KI (2 mM), NaHCO3 (5 mM), pH 12.0, 254 nm irradiation (an 18 W low-pressure Hg lamp in 600 mL solution), and 20 °C.

b

Values in parentheses are from an “energy-saving” setting (a 10 W lamp in 2 L solution, solution constituents not changed) at 20–36 °C due to the gradual heating by the UV lamp without a cooling water bath.

For the much more recalcitrant PFSAs, the EE/O were also lowered to 11.5–230 kW h/m3. As elucidated in our previous study, (29) one primary mechanism for parent PFCA decay is hydrodefluorination of the α −CF2– moiety, and one primary mechanism for parent PFSA decay is the reductive cleavage of the C–S bond. Because both mechanisms require the reaction with eaq, the significantly accelerated degradation of both PFCAs and PFSAs by UV/S + I can be attributed to the enhanced availability of eaq.
As the goal of PFAS degradation is not merely removing 90% of the parent compound, we propose a new metric that calculates the electrical energy consumed by the time (t) upon reaching 90% of the maximum defluorination (EE/Max.deF, Table 1)
Because defluorination involves multiple pathways and cannot be quantified with a simple kinetic model, the EE/Max.deF values are estimated as a range based on the time when the first sample showing >90% of the maximum defluorination was collected. The maximum defluorination was determined in the sample taken at 24 h when sulfite had been fully depleted (see Figure 6c).
For PFCAs, the parent compound decay was much faster than the stepwise defluorination (Figure 2a vs b,c). The decay and defluorination of n ≥ 2 PFCAs followed similar time profiles, so the average values are shown in those figures. Still, their EE/Max.deF values by UV/S + I were <20 kW h m–3. For PFSA degradation, the parent compound decay and defluorination were much more synchronous (Figure 2d–f) than PFCA degradation. Because >90% parent compound decay of PFSAs is the pre-requisite for reaching 90% of the maximum defluorination, the EE/Max.deF values for PFSAs were also higher than their EE/O. The current reactor configuration and solution composition of UV/S + I have resulted in the low-end EE/O values for PFOA/PFOS degradation among existing reports on various technologies. (16,19,23,49) Currently, the comparison of EE/Max.deF with other treatment methods is not feasible due to the lack of reported defluorination-time profiles, especially for PFBS degradation.

Figure 2

Figure 2. (a) Parent compound decay and (b) defluorination of n = 1–7 PFCAs by UV/S (10 mM), and UV/S (10 mM) + I (2 mM). Panel (c) shows the magnified defluorination profiles for n = 2–7 PFCAs (averaged) within the first 1 h. (d–f) Parent compound decay and defluorination of n = 4, 6, 8 PFSAs. Reaction conditions: individual PFAS (0.025 mM), NaHCO3 (5 mM), pH 12.0, 254 nm irradiation (an 18 W low-pressure Hg lamp in 600 mL solution), and 20 °C.

Mechanistic Insights from Reaction Kinetics and TPs

To understand how the new UV/S + I system accelerated both parent compound decay and defluorination, we used n = 1 and 2 PFCAs (CnF2n+1COO) as mechanistic probes. Short-chain PFCAs have a limited number of fluorinated carbons, and many possible TPs have commercial standards for quantitation. The decay and defluorination of the three n = 1 structures, CF3–COO (trifluoroacetate, TFA), CHF2–COO (difluoroacetate, DFA), and CH2F–COO (monofluoroacetate, MFA) by UV/S + I were significantly faster than those by UV/S (Figure 3a–c vs d–f). While the complete defluorination by UV/S took 2–4 h, only 1 h was needed for UV/S + I. The formation of acetate (CH3–COO) indicates the hydrodefluorination pathway, where F atoms were stepwise replaced by H atoms via reduction with eaq (Figure 3g). The generation (and degradation) of the stepwise hydrodefluorination TPs (e.g., DFA and MFA from TFA) was also faster in the UV/S + I system. Meanwhile, the various gaps of the acetate yield from 100% indicate the competing decarboxylation pathway (Figure 3g). At the end of reactions where 100% defluorination was achieved, the acetate yields from DFA and MFA by UV/S + I (72 and 93%, respectively) were higher than those by UV/S (45 and 72%, respectively). The preferred hydrodefluorination indicates enhanced availability of eaq in the UV/S + I system.

Figure 3

Figure 3. Parent compound decay, TP formation, and defluorination of tri-, di-, and monofluoro acetates by (a–c) UV/S + I and (d–f) UV/S and (g) reaction schemes. Reaction conditions: individual PFAS (0.025 mM), Na2SO3 (10 mM), KI (2 mM), NaHCO3 (5 mM), pH 12.0, 254 nm irradiation (an 18 W low-pressure Hg lamp in 600 mL solution), and 20 °C. The vertical red dotted lines in panels (d–f) indicate the 1 h time window where 100% defluorination by UV/S + I was achieved.

However, for the degradation of TFA, the acetate yield by UV/S + I (34%) was lower than that by UV/S (52%). The calculated C–F bond dissociation energy (BDE) in TFA (116.8 kcal mol–1) is significantly higher than those in DFA (109.7 kcal mol–1) and MFA (108.6 kcal mol–1). (29) We have also found that the C–F bonds in CF3CH2–COO (121.5 kcal mol–1) and HCF2CH2–COO (118.8 kcal mol–1) were both sluggish in the reaction with eaq. (37,50) Therefore, the dominating pathway for TFA degradation by UV/S + I should be decarboxylation.
The degradation of CF3CF2–COO (PFPrA, see Table 1 for acronyms) using UV/S + I was also significantly faster than that by UV/S (Figure 4a,b vs c,d). The recalcitrant hydrodefluorination TP, CF3CH2–COO, has very strong sp3 C–F bonds (Figure 4e) and thus is accumulated in the solution. Both UV/S + I and UV/S achieved a similar deF% of PFPrA (79% in Figure 4a and 76% in 4c), suggesting that UV/S + I cannot cleave additional C–F bonds that are recalcitrant under UV/S treatment. However, upon reaching the maximum deF%, the yield of CF3CH2–COO from UV/S + I treatment was higher than that from UV/S treatment (62 vs 39 from 250 μM of PFPrA), indicating enhanced availability of eaq for hydrodefluorination of the relatively weak C–F bonds. Notably, the limit of deF% via hydrodefluorination is 40% (i.e., two α C–F bonds out of the five C–F bonds in PFPrA are cleaved), whereas the overall deF% reached 79%. The deeper defluorination is thus contributed by the decarboxylation pathway, which mineralized the α CF2 moiety and yielded TFA (Figure 4e). TFA can be 100% defluorinated by either pathway (Figure 3).

Figure 4

Figure 4. Parent compound decay, TP formation, and defluorination of PFPrA by (a,b) UV/S + I and (c,d) UV/S. Panels (b,d) show magnified time profiles within the first hour of reaction. Panel (e) shows the proposed reaction schemes. Reaction conditions: PFPrA (0.025 mM), Na2SO3 (10 mM), KI (2 mM), NaHCO3 (5 mM), pH 12.0, 254 nm irradiation (an 18 W low-pressure Hg lamp in 600 mL solution), and 20 °C.

It is worth noting that other degradation pathways also occurred. As shown by the F balance counting F and known TPs from hydrodefluorination (CF3CFH–COO and CF3CH2–COO) and from decarboxylation (TFA, DFA, and MFA were negligible), the maximum gap reached >50% at the beginning (15 min for UV/S + I and 1 h for UV/S). After the maximum deF% was achieved (8 h for UV/S + I and 24 h for UV/S), the missed F balance was decreased to 6% for UV/S + I and 15% for UV/S. While we are investigating novel TPs with other mechanistic hypotheses, we noted that most F-containing TPs did not escape the system as gaseous products because the following oxidation with HO (generated from heat-activated persulfate at pH > 12) achieved ∼100% defluorination. (38)
The abovementioned mechanistic insights were further corroborated by the results from C7F15–COO (PFOA). The generation and the following degradation of the decarboxylation product, C6F13–COO (PFHpA), and the hydrodefluorination product, C7F14H–COO, were much faster by UV/S + I than by UV/S (Figure 5a vs b). We note that the detected concentration of a TP is the result of the balance between generation and degradation at the same time. Although the maximum concentrations of those two TPs under UV/S + I and UV/S were similar, their total generated amounts under UV/S + I should be more than those under UV/S.

Figure 5

Figure 5. Parent compound decay and TP formation for the degradation of (a,b) PFOA and (c,d) PFBS by UV/S + I and UV/S. Panel (e) shows the proposed reaction schemes. Reaction conditions were the following: individual PFAS (0.025 mM), Na2SO3 (10 mM), KI (2 mM), NaHCO3 (5 mM), pH 12.0, 254 nm irradiation (an 18 W low-pressure Hg lamp in 600 mL solution), and 20 °C.

TP analysis for the degradation of C4F9–SO3 (PFBS) also confirmed that the formation of the hydrodefluorination product, C4F7H2–SO3, was significantly enhanced by UV/S + I (Figure 5c vs d). According to the calculated C–F BDEs for PFSAs, (29) the most probable location for the initial hydrodefluorination is on the β carbon (Figure 5e). Further hydrodefluorination only yielded C4F6H3–SO3, which accumulated under UV/S + I treatment. Hence, the high overall deF% of 79% should also be attributed to C–S cleavage upon the reaction between C4F9–SO3 and eaq to yield n ≤ 3 PFCAs (Figure 5e). However, PFCA TPs (e.g., PFBA) were not detected, probably due to the rate-limiting decay of the PFBS and the rapid degradation of daughter PFCAs.

Mechanistic Insights from Reactive Species Analyses

The significantly accelerated parent compound decay and the increased yield of hydrodefluorination products by UV/S + I prompted us to explain the enhanced availability of eaq. We compared the abundance and lifetime of eaq produced from UV/S + I, UV/S, and UV/I by measuring the absorption at 700 nm in laser flash photolysis (LFP) experiments (Figure 6a). As expected from the reported QY, the abundance of eaq generated by the 266 nm laser from 2 mM I (absorbance 0.012, QY = 0.22–0.29 mol/einstein) (40,41) was approximately two times higher than that from 10 mM SO32– (absorbance 0.016, QY = 0.12–0.14 mol/einstein) (35,36) at pH 12. However, the eaq from UV/I has a 22-fold shorter lifetime than that from UV/S. Besides the quenching by H+ (eq 5), a series of RIS can rapidly form (eqs 68) and become strong scavengers of eaq (eqs 911). (25,39) This may be one reason for the limited performance of UV/I for PFAS degradation. (39,44,51,52)
(5)
(6)
(7)
(8)
(9)
(10)
(11)

Figure 6

Figure 6. (a) Kinetic profiles of transient absorption at 700 nm after the 266 nm laser pulse excitation of I, S, and S + I solutions; (b) kinetic profiles of transient absorption at 350 nm after 266 nm laser pulse excitation of I and S + I solutions; (c) decay of sulfite in UV/S and UV/S + I in the absence of PFAS; and (d) lifetime of eaq in UV/S + I at varied solution pH. Conditions: Na2SO3 (S, 10 mM), KI (I, 2 mM), and pH 12.0.

For comparison, UV/S + I generated eaq in a similar abundance to UV/I. Meanwhile, the lifetime of eaq in UV/S + I was eightfold longer than that in UV/I (Figure 6a). Besides serving as the source of eaq (eq 2), SO32– has also been found to rapidly reduce RIS back to I (eqs 1215) (53,54)
(12)
(13)
(14)
(15)
As I3 has a characteristic absorption at 352 nm, (39,44) the rapid quenching of I3 by SO32– was also confirmed by the LFP experiment (Figure 6b). The use of 10 mM SO32– shortened the lifetime of I3 by 23-fold. Although a radical reaction network remains challenging to build, a simplified interpretation is that SO32– donates up to two electrons to convert RIS back to I via eqs 1215. This interpretation is consistent with (i) the faster consumption of SO32– in UV/S + I than in UV/S (Figure 6c), (ii) the generation of eaq in high abundance and long lifetime in UV/S + I, and (iii) the enhanced reductive PFAS degradation. These experimental findings complement the previous study on monochloroacetate degradation, which conducted meticulous kinetic measurements for a comprehensive mechanistic understanding of the UV/S + I system. (44) Moreover, SO32– is an instant scavenger of DO without UV irradiation. (43) Upon adding Na2SO3 in water, the measured DO immediately dropped from 7.3 to <0.1 mg/L. Thus, the use of SO32– effectively prevented eaq quenching by DO, (25,26) and the UV/S + I system does not need N2 purge.
In addition, the lifetime of eaq at pH from 9 to 12 showed an increasing trend (Figure 6d). Because the speciation of both I and SO32– is not impacted at pH > 10, the longer lifetime at higher pH is primarily attributed to the reduced eaq quenching by H+ (eq 5). This result is consistent with the excellent PFBS degradation kinetics at pH 12 (Figure 1f).

Iodide Significantly Enhanced the Utilization of Sulfite

In our previous studies on the UV/S system, the typical concentration of PFOA was 25 μM because higher concentrations resulted in inferior performance. For example, using 10 mM SO32– at pH 12, the defluorination of 25 μM of individual n = 2–7 PFCAs reached 73–93% at 8 h. The kinetics did not show a significant dependence on the fluoroalkyl chain length (n). However, the deF% from 250 μM (or 0.25 mM) of individual PFCAs were significantly lowered and exhibited an obvious decrease with the increased n (Figure 7a). For PFOA, 0.25 mM of C7F15–COO contained 3.75 mM of the C–F bond. If 100% defluorination could be achieved solely by the hydrodefluorination mechanism (eq 1), up to 7.5 mM of eaq would be consumed. However, the use of 10 mM SO32– in UV/S only achieved 34% defluorination at 24 h, where most SO32– had been depleted (Figure 6c). Under UV irradiation, the major portion of SO32– that was not utilized for PFAS degradation probably generated H2 gas. (35) However, this aspect warrants further investigation. Additional spikes of 10 mM SO32– enabled further defluorination, but a total of 40 mM SO32– only defluorinated PFOA up to 73% (Figure 7b). In stark contrast, the initially added 10 mM SO32– and 2 mM I directly achieved 84% defluorination from the 0.25 mM PFOA within 12 h. Thus, the utilization of sulfite was significantly enhanced by the addition of iodide. Further increasing the PFOA concentration to 0.5 mM led to significantly lower defluorination (Figure 7c). Hence, we increased the concentration of all PFCAs and PFSAs from 25 to 250 μM. The UV/S + I treatment of PFCAs achieved a similar deF% for both concentrations (Figure 7d). The UV/S + I treatment of n = 4, 6, 8 PFSAs (0.25 mM) without spiking additional SO32– also achieved 68, 81, and 87% of defluorination, respectively (Figure 7e). These values are consistent with the maximum deF% of n = 4, 6, 8 PFSAs (25 μM) by UV/S with additional SO32– spikes (69, 72, and 84%, respectively). (38)

Figure 7

Figure 7. (a) DeF% of PFCAs by UV/S after 8 h; (b) time profiles for the defluorination from 0.25 mM PFOA by UV/S (with SO32– spikes every 24 h) and UV/S + I (no SO32– spike); (c) time profiles for the defluorination from various [PFOA]0 by UV/S + I; (d) deF% of PFCAs after 24 h by UV/S + I; (e) deF% of PFSAs by UV/I and UV/S + I after 24 h; (f) time profiles for the defluorination from the mixed PFAS (see Table 2) in NaCl brine after a 1:1 dilution with deionized (DI) water. Default reaction conditions: 10 mM Na2SO3 (10 mM), KI (2 mM for UV/S + I), NaHCO3 (5 mM), pH 12.0, and 254 nm irradiation (an 18 W low-pressure Hg lamp for 600 mL solution), and 20 °C.

Motivated by these positive results, we further challenged the UV/S + I system to treat a synthetic brine, which was prepared based on the NaCl content and quantified PFAS constituents in a recently reported “still bottom” waste brine from ion-exchange resin regeneration (Table 2). (16) Natural organic matters and other potential inhibiting species that require pretreatment efforts were not included in the present study. Individual PFAS component concentrations were chosen as the averaged values of the six reported brines. (16) Commercially unavailable structures (e.g., n = 5, 7, 9 PFSAs), minor structures (e.g., n = 9 PFCA and n = 8 sulfonamide), and unknown PFAS precursors were not included. Hence, the synthetic brine contained 3 wt% NaCl and a total of 10.9 mM of C–F bonds in the mixture of three fluorotelomer sulfonates (FTSAs, n = 4, 6, 8 CnF2n+1–CH2CH2–SO3), six PFCAs (n = 3–8), and three PFSAs (n = 4, 6, 8). We diluted the brine for twofold, so that the eaq generated from 10 mM SO32– and 2 mM I would suffice the need for cleaving all C–F bonds (5.5 mM after dilution) via the reductive pathway (eq 1). The UV/S + I treatment achieved 90 and 93% defluorination from this brine sample after 12 and 24 h, respectively (Figure 7f). The UV/S treatment using 50 mM SO32– exhibited an inferior rate and extent of defluorination. Quantitation of the individual PFAS in the treated brine found that a majority of parent structures have been removed for >99.5%. The outstanding substrates exhibiting high recalcitrance included PFBS (94% removal) and 4:2 FTS (52% removal). Hence, short-chain fluorotelomers are not ideal for direct UV/S + I treatment. The isolation of the very short fluoroalkyl chain by −CH2– moieties voids the favorable mechanistic factors for PFAS degradation: (1) vulnerable C–F bonds in the middle of long fluoroalkyl chains, (2) weak α C–F bonds in PFCAs, and (3) the dissociation of C–S bonds in PFSAs. (29)
Table 2. Treatment of a Concentrated PFAS Mixture in NaCl Brine by UV/S + Ia
PFAS category compound PFAS concentration (μg L–1) total fluorine in C–F bonds (μg L–1) residual PFAS after 24 h (μg L–1)b parent PFAS removal ratio (%)
FTSAs 4:2 FTS 131 68 63 51.9
  6:2 FTS 27,950 16,124 296 98.9
  8:2 FTS 875 535 <53 >93.9
PFCAs PFBA 1950 1212 <5 >99.7
  PFPeA 5872 3803 <13 >99.7
  PFHxA 16,433 10,936 <8 >99.9
  PFHpA 7492 5083 <9 >99.8
  PFOA 101,550 69,896 <10 >99.9
  PFNA 228 159 <12 >94.7
PFSAs PFBS 4988 2522 279 94.4
  PFHxS 72,550 40,893 <88 >99.8
  PFOS 86,950 56,155 <100 >99.8
sum   327.0 mg L–1 207.4 mg L–1 (10.9 mM) <0.9 mg L–1 >99.7
a

Reaction conditions: the 3% NaCl brine with the tabulated components was diluted with 1:1 DI water, Na2SO3 (10 mM), KI (2 mM), NaHCO3 (5 mM), pH 12.0, 254 nm irradiation (an 18 W low-pressure Hg lamp in 600 mL solution), and 20 °C.

b

Concentrations without a definitive value indicate the limit of quantitation by the HRMS instrument in our study. These values are the HRMS-detected concentration multiplied by 2 (due to the 1:1 dilution before treatment). Solid-phase extraction for accurate quantitation in low concentration ranges (e.g., the prevalent 70 ng L–1 reference for drinking water) was not used in this proof-of-concept study.

Implications to PFAS Degradation Research and Engineering

Our experimental results collectively suggest that the use of iodide in UV/S + I significantly benefits PFAS degradation by (i) increasing the concentration of eaq (Figure 6a) and (ii) increasing the utilization of sulfite (Figure 7b). Mechanistic probing experiments (Figures 35) identified similar transformation pathways at much higher rates by UV/S + I than by UV/S. While all reaction steps were significantly accelerated upon the addition of I, the maximum deF% remained the same. The consumptions of both electricity (Figure 2 and Table 1) and chemical (Figure 7) by UV/S + I are significantly lower than those by UV/S. Although the results have established a highly efficient photochemical system and provided mechanistic insights to the best of our current research capability, this single report cannot cover all details from fundamental understanding to practical application. For example, the interactions between various sulfur and iodine species are much more complex than the proposed, as shown in eqs 515. (39,44,55) However, a comprehensive and accurate reaction network modeling can be established upon deeper understandings of reactions for individual reactive species with PFASs and their TPs. Further elucidation of PFAS degradation pathways and mechanisms should be prioritized for future research efforts. After all, we have provided detailed data sets that confirm UV/S + I as a potential engineering solution for PFAS destruction. Although iodinated gaseous products were reported in earlier studies using UV/I, (25) the accumulation of polar iodinated organics from UV/S + I treatment is less likely because (i) C–I bonds are highly vulnerable to eaq reduction and UV photolysis (44) and (ii) RIS are rapidly scavenged by SO32–.
For PFAS degradation technology development, we highlight the importance of investigating a broad range of structures rather than the legacy PFOA/PFOS. We have shown that the reactivity of PFAS highly depends on the chain length and functional groups. (29,32,50) This study provides an example of using the highly challenging C4 PFBS as the probe for technology development. As expected, the optimized UV/S + I system exhibits significantly enhanced performance for all PFCAs and C6 and C8 PFSAs. Still, like UV/S, the UV/S + I system is mechanistically limited for treating short-chain fluorotelomers. (38) Thus, oxidation is required both before and after the UV/S + I treatment to ensure efficient and complete defluorination of most PFAS pollutants. (38) NOM and other potential inhibiting species in real “still bottom” brines will also be addressed in our near-future study on sequential treatment strategies.
Lastly, iodide is a relatively expensive chemical enriched from seawater, but we do not consider the cost of iodide as a major barrier for practical applications of UV/S + I. The volume of waste brine from ion-exchange resin regeneration is minimal. The real cost of both the electricity and chemicals for PFAS destruction in the waste brine is to be divided by the large volumes of drinking water or groundwater treated by ion exchange. However, when iodide recycling is necessary, iodine species in the treated brine can be rapidly transformed into either I (e.g., eqs 1215) or I2 (e.g., by H2O2) (56) for facile separation. I has a much higher affinity to ordinary anion exchange resins than common anions (e.g., F, Cl, SO42–, and HCO3/CO32–), (57) and I2 can be extracted by common organic solvents from water. For example, at 25 °C the solubility of I2 in diethyl ether (25 wt%) is substantially higher than that in water (∼300 mg/L). (58)

Supporting Information

ARTICLE SECTIONS
Jump To

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.est.1c07608.

  • Details of PFAS chemicals; quantitation of PFAS parent compounds, TPs, and anions; extended discussions on kinetic data fitting and photoreactor configurations; and defluorination profiles in the two photoreactor configurations (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

ARTICLE SECTIONS
Jump To

  • Corresponding Author
  • Authors
    • Zekun Liu - Department of Chemical & Environmental Engineering, University of California, Riverside, California 92521, United States
    • Zhanghao Chen - State Key Laboratory of Pollution Control and Resource Reuse, School of the Environment, Nanjing University, Nanjing 210023, Jiangsu, China
    • Jinyu Gao - Department of Chemical & Environmental Engineering, University of California, Riverside, California 92521, United StatesOrcidhttps://orcid.org/0000-0002-1751-3430
    • Yaochun Yu - Department of Chemical & Environmental Engineering, University of California, Riverside, California 92521, United StatesDepartment of Civil & Environmental Engineering, University of Illinois at Urbana-Champaign, Urbana, Illinois 61801, United StatesOrcidhttps://orcid.org/0000-0001-9231-6026
    • Yujie Men - Department of Chemical & Environmental Engineering, University of California, Riverside, California 92521, United StatesDepartment of Civil & Environmental Engineering, University of Illinois at Urbana-Champaign, Urbana, Illinois 61801, United StatesOrcidhttps://orcid.org/0000-0001-9811-3828
    • Cheng Gu - State Key Laboratory of Pollution Control and Resource Reuse, School of the Environment, Nanjing University, Nanjing 210023, Jiangsu, ChinaOrcidhttps://orcid.org/0000-0002-6939-7101
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

ARTICLE SECTIONS
Jump To

Financial support was provided by the Strategic Environmental Research and Development Program Project ER18-1289 (for Z.L., J.G., and J.L.) and Project ER20-1541 (for Y.Y. and Y.M.).

References

ARTICLE SECTIONS
Jump To

This article references 58 other publications.

  1. 1
    Ng, C.; Cousins, I. T.; DeWitt, J. C.; Glüge, J.; Goldenman, G.; Herzke, D.; Lohmann, R.; Miller, M.; Patton, S.; Scheringer, M.; Trier, X.; Wang, Z. Addressing urgent questions for PFAS in the 21st Century. Environ. Sci. Technol. 2021, 55, 1275512765,  DOI: 10.1021/acs.est.1c03386
  2. 2
    Rericha, Y.; Cao, D.; Truong, L.; Simonich, M.; Field, J. A.; Tanguay, R. L. Behavior effects of structurally diverse per- and polyfluoroalkyl substances in zebrafish. Chem. Res. Toxicol. 2021, 34, 14091416,  DOI: 10.1021/acs.chemrestox.1c00101
  3. 3
    Dasgupta, S.; Reddam, A.; Liu, Z.; Liu, J.; Volz, D. C. High-content screening in zebrafish identifies perfluorooctanesulfonamide as a potent developmental toxicant. Environ. Pollut. 2020, 256, 113550,  DOI: 10.1016/j.envpol.2019.113550
  4. 4
    Nguyen, G. T. H.; Nocentini, A.; Angeli, A.; Gratteri, P.; Supuran, C. T.; Donald, W. A. Perfluoroalkyl substances of significant environmental concern can strongly inhibit human carbonic anhydrase isozymes. Anal. Chem. 2020, 92, 46144622,  DOI: 10.1021/acs.analchem.0c00163
  5. 5
    Harris, M. H.; Rifas-Shiman, S. L.; Calafat, A. M.; Ye, X.; Mora, A. M.; Webster, T. F.; Oken, E.; Sagiv, S. K. Predictors of per- and polyfluoroalkyl substance (PFAS) plasma concentrations in 6–10 year old American children. Environ. Sci. Technol. 2017, 51, 51935204,  DOI: 10.1021/acs.est.6b05811
  6. 6
    Nian, M.; Luo, K.; Luo, F.; Aimuzi, R.; Huo, X.; Chen, Q.; Tian, Y.; Zhang, J. Association between prenatal exposure to PFAS and fetal sex hormones: Are the short-chain PFAS safer?. Environ. Sci. Technol. 2020, 54, 82918299,  DOI: 10.1021/acs.est.0c02444
  7. 7
    Zheng, G.; Schreder, E.; Dempsey, J. C.; Uding, N.; Chu, V.; Andres, G.; Sathyanarayana, S.; Salamova, A. Per- and polyfluoroalkyl substances (PFAS) in breast milk: Concerning trends for current-use PFAS. Environ. Sci. Technol. 2021, 55, 75107520,  DOI: 10.1021/acs.est.0c06978
  8. 8
    Xiao, X.; Ulrich, B. A.; Chen, B.; Higgins, C. P. Sorption of poly- and perfluoroalkyl substances (PFASs) relevant to aqueous film-forming foam (AFFF)-impacted groundwater by biochars and activated carbon. Environ. Sci. Technol. 2017, 51, 63426351,  DOI: 10.1021/acs.est.7b00970
  9. 9
    McCleaf, P.; Englund, S.; Östlund, A.; Lindegren, K.; Wiberg, K.; Ahrens, L. Removal efficiency of multiple poly- and perfluoroalkyl substances (PFASs) in drinking water using granular activated carbon (GAC) and anion exchange (AE) column tests. Water Res. 2017, 120, 7787,  DOI: 10.1016/j.watres.2017.04.057
  10. 10
    Fang, Y.; Ellis, A.; Choi, Y. J.; Boyer, T. H.; Higgins, C. P.; Schaefer, C. E.; Strathmann, T. J. Removal of per-and polyfluoroalkyl substances (PFASs) in aqueous film-forming foam (AFFF) using ion-exchange and nonionic resins. Environ. Sci. Technol. 2021, 55, 50015011,  DOI: 10.1021/acs.est.1c00769
  11. 11
    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,  DOI: 10.1016/j.watres.2021.117781
  12. 12
    Steinle-Darling, E.; Reinhard, M. Nanofiltration for trace organic contaminant removal: Structure, solution, and membrane fouling effects on the rejection of perfluorochemicals. Environ. Sci. Technol. 2008, 42, 52925297,  DOI: 10.1021/es703207s
  13. 13
    Appleman, T. D.; Higgins, C. P.; Quiñones, O.; Vanderford, B. J.; Kolstad, C.; Zeigler-Holady, J. C.; Dickenson, E. R. V. Treatment of poly- and perfluoroalkyl substances in US full-scale water treatment systems. Water Res. 2014, 51, 246255,  DOI: 10.1016/j.watres.2013.10.067
  14. 14
    Sonmez Baghirzade, B.; Zhang, Y.; Reuther, J. F.; Saleh, N. B.; Venkatesan, A. K.; Apul, O. G. Thermal regeneration of spent granular activated carbon presents an opportunity to break the forever PFAS cycle. Environ. Sci. Technol. 2021, 55, 56085619,  DOI: 10.1021/acs.est.0c08224
  15. 15
    Liu, S.; Zhao, S.; Liang, Z.; Wang, F.; Sun, F.; Chen, D. Perfluoroalkyl substances (PFASs) in leachate, fly ash, and bottom ash from waste incineration plants: Implications for the environmental release of PFAS. Sci. Total Environ. 2021, 795, 148468,  DOI: 10.1016/j.scitotenv.2021.148468
  16. 16
    Singh, R. K.; Multari, N.; Nau-Hix, C.; Woodard, S.; Nickelsen, M.; Mededovic Thagard, S.; Holsen, T. M. Removal of poly- and per-fluorinated compounds from ion exchange regenerant still bottom samples in a plasma reactor. Environ. Sci. Technol. 2020, 54, 1397313980,  DOI: 10.1021/acs.est.0c02158
  17. 17
    Qanbarzadeh, M.; Wang, D.; Ateia, M.; Sahu, S. P.; Cates, E. L. Impacts of reactor configuration, degradation mechanisms, and water matrices on perfluorocarboxylic acid treatment efficiency by the UV/Bi3O(OH)(PO4)2 photocatalytic process. ACS ES&T Engg 2021, 1, 239248,  DOI: 10.1021/acsestengg.0c00086
  18. 18
    Zhu, Y.; Xu, T.; Zhao, D.; Li, F.; Liu, W.; Wang, B.; An, B. Adsorption and solid-phase photocatalytic degradation of perfluorooctane sulfonate in water using gallium-doped carbon-modified titanate nanotubes. Chem. Eng. J. 2021, 421, 129676,  DOI: 10.1016/j.cej.2021.129676
  19. 19
    Yang, S.; Fernando, S.; Holsen, T. M.; Yang, Y. Inhibition of perchlorate formation during the electrochemical oxidation of perfluoroalkyl acid in groundwater. Environ. Sci. Technol. Lett. 2019, 6, 775780,  DOI: 10.1021/acs.estlett.9b00653
  20. 20
    Huang, D.; Wang, K.; Niu, J.; Chu, C.; Weon, S.; Zhu, Q.; Lu, J.; Stavitski, E.; Kim, J.-H. Amorphous Pd-loaded Ti4O7 electrode for direct anodic destruction of perfluorooctanoic acid. Environ. Sci. Technol. 2020, 54, 1095410963,  DOI: 10.1021/acs.est.0c03800
  21. 21
    Lewis, A. J.; Joyce, T.; Hadaya, M.; Ebrahimi, F.; Dragiev, I.; Giardetti, N.; Yang, J.; Fridman, G.; Rabinovich, A.; Fridman, A. A.; McKenzie, E. R.; Sales, C. M. Rapid degradation of PFAS in aqueous solutions by reverse vortex flow gliding arc plasma. Environ. Sci.: Water Res. Technol. 2020, 6, 10441057,  DOI: 10.1039/c9ew01050e
  22. 22
    Saleem, M.; Biondo, O.; Sretenović, G.; Tomei, G.; Magarotto, M.; Pavarin, D.; Marotta, E.; Paradisi, C. Comparative performance assessment of plasma reactors for the treatment of PFOA; reactor design, kinetics, mineralization and energy yield. Chem. Eng. J. 2020, 382, 123031,  DOI: 10.1016/j.cej.2019.123031
  23. 23
    Singh, R. K.; Brown, E.; Thagard, S. M.; Holsen, T. M. Treatment of PFAS-containing landfill leachate using an enhanced contact plasma reactor. J. Hazard. Mater. 2021, 408, 124452,  DOI: 10.1016/j.jhazmat.2020.124452
  24. 24
    Huang, L.; Dong, W.; Hou, H. Investigation of the reactivity of hydrated electron toward perfluorinated carboxylates by laser flash photolysis. Chem. Phys. Lett. 2007, 436, 124128,  DOI: 10.1016/j.cplett.2007.01.037
  25. 25
    Park, H.; Vecitis, C. D.; Cheng, J.; Choi, W.; Mader, B. T.; Hoffmann, M. R. Reductive defluorination of aqueous perfluorinated alkyl surfactants: Effects of ionic headgroup and chain length. J. Phys. Chem. A 2009, 113, 690696,  DOI: 10.1021/jp807116q
  26. 26
    Song, Z.; Tang, H.; Wang, N.; Zhu, L. Reductive defluorination of perfluorooctanoic acid by hydrated electrons in a sulfite-mediated UV photochemical system. J. Hazard. Mater. 2013, 262, 332338,  DOI: 10.1016/j.jhazmat.2013.08.059
  27. 27
    Tian, H.; Gao, J.; Li, H.; Boyd, S. A.; Gu, C. Complete defluorination of perfluorinated compounds by hydrated electrons generated from 3-indole-acetic-acid in organomodified montmorillonite. Sci. Rep. 2016, 6, 32949,  DOI: 10.1038/srep32949
  28. 28
    Sun, Z.; Zhang, C.; Xing, L.; Zhou, Q.; Dong, W.; Hoffmann, M. R. UV/nitrilotriacetic acid process as a novel strategy for efficient photoreductive degradation of perfluorooctanesulfonate. Environ. Sci. Technol. 2018, 52, 29532962,  DOI: 10.1021/acs.est.7b05912
  29. 29
    Bentel, M. J.; Yu, Y.; Xu, L.; Li, Z.; Wong, B. M.; Men, Y.; Liu, J. Defluorination of per- and polyfluoroalkyl substances (PFASs) with hydrated electrons: Structural dependence and implications to PFAS remediation and management. Environ. Sci. Technol. 2019, 53, 37183728,  DOI: 10.1021/acs.est.8b06648
  30. 30
    Cui, J.; Gao, P.; Deng, Y. Destruction of per- and polyfluoroalkyl substances (PFAS) with advanced reduction processes (ARPs): A critical review. Environ. Sci. Technol. 2020, 54, 37523766,  DOI: 10.1021/acs.est.9b05565
  31. 31
    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,  DOI: 10.1016/j.watres.2019.114907
  32. 32
    Bentel, M. J.; Yu, Y.; Xu, L.; Kwon, H.; Li, Z.; Wong, B. M.; Men, Y.; Liu, J. Degradation of perfluoroalkyl ether carboxylic acids with hydrated electrons: Structure-reactivity relationships and environmental implications. Environ. Sci. Technol. 2020, 54, 24892499,  DOI: 10.1021/acs.est.9b05869
  33. 33
    Chen, Z.; Li, C.; Gao, J.; Dong, H.; Chen, Y.; Wu, B.; Gu, C. Efficient reductive destruction of perfluoroalkyl substances under self-assembled micelle confinement. Environ. Sci. Technol. 2020, 54, 51785185,  DOI: 10.1021/acs.est.9b06599
  34. 34
    Chen, Z.; Teng, Y.; Mi, N.; Jin, X.; Yang, D.; Wang, C.; Wu, B.; Ren, H.; Zeng, G.; Gu, C. Highly efficient hydrated electron utilization and reductive destruction of perfluoroalkyl substances induced by intermolecular interaction. Environ. Sci. Technol. 2021, 55, 39964006,  DOI: 10.1021/acs.est.0c07927
  35. 35
    Huang, C.; Linkous, C. A.; Adebiyi, O.; T-Raissi, A. Hydrogen production via photolytic oxidation of aqueous sodium sulfite solutions. Environ. Sci. Technol. 2010, 44, 52835288,  DOI: 10.1021/es903766w
  36. 36
    Li, X.; Ma, J.; Liu, G.; Fang, J.; Yue, S.; Guan, Y.; Chen, L.; Liu, X. Efficient reductive dechlorination of monochloroacetic acid by sulfite/UV process. Environ. Sci. Technol. 2012, 46, 73427349,  DOI: 10.1021/es3008535
  37. 37
    Bentel, M. J.; Liu, Z.; Yu, Y.; Gao, J.; Men, Y.; Liu, J. Enhanced degradation of perfluorocarboxylic acids (PFCAs) by UV/sulfite treatment: Reaction mechanisms and system efficiencies at pH 12. Environ. Sci. Technol. Lett. 2020, 7, 351357,  DOI: 10.1021/acs.estlett.0c00236
  38. 38
    Liu, Z.; Bentel, M. J.; Yu, Y.; Ren, C.; Gao, J.; Pulikkal, V. F.; Sun, M.; Men, Y.; Liu, J. Near-quantitative defluorination of perfluorinated and fluorotelomer carboxylates and sulfonates with integrated oxidation and reduction. Environ. Sci. Technol. 2021, 55, 70527062,  DOI: 10.1021/acs.est.1c00353
  39. 39
    Park, H.; Vecitis, C. D.; Cheng, J.; Dalleska, N. F.; Mader, B. T.; Hoffmann, M. R. Reductive degradation of perfluoroalkyl compounds with aquated electrons generated from iodide photolysis at 254 nm. Photochem. Photobiol. Sci. 2011, 10, 19451953,  DOI: 10.1039/c1pp05270e
  40. 40
    Sauer, M. C.; Crowell, R. A.; Shkrob, I. A. Electron photodetachment from aqueous anions. 1. Quantum yields for generation of hydrated electron by 193 and 248 nm laser photoexcitation of miscellaneous inorganic anions. J. Phys. Chem. A 2004, 108, 54905502,  DOI: 10.1021/jp049722t
  41. 41
    Iwata, A.; Nakashima, N.; Kusaba, M.; Izawa, Y.; Yamanaka, C. Quantum yields of hydrated electrons by UV laser irradiation. Chem. Phys. Lett. 1993, 207, 137142,  DOI: 10.1016/0009-2614(93)87004-m
  42. 42
    Qu, Y.; Zhang, C.; Li, F.; Chen, J.; Zhou, Q. Photo-reductive defluorination of perfluorooctanoic acid in water. Water Res. 2010, 44, 29392947,  DOI: 10.1016/j.watres.2010.02.019
  43. 43
    Wong, G. T. F.; Zhang, L.-S. Chemical removal of oxygen with sulfite for the polarographic or voltammetric determination of iodate or iodide in seawater. Mar. Chem. 1992, 38, 109116,  DOI: 10.1016/0304-4203(92)90070-q
  44. 44
    Yu, K.; Li, X.; Chen, L.; Fang, J.; Chen, H.; Li, Q.; Chi, N.; Ma, J. Mechanism and efficiency of contaminant reduction by hydrated electron in the sulfite/iodide/UV process. Water Res. 2018, 129, 357364,  DOI: 10.1016/j.watres.2017.11.030
  45. 45
    Zhang, T.; Wang, J.; Yan, D.; Wang, L.; Liu, X. Efficient reduction of bromate by iodide-assisted UV/sulfite process. Catalysts 2018, 8, 652,  DOI: 10.3390/catal8120652
  46. 46
    Tenorio, R.; Liu, J.; Xiao, X.; Maizel, A.; Higgins, C. P.; Schaefer, C. E.; Strathmann, T. J. Destruction of per- and polyfluoroalkyl substances (PFASs) in aqueous film-forming foam (AFFF) with UV-sulfite photoreductive treatment. Environ. Sci. Technol. 2020, 54, 69576967,  DOI: 10.1021/acs.est.0c00961
  47. 47
    Kireev, S. V.; Shnyrev, S. L. Study of molecular iodine, iodate ions, iodide ions, and triiodide ions solutions absorption in the UV and visible light spectral bands. Laser Phys. 2015, 25, 075602,  DOI: 10.1088/1054-660x/25/7/075602
  48. 48
    Bolton, J. R.; Bircher, K. G.; Tumas, W.; Tolman, C. A. Figures-of-merit for the technical development and application of advanced oxidation processes. J. Adv. Oxid. Technol. 1996, 1, 1317,  DOI: 10.1515/jaots-1996-0104
  49. 49
    Nzeribe, B. N.; Crimi, M.; Mededovic Thagard, S.; Holsen, T. M. Physico-chemical processes for the treatment of per- and polyfluoroalkyl substances (PFAS): A review. Crit. Rev. Environ. Sci. Technol. 2019, 49, 866915,  DOI: 10.1080/10643389.2018.1542916
  50. 50
    Gao, J.; Liu, Z.; Bentel, M. J.; Yu, Y.; Men, Y.; Liu, J. Defluorination of omega-hydroperfluorocarboxylates (ω-HPFCAs): Distinct reactivities from perfluoro and fluorotelomeric carboxylates. Environ. Sci. Technol. 2021, 55, 1414614155,  DOI: 10.1021/acs.est.1c04429
  51. 51
    Buxton, G. V.; Greenstock, C. L.; Helman, W. P.; Ross, A. B. Critical review of rate constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl radicals (·OH/·O) in aqueous solution. J. Phys. Chem. Ref. Data 1988, 17, 513886,  DOI: 10.1063/1.555805
  52. 52
    Elliot, A. J. A pulse radiolysis study of the reaction of OH with I2 and the decay of I2. Can. J. Chem. 1992, 70, 16581661,  DOI: 10.1139/v92-207
  53. 53
    Yiin, B. S.; Margerum, D. W. Nonmetal redox kinetics: reactions of iodine and triiodide with sulfite and hydrogen sulfite and the hydrolysis of iodosulfate. Inorg. Chem. 1990, 29, 15591564,  DOI: 10.1021/ic00333a023
  54. 54
    Neta, P.; Huie, R. E.; Ross, A. B. Rate constants for reactions of inorganic radicals in aqueous solution. J. Phys. Chem. Ref. Data 1988, 17, 10271284,  DOI: 10.1063/1.555808
  55. 55
    Li, X.; Fang, J.; Liu, G.; Zhang, S.; Pan, B.; Ma, J. Kinetics and efficiency of the hydrated electron-induced dehalogenation by the sulfite/UV process. Water Res. 2014, 62, 220228,  DOI: 10.1016/j.watres.2014.05.051
  56. 56
    Bray, W. C.; Liebhafsky, H. A. Reactions involving hydrogen peroxide, iodine and iodate ion. I. Introduction. J. Am. Chem. Soc. 1931, 53, 3844,  DOI: 10.1021/ja01352a006
  57. 57
    Gregor, H. P.; Belle, J.; Marcus, R. A. Studies on ion-exchange resins. XIII. Selectivity coefficients of quaternary base anion-exchange resins toward univalent anions. J. Am. Chem. Soc. 1955, 77, 27132719,  DOI: 10.1021/ja01615a011
  58. 58
    Hildebrand, J. H.; Benesi, H. A.; Mower, L. M. Solubility of iodine in ethyl alcohol, ethyl ether, mesitylene, p-xylene, 2,2-dimethylbutane, cyclohexane and perfluoro-n-heptane. J. Am. Chem. Soc. 1950, 72, 10171020,  DOI: 10.1021/ja01158a096

Cited By

ARTICLE SECTIONS
Jump To

This article is cited by 24 publications.

  1. Yuwen Qi, Yinbo Yang, Shengyan Cui, Xuejiao Tang, Peng Zhang, Cuiping Wang, Yanna Liang, Hongwen Sun, Chuanxin Ma, Baoshan Xing. Novel Defluorination Pathways of Perfluoroether Compounds (GenX): α-Fe2O3 Nanoparticle Layer Retains Higher Concentrations of Effective Hydrated Electrons. Environmental Science & Technology 2024, 58 (12) , 5567-5577. https://doi.org/10.1021/acs.est.3c09879
  2. Zhanghao Chen, Ruochen Dong, Xinhao Wang, Liuqing Huang, Longlong Qiu, Ming Zhang, Na Mi, Min Xu, Huan He, Cheng Gu. Efficient Decomposition of Perfluoroalkyl Substances by Low Concentration Indole: New Insights into the Molecular Mechanisms. Environmental Science & Technology 2024, 58 (7) , 3530-3539. https://doi.org/10.1021/acs.est.3c08453
  3. Kejian Li, Wenbo You, Wei Wang, Kedong Gong, Yangyang Liu, Longqian Wang, Qiuyue Ge, Xuejun Ruan, Jianpeng Ao, Minbiao Ji, Liwu Zhang. Significantly Accelerated Photochemical Perfluorooctanoic Acid Decomposition at the Air–Water Interface of Microdroplets. Environmental Science & Technology 2023, 57 (50) , 21448-21458. https://doi.org/10.1021/acs.est.3c05470
  4. Bosen Jin, Yiwen Zhu, Weiyang Zhao, Zekun Liu, Shun Che, Kunpeng Chen, Ying-Hsuan Lin, Jinyong Liu, Yujie Men. Aerobic Biotransformation and Defluorination of Fluoroalkylether Substances (ether PFAS): Substrate Specificity, Pathways, and Applications. Environmental Science & Technology Letters 2023, 10 (9) , 755-761. https://doi.org/10.1021/acs.estlett.3c00411
  5. Zhanghao Chen, Shuoqi Zhang, Xinhao Wang, Na Mi, Ming Zhang, Guixiang Zeng, Hailiang Dong, Jinyong Liu, Bing Wu, Si Wei, Cheng Gu. Amine-Functionalized A-Center Sphalerite for Selective and Efficient Destruction of Perfluorooctanoic Acid. Environmental Science & Technology 2023, 57 (28) , 10438-10447. https://doi.org/10.1021/acs.est.3c01266
  6. Camille K. Amador, Haden Cavalli, Raul Tenorio, Hanna Tetu, Christopher P. Higgins, Shubham Vyas, Timothy J. Strathmann. Influence of Carbonate Speciation on Hydrated Electron Treatment Processes. Environmental Science & Technology 2023, 57 (20) , 7849-7857. https://doi.org/10.1021/acs.est.2c09451
  7. Yue Wang, Ye Xiao, Yafei Wang, Qingqi Lin, Yanping Zhu, Zhuobiao Ni, Rongliang Qiu. Electroreductive Defluorination of Unsaturated PFAS by a Quaternary Ammonium Surfactant-Modified Cathode via Direct Cathodic Reduction. Environmental Science & Technology 2023, 57 (19) , 7578-7589. https://doi.org/10.1021/acs.est.2c08182
  8. Min Yang, Zhongcheng Du, Hongjia Bao, Xiaolei Zhang, Qiang Liu, Wenshan Guo, Huu-Hao Ngo, Long D. Nghiem. Experimental and Theoretical Insight of Perfluorooctanoic Acid Destruction by Alkaline Hydrothermal Treatment Enhanced with Zero-Valent Iron in Biochar. ACS ES&T Water 2023, 3 (5) , 1286-1293. https://doi.org/10.1021/acsestwater.2c00614
  9. Zhiwen Jiang, Daniel Adjei, Sergey A. Denisov, Mehran Mostafavi, Jun Ma. Transient Kinetics of Short-Chain Perfluoroalkyl Sulfonate with Radiolytic Reducing Species. Environmental Science & Technology Letters 2023, 10 (1) , 59-65. https://doi.org/10.1021/acs.estlett.2c00837
  10. Ziren Wan, Lifeng Cao, Wan Huang, Di Zheng, Guanghe Li, Fang Zhang. Enhanced Electrochemical Destruction of Perfluorooctanoic Acid (PFOA) Aided by Overlooked Cathodically Produced Bubbles. Environmental Science & Technology Letters 2023, 10 (1) , 111-116. https://doi.org/10.1021/acs.estlett.2c00897
  11. Hubertus Brunn, Gottfried Arnold, Wolfgang Körner, Gerd Rippen, Klaus Günter Steinhäuser, Ingo Valentin. PFAS: forever chemicals—persistent, bioaccumulative and mobile. Reviewing the status and the need for their phase out and remediation of contaminated sites. Environmental Sciences Europe 2023, 35 (1) https://doi.org/10.1186/s12302-023-00721-8
  12. Fenghui Qin, Wenli Yao, Yanlong Liu, Bingjie Zhu, Qianxi Yang, Yian Zheng. Polyethyleneimine functionalized cellulose-rich agroforestry residues for removing perfluorooctanoic acid: adsorption performance and mechanism. Cellulose 2023, 287 https://doi.org/10.1007/s10570-023-05090-2
  13. Yulong Li, Lixin Lv, Lie Yang, Liuyang He, Yulin Chen, Li Wu, Zulin Zhang. Boosting degradation and defluorination efficiencies of PFBS in a vacuum-ultraviolet/S(Ⅳ) process with iodide involvement. Chemosphere 2023, 313 , 137531. https://doi.org/10.1016/j.chemosphere.2022.137531
  14. Qifeng Fan, Tiantian Gong, Qianling Dong, Wenbing Wang. Uncovering hydrothermal treatment of per- and polyfluoroalkyl substances. Eco-Environment & Health 2023, 377 https://doi.org/10.1016/j.eehl.2023.02.002
  15. Camille K. Amador, Daniel J. Van Hoomissen, Jiaoqin Liu, Timothy J. Strathmann, Shubham Vyas. Ultra-short chain fluorocarboxylates exhibit wide ranging reactivity with hydrated electrons. Chemosphere 2023, 311 , 136918. https://doi.org/10.1016/j.chemosphere.2022.136918
  16. Jarno Reuna, Raul Kanter, Niklas Kandelin, Kaisa Kiuru, Heli Koivuluoto, Mircea Guina. Ice resistance of hydrophobic fluoropolymerized nanostructured alumina films for antireflective coatings. Nano Express 2022, 3 (4) , 044002. https://doi.org/10.1088/2632-959X/acad19
  17. Bhavini Saawarn, Byomkesh Mahanty, Subrata Hait, Sahid Hussain. Sources, occurrence, and treatment techniques of per- and polyfluoroalkyl substances in aqueous matrices: A comprehensive review. Environmental Research 2022, 214 , 114004. https://doi.org/10.1016/j.envres.2022.114004
  18. Gongde Chen, Sitao Liu, Qingyang Shi, Jay Gan, Bosen Jin, Yujie Men, Haizhou Liu. Hydrogen-polarized vacuum ultraviolet photolysis system for enhanced destruction of perfluoroalkyl substances. Journal of Hazardous Materials Letters 2022, 3 , 100072. https://doi.org/10.1016/j.hazl.2022.100072
  19. Fuqiang Liu, Xiaohong Guan, Feng Xiao. Photodegradation of per- and polyfluoroalkyl substances in water: A review of fundamentals and applications. Journal of Hazardous Materials 2022, 439 , 129580. https://doi.org/10.1016/j.jhazmat.2022.129580
  20. Zhanghao Chen, Xinhao Wang, Ruochen Dong, Yutong Zhang, Xin Jin, Cheng Gu. Challenging the contamination of per- and polyfluoroalkyl substances in water: advanced oxidation or reduction?. Environmental Functional Materials 2022, 1 (3) , 325-337. https://doi.org/10.1016/j.efmat.2022.11.002
  21. Yurong Gu, Zi Song, Zijun Dong, Feiyun Sun, Chengchun Jiang, Jikun Qi. Efficient degradation and deiodination of iopamidol by UV/sulfite process: Assessment of typical process parameters and transformation paths. Environment International 2022, 167 , 107383. https://doi.org/10.1016/j.envint.2022.107383
  22. Yongjie Fan, Jing Yang, Kaicong Cai, Zhilei Lu, Jiejie Chen, Xingteng Lai, Zeping Xu, Zhenle He, Yuyi Zheng, Changqing Liu, Qiyuan Sun, Rongkun Jian, Feifeng Wang. Mechanism of 9,10-Dihydro-9-oxa-10-phosphaphenanthrene-10-oxide degradation in UV light-emitting diodes lamp driven sodium sulfite activation process. Process Safety and Environmental Protection 2022, 165 , 704-715. https://doi.org/10.1016/j.psep.2022.07.064
  23. Chenguang Li, Yifei Wang, Yaye Wang, Zunyao Wang, Qingguo Huang. Electrochemical oxidation combined with UV irradiation for synergistic removal of perfluorooctane sulfonate (PFOS) in water. Journal of Hazardous Materials 2022, 436 , 129091. https://doi.org/10.1016/j.jhazmat.2022.129091
  24. Yanqing Cong, Lidong Shen, Baimei Wang, Jianlai Cao, Zixuan Pan, Ziyu Wang, Kai Wang, Qiangbiao Li, Xuchun Li. Efficient removal of Cr(VI) at alkaline pHs by sulfite/iodide/UV: Mechanism and modeling. Water Research 2022, 222 , 118919. https://doi.org/10.1016/j.watres.2022.118919
  • Abstract

    Figure 1

    Figure 1. (a) Parent compound decay and (b) defluorination of PFBS by UV/S (10 mM), UV/I (2 mM), and UV/S (10 mM) + I (2 mM); PFBS degradation by UV/S + I at varied (c,d) iodide concentrations, (e) sulfite concentrations, and (f) solution pH. Default reaction conditions: PFBS (0.025 mM), carbonate (5 mM), pH 12.0, 254 nm irradiation (an 18 W low-pressure Hg lamp in 600 mL solution), and 20 °C.

    Figure 2

    Figure 2. (a) Parent compound decay and (b) defluorination of n = 1–7 PFCAs by UV/S (10 mM), and UV/S (10 mM) + I (2 mM). Panel (c) shows the magnified defluorination profiles for n = 2–7 PFCAs (averaged) within the first 1 h. (d–f) Parent compound decay and defluorination of n = 4, 6, 8 PFSAs. Reaction conditions: individual PFAS (0.025 mM), NaHCO3 (5 mM), pH 12.0, 254 nm irradiation (an 18 W low-pressure Hg lamp in 600 mL solution), and 20 °C.

    Figure 3

    Figure 3. Parent compound decay, TP formation, and defluorination of tri-, di-, and monofluoro acetates by (a–c) UV/S + I and (d–f) UV/S and (g) reaction schemes. Reaction conditions: individual PFAS (0.025 mM), Na2SO3 (10 mM), KI (2 mM), NaHCO3 (5 mM), pH 12.0, 254 nm irradiation (an 18 W low-pressure Hg lamp in 600 mL solution), and 20 °C. The vertical red dotted lines in panels (d–f) indicate the 1 h time window where 100% defluorination by UV/S + I was achieved.

    Figure 4

    Figure 4. Parent compound decay, TP formation, and defluorination of PFPrA by (a,b) UV/S + I and (c,d) UV/S. Panels (b,d) show magnified time profiles within the first hour of reaction. Panel (e) shows the proposed reaction schemes. Reaction conditions: PFPrA (0.025 mM), Na2SO3 (10 mM), KI (2 mM), NaHCO3 (5 mM), pH 12.0, 254 nm irradiation (an 18 W low-pressure Hg lamp in 600 mL solution), and 20 °C.

    Figure 5

    Figure 5. Parent compound decay and TP formation for the degradation of (a,b) PFOA and (c,d) PFBS by UV/S + I and UV/S. Panel (e) shows the proposed reaction schemes. Reaction conditions were the following: individual PFAS (0.025 mM), Na2SO3 (10 mM), KI (2 mM), NaHCO3 (5 mM), pH 12.0, 254 nm irradiation (an 18 W low-pressure Hg lamp in 600 mL solution), and 20 °C.

    Figure 6

    Figure 6. (a) Kinetic profiles of transient absorption at 700 nm after the 266 nm laser pulse excitation of I, S, and S + I solutions; (b) kinetic profiles of transient absorption at 350 nm after 266 nm laser pulse excitation of I and S + I solutions; (c) decay of sulfite in UV/S and UV/S + I in the absence of PFAS; and (d) lifetime of eaq in UV/S + I at varied solution pH. Conditions: Na2SO3 (S, 10 mM), KI (I, 2 mM), and pH 12.0.

    Figure 7

    Figure 7. (a) DeF% of PFCAs by UV/S after 8 h; (b) time profiles for the defluorination from 0.25 mM PFOA by UV/S (with SO32– spikes every 24 h) and UV/S + I (no SO32– spike); (c) time profiles for the defluorination from various [PFOA]0 by UV/S + I; (d) deF% of PFCAs after 24 h by UV/S + I; (e) deF% of PFSAs by UV/I and UV/S + I after 24 h; (f) time profiles for the defluorination from the mixed PFAS (see Table 2) in NaCl brine after a 1:1 dilution with deionized (DI) water. Default reaction conditions: 10 mM Na2SO3 (10 mM), KI (2 mM for UV/S + I), NaHCO3 (5 mM), pH 12.0, and 254 nm irradiation (an 18 W low-pressure Hg lamp for 600 mL solution), and 20 °C.

  • References

    ARTICLE SECTIONS
    Jump To

    This article references 58 other publications.

    1. 1
      Ng, C.; Cousins, I. T.; DeWitt, J. C.; Glüge, J.; Goldenman, G.; Herzke, D.; Lohmann, R.; Miller, M.; Patton, S.; Scheringer, M.; Trier, X.; Wang, Z. Addressing urgent questions for PFAS in the 21st Century. Environ. Sci. Technol. 2021, 55, 1275512765,  DOI: 10.1021/acs.est.1c03386
    2. 2
      Rericha, Y.; Cao, D.; Truong, L.; Simonich, M.; Field, J. A.; Tanguay, R. L. Behavior effects of structurally diverse per- and polyfluoroalkyl substances in zebrafish. Chem. Res. Toxicol. 2021, 34, 14091416,  DOI: 10.1021/acs.chemrestox.1c00101
    3. 3
      Dasgupta, S.; Reddam, A.; Liu, Z.; Liu, J.; Volz, D. C. High-content screening in zebrafish identifies perfluorooctanesulfonamide as a potent developmental toxicant. Environ. Pollut. 2020, 256, 113550,  DOI: 10.1016/j.envpol.2019.113550
    4. 4
      Nguyen, G. T. H.; Nocentini, A.; Angeli, A.; Gratteri, P.; Supuran, C. T.; Donald, W. A. Perfluoroalkyl substances of significant environmental concern can strongly inhibit human carbonic anhydrase isozymes. Anal. Chem. 2020, 92, 46144622,  DOI: 10.1021/acs.analchem.0c00163
    5. 5
      Harris, M. H.; Rifas-Shiman, S. L.; Calafat, A. M.; Ye, X.; Mora, A. M.; Webster, T. F.; Oken, E.; Sagiv, S. K. Predictors of per- and polyfluoroalkyl substance (PFAS) plasma concentrations in 6–10 year old American children. Environ. Sci. Technol. 2017, 51, 51935204,  DOI: 10.1021/acs.est.6b05811
    6. 6
      Nian, M.; Luo, K.; Luo, F.; Aimuzi, R.; Huo, X.; Chen, Q.; Tian, Y.; Zhang, J. Association between prenatal exposure to PFAS and fetal sex hormones: Are the short-chain PFAS safer?. Environ. Sci. Technol. 2020, 54, 82918299,  DOI: 10.1021/acs.est.0c02444
    7. 7
      Zheng, G.; Schreder, E.; Dempsey, J. C.; Uding, N.; Chu, V.; Andres, G.; Sathyanarayana, S.; Salamova, A. Per- and polyfluoroalkyl substances (PFAS) in breast milk: Concerning trends for current-use PFAS. Environ. Sci. Technol. 2021, 55, 75107520,  DOI: 10.1021/acs.est.0c06978
    8. 8
      Xiao, X.; Ulrich, B. A.; Chen, B.; Higgins, C. P. Sorption of poly- and perfluoroalkyl substances (PFASs) relevant to aqueous film-forming foam (AFFF)-impacted groundwater by biochars and activated carbon. Environ. Sci. Technol. 2017, 51, 63426351,  DOI: 10.1021/acs.est.7b00970
    9. 9
      McCleaf, P.; Englund, S.; Östlund, A.; Lindegren, K.; Wiberg, K.; Ahrens, L. Removal efficiency of multiple poly- and perfluoroalkyl substances (PFASs) in drinking water using granular activated carbon (GAC) and anion exchange (AE) column tests. Water Res. 2017, 120, 7787,  DOI: 10.1016/j.watres.2017.04.057
    10. 10
      Fang, Y.; Ellis, A.; Choi, Y. J.; Boyer, T. H.; Higgins, C. P.; Schaefer, C. E.; Strathmann, T. J. Removal of per-and polyfluoroalkyl substances (PFASs) in aqueous film-forming foam (AFFF) using ion-exchange and nonionic resins. Environ. Sci. Technol. 2021, 55, 50015011,  DOI: 10.1021/acs.est.1c00769
    11. 11
      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,  DOI: 10.1016/j.watres.2021.117781
    12. 12
      Steinle-Darling, E.; Reinhard, M. Nanofiltration for trace organic contaminant removal: Structure, solution, and membrane fouling effects on the rejection of perfluorochemicals. Environ. Sci. Technol. 2008, 42, 52925297,  DOI: 10.1021/es703207s
    13. 13
      Appleman, T. D.; Higgins, C. P.; Quiñones, O.; Vanderford, B. J.; Kolstad, C.; Zeigler-Holady, J. C.; Dickenson, E. R. V. Treatment of poly- and perfluoroalkyl substances in US full-scale water treatment systems. Water Res. 2014, 51, 246255,  DOI: 10.1016/j.watres.2013.10.067
    14. 14
      Sonmez Baghirzade, B.; Zhang, Y.; Reuther, J. F.; Saleh, N. B.; Venkatesan, A. K.; Apul, O. G. Thermal regeneration of spent granular activated carbon presents an opportunity to break the forever PFAS cycle. Environ. Sci. Technol. 2021, 55, 56085619,  DOI: 10.1021/acs.est.0c08224
    15. 15
      Liu, S.; Zhao, S.; Liang, Z.; Wang, F.; Sun, F.; Chen, D. Perfluoroalkyl substances (PFASs) in leachate, fly ash, and bottom ash from waste incineration plants: Implications for the environmental release of PFAS. Sci. Total Environ. 2021, 795, 148468,  DOI: 10.1016/j.scitotenv.2021.148468
    16. 16
      Singh, R. K.; Multari, N.; Nau-Hix, C.; Woodard, S.; Nickelsen, M.; Mededovic Thagard, S.; Holsen, T. M. Removal of poly- and per-fluorinated compounds from ion exchange regenerant still bottom samples in a plasma reactor. Environ. Sci. Technol. 2020, 54, 1397313980,  DOI: 10.1021/acs.est.0c02158
    17. 17
      Qanbarzadeh, M.; Wang, D.; Ateia, M.; Sahu, S. P.; Cates, E. L. Impacts of reactor configuration, degradation mechanisms, and water matrices on perfluorocarboxylic acid treatment efficiency by the UV/Bi3O(OH)(PO4)2 photocatalytic process. ACS ES&T Engg 2021, 1, 239248,  DOI: 10.1021/acsestengg.0c00086
    18. 18
      Zhu, Y.; Xu, T.; Zhao, D.; Li, F.; Liu, W.; Wang, B.; An, B. Adsorption and solid-phase photocatalytic degradation of perfluorooctane sulfonate in water using gallium-doped carbon-modified titanate nanotubes. Chem. Eng. J. 2021, 421, 129676,  DOI: 10.1016/j.cej.2021.129676
    19. 19
      Yang, S.; Fernando, S.; Holsen, T. M.; Yang, Y. Inhibition of perchlorate formation during the electrochemical oxidation of perfluoroalkyl acid in groundwater. Environ. Sci. Technol. Lett. 2019, 6, 775780,  DOI: 10.1021/acs.estlett.9b00653
    20. 20
      Huang, D.; Wang, K.; Niu, J.; Chu, C.; Weon, S.; Zhu, Q.; Lu, J.; Stavitski, E.; Kim, J.-H. Amorphous Pd-loaded Ti4O7 electrode for direct anodic destruction of perfluorooctanoic acid. Environ. Sci. Technol. 2020, 54, 1095410963,  DOI: 10.1021/acs.est.0c03800
    21. 21
      Lewis, A. J.; Joyce, T.; Hadaya, M.; Ebrahimi, F.; Dragiev, I.; Giardetti, N.; Yang, J.; Fridman, G.; Rabinovich, A.; Fridman, A. A.; McKenzie, E. R.; Sales, C. M. Rapid degradation of PFAS in aqueous solutions by reverse vortex flow gliding arc plasma. Environ. Sci.: Water Res. Technol. 2020, 6, 10441057,  DOI: 10.1039/c9ew01050e
    22. 22
      Saleem, M.; Biondo, O.; Sretenović, G.; Tomei, G.; Magarotto, M.; Pavarin, D.; Marotta, E.; Paradisi, C. Comparative performance assessment of plasma reactors for the treatment of PFOA; reactor design, kinetics, mineralization and energy yield. Chem. Eng. J. 2020, 382, 123031,  DOI: 10.1016/j.cej.2019.123031
    23. 23
      Singh, R. K.; Brown, E.; Thagard, S. M.; Holsen, T. M. Treatment of PFAS-containing landfill leachate using an enhanced contact plasma reactor. J. Hazard. Mater. 2021, 408, 124452,  DOI: 10.1016/j.jhazmat.2020.124452
    24. 24
      Huang, L.; Dong, W.; Hou, H. Investigation of the reactivity of hydrated electron toward perfluorinated carboxylates by laser flash photolysis. Chem. Phys. Lett. 2007, 436, 124128,  DOI: 10.1016/j.cplett.2007.01.037
    25. 25
      Park, H.; Vecitis, C. D.; Cheng, J.; Choi, W.; Mader, B. T.; Hoffmann, M. R. Reductive defluorination of aqueous perfluorinated alkyl surfactants: Effects of ionic headgroup and chain length. J. Phys. Chem. A 2009, 113, 690696,  DOI: 10.1021/jp807116q
    26. 26
      Song, Z.; Tang, H.; Wang, N.; Zhu, L. Reductive defluorination of perfluorooctanoic acid by hydrated electrons in a sulfite-mediated UV photochemical system. J. Hazard. Mater. 2013, 262, 332338,  DOI: 10.1016/j.jhazmat.2013.08.059
    27. 27
      Tian, H.; Gao, J.; Li, H.; Boyd, S. A.; Gu, C. Complete defluorination of perfluorinated compounds by hydrated electrons generated from 3-indole-acetic-acid in organomodified montmorillonite. Sci. Rep. 2016, 6, 32949,  DOI: 10.1038/srep32949
    28. 28
      Sun, Z.; Zhang, C.; Xing, L.; Zhou, Q.; Dong, W.; Hoffmann, M. R. UV/nitrilotriacetic acid process as a novel strategy for efficient photoreductive degradation of perfluorooctanesulfonate. Environ. Sci. Technol. 2018, 52, 29532962,  DOI: 10.1021/acs.est.7b05912
    29. 29
      Bentel, M. J.; Yu, Y.; Xu, L.; Li, Z.; Wong, B. M.; Men, Y.; Liu, J. Defluorination of per- and polyfluoroalkyl substances (PFASs) with hydrated electrons: Structural dependence and implications to PFAS remediation and management. Environ. Sci. Technol. 2019, 53, 37183728,  DOI: 10.1021/acs.est.8b06648
    30. 30
      Cui, J.; Gao, P.; Deng, Y. Destruction of per- and polyfluoroalkyl substances (PFAS) with advanced reduction processes (ARPs): A critical review. Environ. Sci. Technol. 2020, 54, 37523766,  DOI: 10.1021/acs.est.9b05565
    31. 31
      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,  DOI: 10.1016/j.watres.2019.114907
    32. 32
      Bentel, M. J.; Yu, Y.; Xu, L.; Kwon, H.; Li, Z.; Wong, B. M.; Men, Y.; Liu, J. Degradation of perfluoroalkyl ether carboxylic acids with hydrated electrons: Structure-reactivity relationships and environmental implications. Environ. Sci. Technol. 2020, 54, 24892499,  DOI: 10.1021/acs.est.9b05869
    33. 33
      Chen, Z.; Li, C.; Gao, J.; Dong, H.; Chen, Y.; Wu, B.; Gu, C. Efficient reductive destruction of perfluoroalkyl substances under self-assembled micelle confinement. Environ. Sci. Technol. 2020, 54, 51785185,  DOI: 10.1021/acs.est.9b06599
    34. 34
      Chen, Z.; Teng, Y.; Mi, N.; Jin, X.; Yang, D.; Wang, C.; Wu, B.; Ren, H.; Zeng, G.; Gu, C. Highly efficient hydrated electron utilization and reductive destruction of perfluoroalkyl substances induced by intermolecular interaction. Environ. Sci. Technol. 2021, 55, 39964006,  DOI: 10.1021/acs.est.0c07927
    35. 35
      Huang, C.; Linkous, C. A.; Adebiyi, O.; T-Raissi, A. Hydrogen production via photolytic oxidation of aqueous sodium sulfite solutions. Environ. Sci. Technol. 2010, 44, 52835288,  DOI: 10.1021/es903766w
    36. 36
      Li, X.; Ma, J.; Liu, G.; Fang, J.; Yue, S.; Guan, Y.; Chen, L.; Liu, X. Efficient reductive dechlorination of monochloroacetic acid by sulfite/UV process. Environ. Sci. Technol. 2012, 46, 73427349,  DOI: 10.1021/es3008535
    37. 37
      Bentel, M. J.; Liu, Z.; Yu, Y.; Gao, J.; Men, Y.; Liu, J. Enhanced degradation of perfluorocarboxylic acids (PFCAs) by UV/sulfite treatment: Reaction mechanisms and system efficiencies at pH 12. Environ. Sci. Technol. Lett. 2020, 7, 351357,  DOI: 10.1021/acs.estlett.0c00236
    38. 38
      Liu, Z.; Bentel, M. J.; Yu, Y.; Ren, C.; Gao, J.; Pulikkal, V. F.; Sun, M.; Men, Y.; Liu, J. Near-quantitative defluorination of perfluorinated and fluorotelomer carboxylates and sulfonates with integrated oxidation and reduction. Environ. Sci. Technol. 2021, 55, 70527062,  DOI: 10.1021/acs.est.1c00353
    39. 39
      Park, H.; Vecitis, C. D.; Cheng, J.; Dalleska, N. F.; Mader, B. T.; Hoffmann, M. R. Reductive degradation of perfluoroalkyl compounds with aquated electrons generated from iodide photolysis at 254 nm. Photochem. Photobiol. Sci. 2011, 10, 19451953,  DOI: 10.1039/c1pp05270e
    40. 40
      Sauer, M. C.; Crowell, R. A.; Shkrob, I. A. Electron photodetachment from aqueous anions. 1. Quantum yields for generation of hydrated electron by 193 and 248 nm laser photoexcitation of miscellaneous inorganic anions. J. Phys. Chem. A 2004, 108, 54905502,  DOI: 10.1021/jp049722t
    41. 41
      Iwata, A.; Nakashima, N.; Kusaba, M.; Izawa, Y.; Yamanaka, C. Quantum yields of hydrated electrons by UV laser irradiation. Chem. Phys. Lett. 1993, 207, 137142,  DOI: 10.1016/0009-2614(93)87004-m
    42. 42
      Qu, Y.; Zhang, C.; Li, F.; Chen, J.; Zhou, Q. Photo-reductive defluorination of perfluorooctanoic acid in water. Water Res. 2010, 44, 29392947,  DOI: 10.1016/j.watres.2010.02.019
    43. 43
      Wong, G. T. F.; Zhang, L.-S. Chemical removal of oxygen with sulfite for the polarographic or voltammetric determination of iodate or iodide in seawater. Mar. Chem. 1992, 38, 109116,  DOI: 10.1016/0304-4203(92)90070-q
    44. 44
      Yu, K.; Li, X.; Chen, L.; Fang, J.; Chen, H.; Li, Q.; Chi, N.; Ma, J. Mechanism and efficiency of contaminant reduction by hydrated electron in the sulfite/iodide/UV process. Water Res. 2018, 129, 357364,  DOI: 10.1016/j.watres.2017.11.030
    45. 45
      Zhang, T.; Wang, J.; Yan, D.; Wang, L.; Liu, X. Efficient reduction of bromate by iodide-assisted UV/sulfite process. Catalysts 2018, 8, 652,  DOI: 10.3390/catal8120652
    46. 46
      Tenorio, R.; Liu, J.; Xiao, X.; Maizel, A.; Higgins, C. P.; Schaefer, C. E.; Strathmann, T. J. Destruction of per- and polyfluoroalkyl substances (PFASs) in aqueous film-forming foam (AFFF) with UV-sulfite photoreductive treatment. Environ. Sci. Technol. 2020, 54, 69576967,  DOI: 10.1021/acs.est.0c00961
    47. 47
      Kireev, S. V.; Shnyrev, S. L. Study of molecular iodine, iodate ions, iodide ions, and triiodide ions solutions absorption in the UV and visible light spectral bands. Laser Phys. 2015, 25, 075602,  DOI: 10.1088/1054-660x/25/7/075602
    48. 48
      Bolton, J. R.; Bircher, K. G.; Tumas, W.; Tolman, C. A. Figures-of-merit for the technical development and application of advanced oxidation processes. J. Adv. Oxid. Technol. 1996, 1, 1317,  DOI: 10.1515/jaots-1996-0104
    49. 49
      Nzeribe, B. N.; Crimi, M.; Mededovic Thagard, S.; Holsen, T. M. Physico-chemical processes for the treatment of per- and polyfluoroalkyl substances (PFAS): A review. Crit. Rev. Environ. Sci. Technol. 2019, 49, 866915,  DOI: 10.1080/10643389.2018.1542916
    50. 50
      Gao, J.; Liu, Z.; Bentel, M. J.; Yu, Y.; Men, Y.; Liu, J. Defluorination of omega-hydroperfluorocarboxylates (ω-HPFCAs): Distinct reactivities from perfluoro and fluorotelomeric carboxylates. Environ. Sci. Technol. 2021, 55, 1414614155,  DOI: 10.1021/acs.est.1c04429
    51. 51
      Buxton, G. V.; Greenstock, C. L.; Helman, W. P.; Ross, A. B. Critical review of rate constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl radicals (·OH/·O) in aqueous solution. J. Phys. Chem. Ref. Data 1988, 17, 513886,  DOI: 10.1063/1.555805
    52. 52
      Elliot, A. J. A pulse radiolysis study of the reaction of OH with I2 and the decay of I2. Can. J. Chem. 1992, 70, 16581661,  DOI: 10.1139/v92-207
    53. 53
      Yiin, B. S.; Margerum, D. W. Nonmetal redox kinetics: reactions of iodine and triiodide with sulfite and hydrogen sulfite and the hydrolysis of iodosulfate. Inorg. Chem. 1990, 29, 15591564,  DOI: 10.1021/ic00333a023
    54. 54
      Neta, P.; Huie, R. E.; Ross, A. B. Rate constants for reactions of inorganic radicals in aqueous solution. J. Phys. Chem. Ref. Data 1988, 17, 10271284,  DOI: 10.1063/1.555808
    55. 55
      Li, X.; Fang, J.; Liu, G.; Zhang, S.; Pan, B.; Ma, J. Kinetics and efficiency of the hydrated electron-induced dehalogenation by the sulfite/UV process. Water Res. 2014, 62, 220228,  DOI: 10.1016/j.watres.2014.05.051
    56. 56
      Bray, W. C.; Liebhafsky, H. A. Reactions involving hydrogen peroxide, iodine and iodate ion. I. Introduction. J. Am. Chem. Soc. 1931, 53, 3844,  DOI: 10.1021/ja01352a006
    57. 57
      Gregor, H. P.; Belle, J.; Marcus, R. A. Studies on ion-exchange resins. XIII. Selectivity coefficients of quaternary base anion-exchange resins toward univalent anions. J. Am. Chem. Soc. 1955, 77, 27132719,  DOI: 10.1021/ja01615a011
    58. 58
      Hildebrand, J. H.; Benesi, H. A.; Mower, L. M. Solubility of iodine in ethyl alcohol, ethyl ether, mesitylene, p-xylene, 2,2-dimethylbutane, cyclohexane and perfluoro-n-heptane. J. Am. Chem. Soc. 1950, 72, 10171020,  DOI: 10.1021/ja01158a096
  • Supporting Information

    Supporting Information

    ARTICLE SECTIONS
    Jump To

    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.est.1c07608.

    • Details of PFAS chemicals; quantitation of PFAS parent compounds, TPs, and anions; extended discussions on kinetic data fitting and photoreactor configurations; and defluorination profiles in the two photoreactor configurations (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect