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

Chemosphere 286 (2022) 131685

Contents lists available at ScienceDirect

Chemosphere
journal homepage: www.elsevier.com/locate/chemosphere

Thermal decomposition of perfluorinated carboxylic acids: Kinetic model


and theoretical requirements for PFAS incineration
Mohammednoor Altarawneh a, *, Mansour H. Almatarneh b, Bogdan Z. Dlugogorski c, *
a
United Arab Emirates University, Department of Chemical and Petroleum Engineering, Al-Ain, 15551, United Arab Emirates
b
University of Jordan, Department of Chemistry, Amman, 11942, Jordan
c
Charles Darwin University, Energy and Resources Institute, Darwin, NT, 0909, Australia

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• The study presents a kinetic model on


thermal decomposition of a PFOA model
compound.
• Major initiation routes include subse­
quent elimination of HF/CO and direct
ejection of HF/CO2.
• Profiles of decomposition products
illustrate formation of several CnFm
species.
• Perfluorinated cyclic compounds arise
from ring-closure of singlet-diradical
structures.

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

Handling editor: Derek Muir Thermal decomposition of high-fluorine content PFAS streams for the disposal of old generations of concentrates
of firefighting foams, exhausted ion-exchanged resins and granular activated carbon, constitutes the preferred
Keywords: method for destruction of these materials. This contribution studies the thermal transformation of per­
Fluorotelomer carboxylic acids fluoropentanoic acid (C4F9C(O)OH, PFPA), as a model PFAS species, in gas-phase reactions over broad ranges of
Reaction mechanisms
temperature and residence time, which characterise incinerators and cement kilns. Our focus is only on gas-phase
Pyrolysis
reactions, to formulate a gas-phase submodel that, in future, could be used in comprehensive simulation of
thermal destruction of PFAS; such comprehensive models will need to comprise fluorine mineralisation on flyash
and in clinker material. Our submodel consists of 56 reactions and 45 species, and includes new pathways that
cover the initial decomposition channels of PFPA, including those that lead to the formation of the n-C4F9 radical,
the abstraction of hydroxyl H by O/H radicals, the fragmentation of the n-C4F9 radical, reactions between HF and
perfluoropentanoic acid, as well as between HF and heptafluorobutanoyl fluoride (C3F7COF), and the cyclisation
reactions. The model illustrates the formation of a wide spectrum of small CnFm and CnHFm compounds in the
temperature window of 800–1500 K, 2 and 25 s residence time in a plug flow reactor, providing theoretical
estimates for the operating conditions of PFAS thermal destruction systems. The initiation reactions involve the
loss of HF and formation of the transition α-lactone species that converts to C3F7COF, with C4F9C(O)OH
completely decomposed at 1020 K for 2 s residence time. At 1500 K, we predict the emission of ꞉CF2 (biradical
difluorocarbene), HF, CO2, CO, CF4, C2F6, and C2F4, but at < 1400 K, we note the formation of 1H-nonafluoro­
butane (C4HF9), phosgene (COF2), and heptafluorobutanoyl fluoride (C3F7COF), with 1-C4F8, 2-C4F8 and C3HF7
persisting to 1500 K. We demonstrate that, the gas-phase pyrolysis processes by themselves convert PFAS to HF

* Corresponding authors.
E-mail addresses: mn.altarawneh@uaeu.ac.ae (M. Altarawneh), bogdan.dlugogorski@cdu.edu.au (B.Z. Dlugogorski).

https://doi.org/10.1016/j.chemosphere.2021.131685
Received 27 March 2021; Received in revised form 21 July 2021; Accepted 25 July 2021
Available online 29 July 2021
0045-6535/© 2021 Elsevier Ltd. All rights reserved.
M. Altarawneh et al. Chemosphere 286 (2022) 131685

and short-chain fluorocarbons, with similar product distribution for short (2 s) and long (25 s) residence times, as
long as the treatment temperature exceeds 1500 K. These residence times reflect those encountered in in­
cinerators and cement kilns, respectively. Thermokinetic and mechanistic insights revealed herein shall assist to
innovate PFAS thermal disposal technologies, and, from a fundamental perspective, to accelerate research
progress in modelling of gas/solid reactions that mineralise PFAS-derived fluorine.

1. Introduction Apparently, PFOA was used as a synthesis aid in the production of flu­
orotelomer surfactants. Thus, destruction of concentrates of firefighting
The most common scenarios of remediation of PFAS (per and poly­ foams involves treatment of significant amounts of PFOA and its ho­
fluoroalkyl substances) include the destruction of concentrates of PFAS- mologues, making one of these species a convenient molecule to
based firefighting foams and sewage sludge as well as ion-exchange investigate in mechanistic studies, as we do in the present work. The
resins (IER), granulated activated carbon (GAC) and similar materials same reasoning applies to the thermal disposal of IER and GAC that may
loaded with adsorbed PFAS extracted from contaminated water (Garg contain high loads of PFOA, say from the clean-up of ground water
et al., 2021; Horst et al., 2021; Winchell et al., 2021). The destruction contaminated with products of decomposition of ammonium per­
technologies used for this purpose must be versatile and non-specific to fluorooctanoate used in manufacturing of fluoropolymers (Schroeder
decompose a large variety of PFAS chemicals and mineralise their et al., 2021).
fluorine content, such as in CaF2 or as F‾ in aqueous phase (Wang et al., The total organic fluorine (TOF) of the most common concentrates of
2015; Hori et al., 2021; Riedel et al., 2021). For example, the so-called defence-grade Light Water corresponded to between 1.8 and 2.1%. This
Light Water formulations of firefighting foams, produced by the 3M can be compared with 0.8–1.2% of TOF present in the concentrates
Company until 2002, accounted for 55% of the total amount of the based on fluorotelomer surfactants, produced by the competitors of the
stored concentrates in 2004 in the USA (Darwin, 2004). Light Water 3M Company for the military market (Snow et al., 2017). Concentrates
typically contained two classes of fluorosurfactants: (i) filmers that in of PFAS foams prepared for the civilian applications, to pass UL 162 and
post-1970 formulations consisted mainly of potassium PFOS (per­ EN 1568:3 standards, typically comprise around 0.3–0.4% TOF (Dubocq
fluorooctane sulfonate) added to the so-called 3% concentrates at levels et al., 2020; Dlugogorski and Schaefer, 2021). Treatment processes to
of between 1 and 1.84 %, as active matter; and, (ii) foamers, based on handle concentrates of firefighting foams must, therefore, be capable of
the two species illustrated in Scheme 1, at levels of between 2.56 and destruction and mineralisation of the very high abundance of organic
2.74% (active matter) (Alm and Stern, 1992). fluorine in these materials. At present, only the combustion-based
In practice, the formulations produced by the 3M Company destruction processes, especially cement kilns, can effectively reme­
comprised many impurities and byproduct chemicals because of the diate such concentrated streams of PFAS (Holmes and Klein, 2020;
non-selective operation of the Simons electrofluorination cell and the Winchell et al., 2021). Cement kilns provide long gas residence time that
subsequent synthesis steps that included no purification of target species typically exceed 10 s at temperatures above 1700 K, with clinker min­
(Dlugogorski and Schaefer, 2021). Hundreds of fluorosurfactant species eralising the fluorine (Meystre and Silva, 2015; Holmes and Klein,
detected both in the environment and in the concentrates represent the 2020). The latter authors also reported the destruction and removal
legacy of the Light Water concentrates (Backe et al., 2013; Barzen-­ efficiency of PFOS, PFOA and PFHxS (perfluorohexanesulfonic aicd/­
Hanson et al., 2017). One such group of byproducts, from the Simons sulfonate) in cement kilns to be better than 99.9993%. In comparison,
cell and the subsequent hydrolysis, comprises potassium PFOA (per­ incinerators operate under a range of conditions that depend on the
fluorooctanoate) and its homologues. (In this article, PFOA denotes both application and technology, with multiple hearth furnaces and fluidised
perfluorooctanoate and perfluorooctanoic acid and, similarly, for other bed furnaces used in sewage sludge incineration, and moving grate,
acidic PFAS.) While not added on purpose to the post-1970 formulations rotary kiln and liquid injection incinerators more common for handling
of Light Water, these species represented about 2% of the entire PFAS municipal and hazardous waste (Winchell et al., 2021). Except for rotary
load in the concentrates (Dlugogorski and Schaefer, 2021). PFOA also kilns and liquid injection, these systems do not operate at temperatures
appeared in fluorotelomer-based concentrates of the pre-2015 fire­ higher than 1300 K, typically below 1150 K, with the residence time of 2
fighting foams, with fluorotelomer technology used by the competitors s for moving grade and several seconds for multiple hearts and fluidised
of the 3M Company. The fluorotelomer based formulations accounted bed installations (Winchell et al., 2021). This comparison of the vari­
for the remaining 45% of the stored concentrates (Darwin, 2004). ability in the operating conditions of incinerators has prompted us to

Scheme 1. Examples of foamers used by the 3M Company in its formulations of PFOS foams (Dlugogorski and Schaefer, 2021); Light Water contained hundreds of
congeners of these chemicals.

2
M. Altarawneh et al. Chemosphere 286 (2022) 131685

develop the first comprehensive submodel of the gas-phase reactions decomposition of the perfluorobutanoic acid, similarly to the formation
that occur in incinerators to provide the theoretical estimates of PFAS of a transitional species of α-sultone that may arise in the pyrolysis of
destruction and mineralisation as function of the operating conditions. PFOS (Altarawneh, 2021; Weber et al., 2021), but is too unstable to be
While, in this work, we execute the submodel under the pyrolysis con­ observed experimentally. Computed reaction rate constants established
ditions, the submodel can accept concentrations of moisture and oxygen the fate of the initially-formed α-lactone intermediate to be dominated
typically encountered in practical systems (Taylor et al., 2014). by two temperature-dependent competing channels affording penta­
Detailed mechanistic studies on degradation of PFAS compounds fluoropropanoyl fluoride (and CO), and a closed-shell singlet CF3CF2CF꞉
have generally focused on catalysed reactions in which calcium hy­ species (and CO2). A significant hyperconjugation across the carbon
droxide (Wang et al., 2015), alumina (Wang et al., 2019) and hematite alkyl chain, induced by fluorine atoms elevated the barrier height for a
(Gao and Chorover, 2012) effectively destroy the pollutants at temper­ channel directly eliminating CO2 to produce 1H-perfluorobutane
atures as low as 670 K. The merits of the catalyst-assisted process are the (C4HF9, 1H-nonafluorobutane).
subsequent capture of HF, leading to fluorination of the metal oxide, and Our previous study has only focused on the initial decomposition
the release of water. Unfortunately, at this temperature, the capture pathways without attempting to construct a comprehensive kinetic
operates only at about 80% efficiency, with the thermal treatment model that considers channels leading to smaller CnFm-type fragments
emitting short-chair fluorocarbons (Wang et al., 2015). Measured XRD from subsequent decomposition reactions. Crucially, the previous DFT
patterns confirm the generation of CaF2 phases upon reaction of PFOS study (Altarawneh, 2012) did not include important parallel bimolec­
with Ca(OH)2 across the temperature window of 570–1170 K (Wang ular reactions involving HF and the H/O radical pool. These reactions
et al., 2015). Conversion of metal oxides into their metal fluorides was significantly contribute to the PFAS decomposition process, especially at
suggested to take place via the interaction of the metal oxide (or hy­ elevated temperatures. Likewise, the recent detection of novel per­
droxide) with the degradation fragments of PFAS rather than with the fluorinated cyclic compounds (Singh et al., 2019) warrants mapping out
PFAS molecule itself; namely, with a wide array of CnFm radicals. relevant reaction pathways that were not addressed in previous in­
Depending on the operational temperature, partial or full mineralisation vestigations. In a recent study (Altarawneh, 2021), we investigated
of fluorine follows the initial decomposition of the parent PFAS molecule pathways that operate in the thermal decomposition of
(Wang et al., 2012, 2015, 2019). perfluoro-1-butanesulfonic acid (C4F9SO2OH), as a model compound of
Along the same line of enquiry, other investigations addressed the PFOS. Our kinetic modelling (Altarawneh, 2021) demonstrated SO2 as
thermal decomposition of PFOA. A recent study by Xiao et al. (2020) the major sulfur species with a noticeable contribution by SO3. The
investigated thermal decomposition of different PFAS compounds with a investigated reaction pathways included hydrofluorination, hydrolysis,
varying chain length of the CF2 groups over spent granular activated and fragmentation of the alkyl chain.
carbon (GAC). The mass spectra revealed both the sublimation of PFOA To this end, this study revisits reactions operating in the initial
from GAC and its partial decomposition at temperatures as low as 473 K. thermal decomposition of perfluoropentanoic acid (as a model com­
Stockenhuber et al. (2019) reported the formation of pound for PFOA and, in general, for PFAS) in light of the recent exper­
perfluoro-1-heptene (1-C7F14), 1H-perfluorohexane (C6HF13, 1H-tride­ imental findings. As reported by Khan et al. (2020) for PFOS, the length
cafluorohexane) and 1H-perfluoroheptane (C7HF15, 1H-pentadecafluor­ of the perfluoroalkyl chain exerts a rather minor influence on the re­
oheptane) from the thermal degradation of PFOA under an inert N2 action and activation energies for the initial decomposition channels.
atmosphere in a gas-phase process between 300 and 750 ◦ C for long Similarly, the findings of Bentel et al. (2019), albeit gained in the context
residence time of up to 46 s. The FT-IR spectra highlighted the formation of PFAS reduction by solvated electrons, demonstrated very similar
of HF, CO, CO2. Production of HF was shown to be sensitive to the decomposition trends among perfluoroalkyl carboxylic acids (n = 2-10;
applied residence time, in which longer residence time suppressed the where n denotes the number of CF3 and CF2 segments that form a single
generation of HF, presumably due to its consumption in the hydro­ perfluoroalkyl chain). These authors also described defluorination effi­
fluorination reaction: HF + alkene → fluoroalkane. The temperature for ciency of PFCA (perfluoroalkyl carboxylic acids) in a UV-irradiated
the onset of the decomposition of PFOA, reported by Stockenhuber et al. environment to reside in the very narrow range of 49.1-59.5%
(2019), fall below the expected temperature for the gas-phase initiation (n=2-10). As such, we anticipate that the kinetic model of the thermal
reaction. This indicates promotion of the decomposition process on the decomposition of the perfluoropentanoic acid (n = 4) will represent well
walls of their reactor. Krusic et al. (2005) studied the kinetics of the the decomposition behaviour of longer-chain PFCA (e.g., n = 7). From
thermolysis of PFOA in quartz ampoules, reporting 1H-perfluoroheptane these perspectives, the current investigation has four aims: i) to
and perfluoro-1-heptene as major and minor products, respectively. construct a kinetic model that describes the decay of PFPA and forma­
Adding crushed quartz significantly accelerated the decomposition of tion of most important products reported in experiments; ii) to elucidate
PFOA, indicating a strong catalytic effect, although the nature of this the role of the essential bimolecular reactions induced by the abundant
effect was different for quartz and borosilicate surfaces. Finally, within a HF species that form in the initial stages of the degradation of PFPA; iii)
different context of the thermal degradation of PFOA in aqueous solu­ to address the reactions that contribute to the formation of cyclic per­
tions, as function of temperature and pH, Liu et al. (2017) also identified fluorinated species; and, iv) to establish the theoretical (modelled) re­
perfluoro-1-heptene, as the only species that formed under their quirements for destruction and mineralisation of PFAS in gas reactions
experimental conditions. In summary, all experimental results on the under conditions corresponding to those encountered in incinerators. It
decomposition of PFOA suggest that the process starts with the emission is expected that, the mechanism developed in this contribution will
of HF from the head group and leads to the initial formation of 1H-per­ provide much improved and much more detailed understanding of the
fluoroheptane and perfluoro-1-heptene (Krusic et al., 2005; Liu et al., transformation chemistry of PFAS compounds under a pyrolytic envi­
2017; Stockenhuber et al., 2019; Xiao et al., 2020). However, the ronment, including profiles of degradation products of PFAS species,
mechanistic details of the process remain unknown, and there is no with direct application to modelling of the remediation of these con­
comprehensive kinetic model to predict the yields of the decomposition taminants in practical thermal processes.
products.
In a previous theoretical density functional theory (DFT) study, we 2. Computational detail
illustrated reaction and activation energies for pathways operating in
the initial decomposition of perfluorobutanoic acid (C3F7C(O)OH), as a Gaussian 09 (Frisch et al., 2009) suite of programs afforded all
representative model compound for the degradation of PFOA (Altar­ structural optimisations and energy calculations based on the M062X
awneh, 2012). Among several opening channels, we found the elimi­ DFT method at the polarised basis set of 6-311 + G(d,p) (Montgomery
nation of HF leading to a α-lactone intermediate to dominate the initial et al., 2000). Given the size of the C7F15C(O)OH molecule, we have

3
M. Altarawneh et al. Chemosphere 286 (2022) 131685

found it computationally prohibitive to perform all calculations for the Table 1


PFOA molecule itself, and hence selected the perfluoropentanoic acid as Modified Arrhenius parameters fitted in the temperature range of 500–2000 K.
a model compound. Nonetheless, we report trends in bond dissociation Several of these reactions also appear in our presented kinetic model on the
enthalpies and activation enthalpies for several classes of reactions pyrolysis of a PFOS model compound (Altarawneh, 2021).
involving a varying length of perfluorinated alkyl chain of between 2 Reaction A (s− 1, cm3, n Ea Method
and 7. The meta-hybrid M062X functional performs well in attaining molecule) (kcal/
mol)
thermokinetic values for a broad spectrum of applications in organic
chemistry (Zhao and Truhlar, 2008). Our quantum chemical studies on 1 M0 (Perfluoropentanoic 1.39 × 1012 0.40 61.0 TST
halogenated (i.e., chlorinated/brominated) compounds established a acid) → M1 + HF
9
2 M0 → CO2 + HF + M6 3.59 × 10 0.97 68.8 TST
satisfactory performance of the adopted methodology in computing (Perfluoro-1-butene)
thermochemical and kinetic parameters (Altarawneh et al., 2009; 3 M0 → CO2 + M5 (1H- 6.27 × 1011 0.82 71.2 VTST
Altarawneh and Dlugogorski, 2014a, b; Altarawneh et al., 2019; Lee perfluorobutane)
et al., 2020). Likewise, Table S1 in the Supplementary Material (SM) 4 M0 → H + M4 1.29 × 1013 0.96 89.0 TST
5 M0 → COOH + M7 (n-C4F9) 1.30 × 1013 0.65 86.5 VTST
contrasts computed reaction enthalpies for a few reactions involving
6 M0 + H → H2 + M4 1.87 × 1013 0.0 16.4 TST
fluorinated species with analogous experimental values, calculated 7 M4 → M7 (n-C4F9) + CO2 1.00 × 1013 0.0 0.0 TST
based on standard enthalpies of formation (ΔfHo298.) of involved species 8 M1 → CO2 + M3 (n-C4F8) 3.11 × 1012 0.59 34.0 TST
(Afeefy et al., 2005). The average of deviations between calculated and 9 M3 → M6 1.59 × 109 1.48 43.9 TST
experimentally-based values only amounts to 1.9 kcal/mol. In their 10 M1 → CO + M2 6.88 × 1011 0.49 19.0 TST
(Heptafluorobutanoyl
study on the thermal decomposition of PFOS, Khan et al. (2020) found
fluoride)
that, values calculated by the M062X method generally reproduce 11 M5 → M6 + HF 6.18 × 109 1.2 89.0 TST
analogous figures obtained by the chemistry model G3X. Stability of 12 M7 → M6 + F 1.22 × 1012 0.54 58.3 VTST
wavefunctions of singlet-diradical species was verified. 13 M7 → CF2 + M8 (n-C3F7) 7.47 × 1013 0.86 53.7 VTST
14 M8 → F + M9 1.22 × 1012 0.54 58.3 VTST
Calculations of the intrinsic reaction coordinate (IRC) (Hratchian
15 C2F4 + C2F4 ⇆ cyc-C4F8 5.0 × 1011 0.0 23.0 TST
and Schlegel, 2005) confirmed the atomic connection between transi­ 16 C2F5 → CF3 + CF2 4.13 × 1015 0.0 60.0 TST
tion states and their associated reactants and products. We computed all 18 M7 → C2F4 + C2F5 5.10 × 1011 0.9 42.6 TST
reaction and activation enthalpies at 298.15 K. The KiSThelP package 19 M8 ⇆ C2F4 + CF3 8.66 × 1011 0.9 43.9 TST
(Canneaux et al., 2014) served to obtain the reaction rate constants 20 M17 → M9 + HF 6.18 × 109 1.2 89.0 TST
21 M0 + HF → M10 3.00 × 104 5.13 13.2 TST
based on the conventional transition state (TST). Modified Arrhenius
22 M10 → M12 + HF 3.10 × 1011 0.71 80.2 TST
parameters, k(T) = A Tn e− (Ea/(RT)), were fitted between 500 and 2000 K. 23 M10 → M6 + HF + FC(O)OH 3.17 × 1010 0.73 75.2 TST
Plausible contributions from quantum tunnelling effects on k(T) values 24 M10 → M5 + FC(O)OH 5.70 × 1012 0.40 80.9 TST
were accounted for via the inclusion of a one-dimensional Eckart bar­ 25 M2 + HF → M14 1.56 × 1014 0.0 33.6 TST
26 M14 → M9 3.61 × 1010 0.82 78.3 TST
rier. In order to set a benchmark of accuracy for the computed reaction
(Perfluoropropene) + HF +
rate constants, Figure S1 (SM) portrays a close agreement, in terms of COF2
the comparison of Arrhenius plots, between our results and the corre­ 27 M14 → M17 (1H- 9.19 × 1013 0.90 88.2 TST
sponding literature values (Ainagos, 1991), for the important dissocia­ perfluoropropane) + COF2
tion reaction n-C4F9 → C2F4 + C2F5. The two sets of values are within a 28 M12 → M13 + HF 9.54 × 1010 0.66 55.4 TST
29 M2 + HF → M14 8.56 × 1014 0.00 33.6 TST
factor of 2–5 between 800 and 1200 K. Reaction rate constants for a few
30 C2F4 + HF → C2F5H 6.98 × 1010 0.00 48.0 TST
barrierless reactions were calculated based on the variational transition 31 M2 → COF + M8 5.00 × 1012 0.00 87.0 VTST
state theory (VTST) (Garrett and Truhlar, 2005). This method minimises 32 CHF3 → CF2 + HF 3.95 × 1010 1.11 76.9 TST
the rate constants as function of temperature and reaction coordinates (i.
e., bond stretching). Minimum energy points required in the VTST cal­
culations were attained by varying the relevant C–C and C–F bonds by an 3. Results and discussion
increment of 0.1 Å, until the structure no longer retains a single imag­
inary frequency along the designated reaction coordinate. Table 1 as­ Herein, we update the previously suggested mechanism (Altar­
sembles all estimated constants for the reaction rates introduced in this awneh, 2012) for the decomposition of perfluorobutanoic acid, consid­
study or by (Altarawneh, 2021). ering four new sets of important reactions: namely, (i) pathways leading
The CHEMKIN-RRO software (Reaction Design, 2013) performed to formation of a n-C4F9 radical and its further decomposition into
kinetic modelling based on a plug-flow reactor model (PFR); thereby, smaller CnFm fragments; (ii) the abstraction of the hydroxyl H by O/H
providing simultaneous solution to the governing ordinary differential radicals; (iii) bimolecular reactions involving HF molecules; and, (iv)
equations that describe the change of species concentrations as function cyclisation reactions. We commence the discussion by considering the
of the reactor coordinate. Section 3.6 describes details of the modelled initial decomposition channels of C4F9C(O)OH, including those that lead
reactor, flow conditions, temperature and residence time. While our to the formation of the n-C4F9 radical. Then we proceed to cover the
submodel can admit atmospheres that are typical of incinerators, here, abstraction of hydroxyl H by O/H radicals, before addressing the frag­
we include the results of pyrolysis of PFPA, as a model PFAS species, in mentation of the n-C4F9 radical. We then cover the reactions between HF
N2 bath gas. NASA polynomials were obtained with the aid of the and perfluoropentanoic acid, as well as between HF and hepta­
CHEMRATE software (Mokrushin et al., 2002) using the computed fluorobutanoyl fluoride, and then describe the cyclisation reactions.
vibrational frequencies, rotational constants, and values of ΔfHo298.
Literature provides no account on the value of ΔfHo298 for the per­ 3.1. Initial decomposition of the perfluoropentanoic acid
fluoropentanoic acid. Thus, we computed ΔfHo298 from the isodesmic
reaction given in Scheme S1 of the SM. Other ΔfHo298 values, for initially Our previous DFT study (Altarawneh, 2012) presents ten plausible
formed perfluorinated products and intermediates, followed from the initiation channels operating in the thermal decomposition of the per­
figure of ΔfHo298 for the perfluoropentanoic acid (− 533.2 kcal/mol) and fluorobutanoic acid. Direct dissociation of the C–C bonds at α, β and γ
the estimated enthalpies of the relevant reactions. NASA polynomials for locations away from the carboxylic group and the fission of the hydroxyl
all species are provided in the SM. O–H bond proceed via sizable endothermic reactions of 87.5, 85.3, 93.7
and 89.0 kcal/mol, respectively. The CO2 expulsion reactions require
the activation enthalpy of between 77.3 and 96.2 kcal/mol. For this

4
M. Altarawneh et al. Chemosphere 286 (2022) 131685

reason, these reactions contribute rather minimally to the initial the O–H bond in the hydroxyl group (M0 → M4 + H) and the α C–C bond
degradation of PFPA and its homologues. By examining the relevant (M0 → M7 + C(O)OH). The former is endothermic by 111.8 kcal/mol
activation enthalpies, the HF elimination, requiring 54.2 and 35.1 while the latter by 86.2 kcal/mol. Our calculated bond dissociation free
kcal/mol of activation and reaction enthalpies, in this order, and leading energy for the fission of the C–C(O)OH bond (90.5 kcal/mol) clearly
to the formation of a α-lactone intermediate, appears to be far more overshoots a corresponding value of 63.9 kcal/mol calculated by Liu
important than the other reactions. Further decomposition of the et al. (2017) at the B3LYP/6-31G(d) level of theory. It is well-established
α-lactone intermediate branches into two channels forming penta­ that, the commonly deployed B3LYP functional significantly un­
fluoropropanoyl fluoride molecule and a singlet CF3CF2CF꞉ species. derestimates reaction enthalpies when compared with meta-hybrid DFT
While literature reports no detection of the perfluoroalkyl α-lactones methods (Zhao and Truhlar, 2008), deployed in the present study. The
and perfluoroalkylacyl fluorides as intermediates in the thermal scission of the O–H bond most likely proceeds via bimolecular reactions
decomposition of PFCA (but note that, Weber et al. have recently proved with H/OH/F radicals in the pyrolytic andoxidative environments.
the formation of pentafluoropropanoyl fluoride in the pyrolysis of
PFOS), the channels similar to those shown in Scheme 2 for PFPA 3.2. Abstraction of the hydroxyl by O/H radicals
dominate the initial decomposition of fluoroacetic (Blake and Tomlin­
son, 1971) and trifluoroacetic (David and Philip, 1997) acids. Fig. 2 shows reaction and activation enthalpies for abstraction of the
Fig. 1 depicts the augmented reactions pathways that prevail at the hydroxyl H atom by H/OH/F radicals, providing estimates for the facile
initial stages of the degradation of the PFPA, M0 (to simplify the nota­ barriers. Despite the best efforts, our attempts were unsuccessful in
tion, chemical species are termed as Mx in the subsequent discussion). finding an activation enthalpy for the thermally-neutral H abstraction by
The computed enthalpies, presented in the lowermost part of Fig. 1, F atoms. Likewise, literature values of the kinetic rate constants for the
depart only by small values from their analogous estimates calculated abstraction of the hydroxyl H by F atoms often feature a temperature-
using the G3MP2B3 composite chemistry model and reported in our independent behaviour; i.e., the reaction does not encounter a genuine
previous study (Altarawneh, 2012). The direct formation of 1H-perfluor­ reaction barrier (Assaf et al., 2018). In our kinetic modelling, we
obutane (M5) and perfluoro-1-butene (M6, 1-C4F8), which are analo­ adapted rate parameters for the reaction F + methanol → HF + methoxy
gous to the experimentally observed 1H-perfluoroheptane and (Dóbé et al., 1994) to account for the H abstraction from the per­
perfluoro-1-heptene from the decomposition of PFOA (Krusic et al., fluoropentanoic acid by fluorine atoms.
2005; Stockenhuber et al., 2019), proceeds via sizable activation en­ A decarboxylation reaction along the channel M4 → M7 + CO2
thalpies of 72.4 kcal/mol and 69.3 kcal/mol, respectively. The calcu­ (CF3CF2CF2CF2., n-C4F9) releases 5.3 kcal/mol of enthalpy without
lated barrier for the elimination of a CO2 molecule along the reaction M0 climbing a reaction barrier. When the decomposition process of per­
→ M5 + CO2 matches the corresponding barrier encountered in the fluorinated carboxylic acids proceeds in an environment with estab­
decarboxylation of acetic acid into methane and CO2 of 71.2 kcal/mol lished H/O/F radical pools, long-chain fluoroalkyl radicals constitute
(Nguyen et al., 1995). As suggested by Stockenhuber et al. (2019), direct the prominent initial intermediates. For example, such scenarios prevail
formation of perfluoro-1-heptene (along with HF and CO2) may occur in recent experiments that investigated the decomposition of PFOA to
through a six-centred transition state. Our calculated enthalpic barrier radical species in non-equilibrium plasma that included hot and solvated
for this new pathway amounts to 69.3 kcal/mol through the transition electrons (Singh et al., 2019). Thus, it is of interest to address in details
structure TS3. The high barrier of TS3 arises because of the strength of the subsequent decomposition pathways of long-chain fluoroalkyl rad­
secondary C–F bonds (115.4 kcal/mol). Scheme 3 portrays the icals, also for applications in non-thermal systems.
geometrical features of TS3.
The channel of M0 → M6 + HF + CO2 is endothermic by 13.3 kcal/
3.3. Fragmentation of the n-C4F9 radical
mol. Nonetheless, elimination of HF in the reaction M0 → M1 + HF
remains the leading channel with an activation enthalpy of 60.7 kcal/
The formation of the major products of 1H-perfluoroheptane and
mol. The formed α-lactone intermediate (M1) is unstable because it only
perfluoro-1-heptene from the pyrolysis of PFOA (Krusic et al., 2005)
needs to climb a facile barrier of 18.6 kcal/mol to dissociate into hep­
may take place via a direct loss of a fluorine atom from the n-C7F15
tafluorobutanoyl fluoride (M2) and CO (TS4). In a competing minor
radical (producing the perfluoro-1-heptene molecule) and by HF addi­
pathway, the diradical species CF3CF2CF2CF꞉꞉ (M3) forms through an
tion (producing C4HF9). We envisage that fluorine mineralisation (i.e.,
enthalpic barrier of 33.4 kcal/mol (TS5). A 1,2-F shift transforms M3
capturing the fluorine content of PFAS by Ca(OH)2 (Wang et al., 2015),
into the thermodynamically more stable perfluoro-1-butene (M6)
or as F− in aqueous solutions (Hori et al., 2021), is not limited to the
molecule via an accessible enthalpic barrier of 44.4 kcal/mol.
parent PFAS species, but could also encompass active CnFm radicals and
The two considered bond fission channels characterise the rupture of
intermediates that form from the degradation of PFAS compounds. Thus,

Scheme 2. Previously presented initialisation channels.

5
M. Altarawneh et al. Chemosphere 286 (2022) 131685

Fig. 1. Reaction pathways operating in the unimolecular decomposition of perfluoropentanoic acid (M0). Values in bold and italic fonts denote reaction and
activation enthalpies, respectively. All values were calculated at 298.15 K and are presented kcal/mol.

As shown in Fig. 1, the n-C4F9 radical (M7) branches into three


plausible channels. Expulsion of a difluorocarbene (꞉CF2) molecule takes
place in a barrierless endothermic reaction of 55.2 kcal/mol. This route
affords the n-CF3CF2CF2 radical (M8). Unlike methylidene (꞉꞉CH2, dihy­
dridocarbon), the singlet-state of the ꞉ CF2 molecule is more stable than
its triplet configuration. Our calculated energy gap between the triplet
and singlet difluorocarbenes corresponds to 52.5 kcal/mol, in a close
agreement with the experimental figure of 55.6 kcal/mol of Brahms and
Dailey (1996). In a more energy demanding pathway, departure of a
fluorine atom to produce perfluoro-1-butene is noticeably endothermic
by 58.6 kcal/mol. Nonetheless, this value is significantly lower than
bond dissociation enthalpies for secondary C–F in alkanes (115.4
kcal/mol).
The most plausible channel signifies fragmentation of M7 into C2F4
and C2F5. through an accessible reaction enthalpy of 40.1 kcal/mol
(TS22). Fig. 3 plots the branching ratios for two most probable exit
Scheme 3. Structure of TS. channels of the n-C4F9 radical (M7), based on their reaction rate con­
stants listed in Table 1. Higher entropy of activation for the channel ꞉CF2
in a future contribution, it will be insightful to theoretically address the + n-C3F7. renders it to be competitive with the degradation reaction
sequential conversion of potent metal hydroxides into their fluorides C2F4 + n-C2F5. across the entire temperature window explored in Fig. 3,
through dissociative adsorption of PFAS compounds, and their derived albeit with the change in the dominant reaction that occurs around
products. 1200 K. Thus, both channels assume comparable importance during the

Fig. 2. Reaction (in bold) and activation (in italic) enthalpies for H abstraction from the hydroxyl H in the perfluoropentanoic (M0) molecule by H and OH radicals.
All values are in kcal/mol at 298.15 K.

6
M. Altarawneh et al. Chemosphere 286 (2022) 131685

hydrofluorination of the terminal carbonyl groups (C– – O) in carboxylic


acids, –C(O)OH, and acyl fluorides, –COF.
Very recently Altarawneh (2021) investigated a series of hydrolysis
and hydrofluorination reactions involving the transformation of hepta­
fluorobutanoyl fluoride and subsequently-formed molecules. Kinetics of
these reactions were discussed to establish the competitive nature of
three pathways, namely, HF addition, hydrolysis, and C–C bond fission.
Herein, we limit our analysis to HF addition reactions to simulate the
effect of purely pyrolytic environment. We emphasise the importance of
these reactions, HF + perfluoropentanoic acid/heptafluorobutanoyl
fluoride, for the formation of experimentally observed products from the
decomposition of PFOA. Fig. 4 presents the potential energy surfaces
(PES) of the bimolecular reactions of HF with PFPA and C3F7COF. In
Fig. 4a, the fission of the H–F bond across the aldehydic C–O bond in the
M0 structure demands a low barrier of only 14.1 kcal/mol. This reaction
results in the formation of the double hydroxylated adduct of M10 that
exists in a well-depth of 25.0 kcal/mol, in reference to the initially
separated reactants.
The M10 intermediate branches into three channels, all with sizable
enthalpic barriers in the range of 75.9–81.7 kcal/mol. The transition
TS11 signifies the formation of carbonofluoridic acid (FC(O)OH) and the
Fig. 3. Branching ratios for the two competing channels in the unimolecular 1H-perfluorobutane molecule. An unimolecular elimination of HF from
decomposition of the n-C4F9 radical. C4HF9 generates 1-C4F8 through a high enthalpic barrier of 90.2 kcal/
mol in a reaction that is endothermic by 26.6 kcal/mol. In a parallel
decomposition of PFAS compounds with long perfluoroalkyl chains. pathway, the six-centred transition state TS12 marks the direct forma­
Next, we turn our attention to elucidate the effect of the length of the tion of 1-C4F8, HF, and FC(O)OH molecules. In the third pathway, for­
perfluoroalkyl chain on the energy requirements for the expulsion of mation of the epoxy intermediate M12 takes place via TS13 (80.0 kcal/
꞉CF2 and C2F4 molecules. Figure S2a (SM) depicts enthalpic values for mol). The relative importance of the pathways shown in Fig. 4a rests on
the sequential loss of CF2 groups, commencing from a n-C7F15 radical. a competition between the reverse reaction M10 → M0 + HF and frag­
The five considered C–C fissions in Figure S2 essentially exhibit similar mentation of M10 via the three pathways. The enthalpic barriers for
bond dissociation enthalpies in the very narrow range of 54.9–55.9 kcal/ these pathways (75.9–81.7 kcal/mol) significantly surpass that of M10
mol. Similarly, Figure S2b illustrates the loss of C2F4 molecules from → M0 + HF (39.2 kcal/mol). The shallow enthalpic barrier for the HF
perfluorinated n-C4-7F7-15 radicals arising via enthalpy barriers of removal from the M10 intermediate diminishes the importance of the
38.2–40.8 kcal/mol. However, the very shallow reverse barriers (1–3 HF-assisted decomposition pathways depicted in Fig. 4a. Overall, the
kcal/mol) plausibly diminish the overall contribution of these reactions length of the carbon chain (4 in this study versus 5 in (Altarawneh,
in a practical pyrolytic environment. Nonetheless, fission of the 2021) exerts a minor influence on values of the activation energies
F3C–C2F4 bond requires a slightly higher barrier of 44.2 kcal/mol. In a pertinent to HF addition.
nutshell, we infer that, the length of the perfluorinated alkyl n-CnFm Fig. 4b illustrates the dissociation of the HF molecule over the C– –O
radicals induces a rather minor influence on energies pertinent to their bond in the heptafluorobutanoyl fluoride via an accessible reaction
gas-phase fragmentation, i.e., via the ꞉CF2 and C2F4 routes. We note the enthalpy of 34.9 kcal/mol, generating the M14 adduct in a noticeably
recent B3LYP/6-311 + G(2 d, 2p) calculations of Bentel at al (Bentel exothermic reaction of − 132 kcal/mol. Decomposition of the M14 in­
et al., 2019) that reveal a somehow noticeable variations (within 3 termediate proceeds via three pathways similar to those operating in the
kcal/mol) in bond dissociation enthalpies of C–F in ionic fragmentation of M10. The three corridors produce perfluoropropene,
CF3(CF2)n-COO- and CF3(CF2)n-SO2O− species. With a very strong bond 1H-perfluoropropane, carbonyl difluoride (COF2), in addition to the
dissociation enthalpy (~110 kcal/mol), fission of secondary C–F bonds epoxy intermediate M16. The enthalpic barrier for the back dissociation
is unlikely to contribute significantly in the homogenous gas phase corresponds to 66.4 kcal/mol, lower than the barrier for transition state
decomposition of PFCA. TS16 (78.6 kcal/mol). However, at high temperatures both reaction
routes may be competitive affording the corridors in Fig. 4b to consume
3.4. Hydrofluorination of perfluoropentanoic acid and heptafluorobutanoyl fluoride.
heptafluorobutanoyl fluoride Preliminary kinetic modelling of pathways shown in Fig. 1 discloses
that, ꞉CF2 and HF constitute the ultimate fluorine carriers at high tem­
Hydrogen halides (HCl and HBr) are by far the most important peratures. For this reason, it is vital to include the bimolecular ꞉CF2 + HF
halogen carriers during the thermal decomposition of halogenated reaction (and subsequent steps) in the kinetic model for the thermal
compounds (Altarawneh et al., 2009, 2019). For instance, it has been degradation of PFAS. Pyrolysis of CHF3 is dominated by its fragmenta­
consistently found that approximately between 50 and 60% of the initial tion into ꞉CF2 and HF (Hidaka et al., 1991; Han et al., 2010). Several
bromine present in brominated flame retardants transfers into HBr experimental studies (mainly behind shock waves) report reaction rate
during their pyrolysis. The noticeable absence of the C6F13COF homo­ parameters for dissociation of CHF3 into ꞉CF2 and HF with apparent
logue of heptafluorobutanoyl fluoride (M2) from the products of activation energies in the range of 58.3–72.3 kcal/mol. We are unaware
degradation of PFOA as evident in the work of Stockenhuber et al. of any theoretically obtained values. Our calculated activation en­
(2019), coupled with the disappearance of H–F vibrations (i.e., thalpies for the reaction ꞉CF2 + HF ⇆ CHF3 entail values of 18.6 and
3700–4200 cm− 1), suggest a potential hydrofluorination route that 78.0 kcal/mol in the forward and reverse direction, respectively. The
consumes both molecules. It should also be noted that, Weber et al. shallow forward barrier indicates a rapid and facile consumption of ꞉CF2
(2021) detected the formation of C7F15COF from the pyrolysis of PFOS. and HF into CHF3. Fig. 5 contrasts the computed rate constants for the
Likewise, abundant HF molecules appearing at the onset of the pyrolysis backward dissociation reaction CHF3 → CF2 + HF with the analogous
of PFOA may induce the degradation of PFOA via bimolecular reactions. experimental values of Schug et al. (1979), and the recommended values
However, literature presents no account on a direct homogenous of Han et al. (2010). The agreement is rather satisfactory, especially at

7
M. Altarawneh et al. Chemosphere 286 (2022) 131685

Fig. 4. Reaction pathways operating in the bimolecular reaction of HF with perfluoropentanoic acid (a) and heptafluorobutanoyl fluoride (b). Values in bold and
italic fonts denote reaction and activation enthalpies, respectively. All values are in kcal/mol.

high temperatures relevant to the temperature window for the formation of cyc-C4F8 as one of the main products in the oxidative thermolysis of
of ꞉CF2 and HF (Han et al., 2010). fluorinated polymers, and so did many others, starting from the pio­
neering work of Lewis and Naylor (1947). The Diels-Alder cyclo-addi­
3.5. Formation of perfluorinated cyclic compounds tion of small halogenated chains, followed by successive addition of
halogen atoms at high temperature, constitutes a key step in these
Oxidation and pyrolysis of halogenated alkanes and alkenes often mechanisms (Ahubelem et al., 2015).
result in the formation of cyclic halogenated compounds. For example, While cyclic perfluorinated compounds have not yet been observed
pyrolysis of hexachloropropene (C3Cl6) above ~1200 K generated a in the pyrolysis of PFOA, several gaseous cyclic perfluorinated com­
wide array of cyclic chlorinated compounds, including C6Cl6 (hexa­ pounds (C4F8, C5F10, C6F12, C7F14, and C8F16) formed in the non-
chlorobenzene), C6Cl8 (1,3,5-octachlorohexatriene), and C8Cl8 (octa­ equilibrium plasma generated above solutions of PFAS (Singh et al.,
chlorostyrene) (Dellinger et al., 1989). DFT calculations established 2019), including that of PFOA. The mechanism proposed by Singh et al.
reaction pathways for the formation of chlorinated (Altarawneh et al., (2019) entails the generation of singlet diradical species (where the
2009) and brominated (Altarawneh et al., 2019) cyclic compounds. A radical character appears at the two ends of the carbon chain),
minor amount of cyc-C4F8 was detected during pyrolysis of n-C4F10 (Yu ring-closure of these radicals, and conversion into stable cyclic com­
et al., 2008). In comparison, Ellis et al. (2001) reported the appearance pounds via loss of a fluorine atom. However, our calculations discovered

8
M. Altarawneh et al. Chemosphere 286 (2022) 131685

noticeably increases the activation enthalpy, compared with transition


states leading to formation of cyc-C6F12, cyc-C7F14, and cyc-C8F16 (53.9
versus 43.5–44.4 kcal/mol). No analogous transition state could be
found for the formation of cyc-C4F8 from CF3(CF2)2CF꞉꞉. As shown in
Fig. 6, the cyclic perfluorinated compounds are more stable than their
parent CF2(CF2)mCF2 di-radicals.
The sequential build-up from shorter building blocks signifies the
second pathway in the formation of cyclic perfluorinated compounds. In
this regard, ꞉CF2 and C2F4 molecules represent potent precursors. Yu
et al. (2008) located a transition state with a modest barrier of only 8.9
kcal/mol for the cyclo-addition of these two moieties leading to
cyc-C3F6. This reaction has been included in our kinetic analysis. (Note
that Table 1 only includes the reactions that we compute in the present
contribution.). Likewise, we found that, the self-condensation of two
C2F4 molecules, yielding cyc-C4F8, passes via a reaction enthalpy of 25.0
kcal/mol (TS23, shown in Scheme S2) in a highly exothermic reaction of
49.6 kcal/mol. Experimental evidence for the formation of cyc-C4F8
comes from the studies of Simon and Kaminsky (1998), who found the
pyrolysis of polytetrafluoroethylene (PTFE) to produce up to 9.9 wt%
cyc-C4F8, and by Mięsowicz (1987) who reported 29–34 wt% cyc-C4F8. It
is anticipated that, further cyclo-addition of CF2/C2F4 (and their radical
Fig. 5. Comparison between computed and literature values for the dissocia­ forms) will grow the size of cyclic perfluorinated compounds. Proposed
tion reaction CHF3 → CF2 + HF. aRef (Schug et al., 1979), bRef (Han mechanisms follow the well-documented channels for the synthesis of
et al., 2010). polyaromatic hydrocarbons (PAHs) (Richter and Howard, 2000).
Finally, we are now in a position to elucidate plausible reasonings for
that, for instance, a CF3(CF2)5CF꞉꞉ diradical (a singlet diradical, where the absence of C6F13COF (the major predicted product herein at low and
the two radical characters appear on a terminal carbon atom) is more intermediates temperatures) from the degradation products of PFOA in
stable than its counterpart .CF2(CF2)5CF2. (the two carbon atoms at the Stockenhuber et al. (2019) study. Decomposition of the trideca­
two ends of the chain bear a radical character) by 72.1 kcal/mol fluoroheptanoyl fluoride is likely to commence via fission of the C–COF
(Scheme 4). bond (i.e., 87.0 kcal/mol). Thus, C6F13COF should be thermally stable at
Furthermore, we were not able to locate a transition state that results relatively high temperatures. However, the non-detection of trideca­
in the formation of cyclic CnF2n from their corresponding .CF2(CF2)2n- fluoroheptanoyl fluoride could be attributed to several factors. In
.
2CF2 diradical intermediates. The formation of the cyclic perfluorinated addition to its removal via the hydrofluorination pathway, similar to
compounds in Singh’s et al. (Singh et al., 2019) experiments indicates that shown in Fig. 4b for heptafluorobutanoyl fluoride, the catalytic
the direct cyclisation routes from corresponding unsaturated aliphatic nature of alumina surfaces deployed in Stockenhuber’s et al. experi­
chains. Fig. 6 presents the reaction and activation enthalpies for the ments may provide a surface-assisted degradation channel (Hou et al.,
cyclisation of CF3(CF2)3–5CF꞉ into perfluorinated cyclic cyc-C5F10, 2013). Aldehydic O presents a site in a molecule of C6F13COF that may
cyc-C6F12, cyc-C7F14, and cyc-C8F16 compounds. The transition states, engender the dissociative adsorption of C6F13COF over the noticeably
TS18-21, signify shifts of fluorine atoms between the two terminal car­ catalytic alumina and/or AlF3 surfaces.
bon atoms, accompanied with a ring closure. Note that, cyc-C4F8 arises More importantly, Stockenhuber et al. (2019) introduced gaseous
from two C2F4 molecules rather than from the singlet diradical, and PFOA into the reactor through slow evaporation of its powdered form. It
significant strain arises in the 4-membered ring to allow this cyclisation is not clear if a thermal decomposition in the condensed medium
to occur. occurred prior to vaporisation, especially if the vaporiser was operated
A ring strain in transition state leading to cyc-C5F10 (TS18) close to the boiling point of PFOA (462–465 K). Likewise, Xiao et al.
(2020) decomposed PFOA in a thermal desorption-pyrolysis unit but did
not mention the heating ramp; pyrolysers from CDS Analytical can be
operated between 20 K/ms and 1000 K/min, followed by a period of a
constant temperature reaction, which Xiao et al. set at 30 s for tem­
peratures higher than the boiling point of PFOA. This makes it impos­
sible to estimate the contribution of the solid-phase reactions to the
overall decomposition of PFOA. Thermal decomposition processes of
polymeric materials in the condensed phase are mainly triggered by
bimolecular reactions that lead to elimination of stable molecules (i.e.,
H2O, CO, CO2, HX), followed by C–C bridging (Starnes, 2002). As such,
these reactions may promote expulsion of CO2 molecules from the
structure of PFOA, prior to evaporation into the gas phase. For instance,
we have shown previously (Altarawneh and Altarawneh, 2012), that
major products from thermal decomposition of acetamide originate
from a series of complex bimolecular reaction that take place in the
condensed medium. Similarly, it was demonstrated (Clark et al., 2015)
that bimolecular reactions of carboxylic acids (acetic, vinylacetic
(3-butenoic acid), butanoic, crotonic ((2E)-but-2-enoic acid), iso­
crotonic ((2Z)-but-2-enoic acid)) significantly reduce activation bar­
riers, and alter the products branching ratios (decarboxylation versus
dehydration), in reference to unimolecular pathways. In summary,
Scheme 4. Thermodynamic stability of CF2(CF2)5CF2 versus CF3(CF2)5. experimental identification of C6F13COF, as a key intermediate as

9
M. Altarawneh et al. Chemosphere 286 (2022) 131685

Fig. 6. Pathways for the formation of perfluorinated cyclic compounds from ring closure of their corresponding closed-shell singlet intermediates. All enthalpy
values are in kcal/mol computed at 298.15 K.

predicted by the quantum chemical calculations, during the pyrolysis of fragmentation and bimolecular reactions of CnFm species were adopted
PFOA (and, more general, any perfluoroalkylacyl fluoride in pyrolysis of from the n-C4F10 combustion subset of Yu et al. (2008). Kinetic param­
PFCA) remains an outstanding research problem. eters for important reactions involving CHF3/CHF2 were extracted from
Hydrolysis may constitute another potent pathway for the removal of a kinetic model of Han et al. (2010) on co-pyrolysis of methane and
CnF2n+1-containing species in a reaction that sequentially produces trifluoromethane. These reactions were also included in our recent study
shorter perfluorinated carboxylic acids: on thermal decomposition of a PFOS model compound (Altarawneh,
2021). Arrhenius coefficients for termination reactions of COF were
CF3(CF2)mCOF + H2O → HF + CF3(CF2)mC(O)OH → CF3(CF2)mCOF + HF adopted from literature: 2COF → CO + COF2 (Ninomiya et al., 2001),
+ CO COF → F + CO (Knyazev et al., 1997), and COF + F → COF2 (Cobos et al.,
CF3(CF2)mCOF + H2O → HF + CF3(CF2)mC(O)OH 1995). Kinetic modelling deploys a PFR model (reactor length of 10 cm
and an internal diameter of 0.5 cm) operated under 1 atm, at tempera­
Singh et al. (2019) suggested the appearance of the hydrolysis re­ ture between 500 and 1500 K, and a residence time of 25 s. The inlet
action in their plasma-induced water treatment of PFAS to decompose stream consists of 2% mol of perfluoropentanoic acid (M0) in N2 (i.e.,
CF3(CF2)mCOF. The unavoidable presence of moisture in real combus­ purely pyrolytic conditions). These conditions cover typical gas resi­
tion environments is expected to reduce the yield of perfluorinated al­ dence time and temperature in cement kiln systems, albeit without ac­
dehydes. For this reason, we plan to investigate the hydrolysis reaction counting for O2, H2O and additional radical pools, which are present in
in combustion systems in future work, both experimentally and the process gases. The gas residence time in kiln systems correspond to
quantum-chemically, to refine further the present model. (all values are averages) 9.3 s at 1475 K in the kiln itself, 3.9 s at 1290 K
in precalciner and 8.9 s at 1045 K in the heating cyclones (Meystre and
Silva, 2015). Thermochemical and kinetic parameters that comprise the
3.6. Kinetic modelling model are enclosed in the SM.
While the relevant experimental investigations point out to the for­
The developed model consists of 56 reactions and 45 species. Table 1 mation of several products, these studies do not provide temperature-
presents the reaction rate constants computed in this study or by dependent profiles of the species emerging from the reactor.
Altarawneh (2021) and incorporated in the model. Additional

10
M. Altarawneh et al. Chemosphere 286 (2022) 131685

Nonetheless, simulated yields of products, obtained in the present study, the reactor walls early in the pyrolysis process and is known for its
should provide a valuable insight into mechanistic trends governing the strong acidic sites (Fang et al., 2019). Our previous work on oxidative
decomposition of PFOA, and other PFCA homologues, in general, and decomposition of 4-chlorobiphenyl (Hou et al., 2013) reported a
provide the theoretical requirements for the operating conditions for remarkably lower onset temperature for its decomposition when
burning of PFAS in incinerators. Our mechanistic analysis provides compared with the analogous value measured in reactors made out of
accessible pathways, that potentially underpin the generation of major high purity quartz and high purity quartz coated with boric oxide to
experimental products such as HF, perfluorinated alkenes, CnHFm spe­ minimise catalytic wall effects (i.e, 573 versus 773 K). Overall, following
cies, and cyclic compounds. The model is expected to satisfactorily ac­ the loss of SO3/SO2 versus CO2/CO, the decomposition of our
count for evolution of products from the decomposition of the PFCA-model compound proceeds via similar degradation pathways to
perfluorinated alkyl chain. This is affirmed by the close agreement be­ these presented in our recent study for PFOS surrogate (Altarawneh,
tween experimental and product profiles in case of n-C4F10 as demon­ 2021).
strated by Yu et al. (2008). As an accuracy benchmark of the adopted Fig. 7b depicts profiles of major products while Fig. 7c and d shows
simulation framework, our computed thermokinetic parameters were mole fractions of minor CnFm and CnFmH species. Analysis of reaction
successfully contrasted with representative experimental measure­ pathways are presented in Fig. 8. Minor species include carbonyl
ments, please refer to discussion in Section 2 and the values provided in difluoride (COF2), C2F4, CF4, C2F6, C3F6 and CHF3, in addition to minute
Table S1 and Figure S1. This indicates that, the calculated activation concentrations of cyc-C4F8, C2HF3 and C2H2F4. At low decomposition
barriers for the initiation steps in the decomposition of the of per­ temperature (i.e., 800–850 K), decay of the perfluoropentanoic acid
fluoropentanoic acid should exhibit an acceptable accuracy margin, i.e., solely occurs via the two-step reaction sequence M0 → HF + M1 → CO +
within 1–3 kcal/mol. heptafluorobutanoyl fluoride (M2). Between 800 and 1100 K, mole
The model predicts an onset temperature at ~780 K for the decom­ fraction of CO closely matches that of M2 (Fig. 7b). As the temperature
position of the perfluoropentanoic acid with a full degradation at ~950 increases, direct formation of perfluoro-1-butene (CO2 + HF + M6), and
K, for the residence time of 25 s, following a steep decomposition profile to a lesser extent C4HF9 (CO2 + M5), become gradually the important
(Fig. 7a). Our predicted onset temperature exceeds the analogous values decomposition channels. At 1200 K, the generation of hepta­
determined experimentally of above 723 K. The difference (by ~60 K) in fluorobutanoyl fluoride, perfluoro-1-butene, and 1H-perfluorobutane
modelled and measured onset temperature can be rationalised consid­ accounts for 58.0%, 34.3% and 7.8%, of the initial decomposition of
ering the construction material of the reactor in Stockenhuber’s et al. the perfluoropentanoic acid (Fig. 8a), respectively. Departure of the CO
(2019) experiments. It is well-documented that minute concentration of profile from that of heptafluorobutanoyl fluoride in Fig. 7b tracks the
transition metal oxides in alumina substantially catalyse the decompo­ consumption of C3F7COF by the hydrofluorination sequence illustated in
sition of halogenated compounds, as does AlF3 that replaces Al2O3 on Fig. 4b as well as by fission of its C–COF bond (46.5%). The yield of

Fig. 7. Decay profile of perfluoropentanoic acid (a), mole fraction of major (b), and minor (c and d) species from pyrolysis of a perfluoropentanoic acid at a residence
time of 25 s.

11
M. Altarawneh et al. Chemosphere 286 (2022) 131685

Fig. 8. Analysis of initiation reaction pathways at different temperatures (a), and subsequent fragmentation pathways at 1200 K (b).

perfluoro-1-butene peaks at ~1300 K accounting for the mole fraction of mainly in CF3. → ꞉CF2 + F and C3F7COF ⇆ COF. + n-C3F7. reactions
17% of the product gases, as illustrated in Fig. 7b. between 900 and 1300 K. The model predicts a key role of F atoms in
The addition of HF consumes 53.5% of formed heptafluorobutanoyl
fluoride at 1200 K (Fig. 8b). The hydrofluorination reaction accounts for
chain shortening reactions, which eventually lead to the formation of
perfluoropropene (58.6%) and of 1H-perfluoropropane (41.4%). None­
theless, perfluoropropene (C3F6) readily degrades into ꞉CF2 and C2F4
(99%, Fig. 8b) with a maximum yield of only 0.005 (based on mole
fraction of products, Fig. 7d) around 1300 K. Formation of 1H-perfluor­
opropane mimics the synthesis of the experimentally detected product
1H-perfluorohexane in the experiment of Stockenhuber et al. (2019). As
shown in Fig. 7a, HF remains the leading product at all temperatures,
despite of its significant depletion via the hydrofluorination reaction:
heptafluorobutanoyl fluoride + HF. Hydrofluorination of per­
fluoropropene into 1H-perfluoropropane is rather limited, consuming
only 2% of the formed perfluoropropene. As pictured in Fig. 7c, the yield
of 1H-perflurobutane attains its maximum value at 0.03 at around 1200
K.
Formation of smaller CnFm fragments commences via fission of the
C–C(O)OH in perfluoropentanoic acid. The fluoropropyl radical (n-C3F7)
decomposes into ꞉CF2 + C2F5. (21.6%) and CF3. + C2F4 (78.4%). Frag­
mentation of CF3. produces a fluorine atom and ꞉CF2. Kinetic parameters
of this important reaction were adopted from a shock-tube study of
Cobos et al. (2010). Computed coefficients from the sensitivity analysis
on the production of F atoms (shown in Fig. 9) reveal that F atoms arise Fig. 9. Normalised sensitivity analysis for production of F atoms at the middle
of the reactor. The equal signs in the legend denote reversible kinetic reactions.

12
M. Altarawneh et al. Chemosphere 286 (2022) 131685

driving decomposition of the perfluoropentanoic acid at elevated tem­ reaction. The yield of C2F4 from pyrolysis of polytetrafluoroethylene
peratures. The major consumption route of perfluoro-1-butene stems (PTFE) reached 76 wt% at 823 K (Simon and Kaminsky, 1998). Per­
from its reaction with F atom to produce the n-C4F9 radical. Abstraction fluoropropene accounted for 7.1 wt% of the initial mass of PTFE (Simon
of the hydroxyl H atom via the reaction F + M0 → M4 + HF accounts for and Kaminsky, 1998). Similarly, the yield of cyc-C4F8 approached 8%
25.6% of the total consumption of the perfluoropentanoic acid at 1500 from the pyrolysis of PTFE, rising to 29–34% at 923–973 K (Mięsowicz,
K. A direct loss of CO2 marks the sole fate of the M4 radical and 1987). The predicted low yield of cyc-C4F8 (2 × 10− 5 at 1200 K) herein
significantly contributes to the formation of CO2 molecules at high can be understood in view of the rapid degradation of its main building
temperatures as documented in Fig. 7b. At low and intermediate tem­ block (C2F4) into ꞉CF2 at temperatures higher than 1200 K. In other
peratures, n-C4F9. forms via addition of a fluorine atom to the n-C4F8 words, formation temperature of C2F4 (from decomposition of
molecule (Fig. 8b, right-hand side). The importance of the channel n-C3F7./n-C4F9.) coincides with its fragmentation into ꞉CF2 molecules.
leading to heptafluorobutanoyl fluoride diminishes with temperature. Overall, in the absence of gas-solid reactions, at > 1500 K, we predict
At 1200 K, the n-C4F9 radical fractures into C2F4/C2F5. (91.2%) and the emission of ꞉CF2 (biradical difluorocarbene), HF, CO2, CO, CF4, C2F6,
꞉CF2/n-C3F7. (8.8%). Decomposition of C2F5 radical gives ꞉CF2 and CF3.. and C2F4, as the terminal products of decomposition of PFAS, but at <
At 1200 K, nearly 50% of the formed C2F4 molecules (from degradation 1400 K, we note the formation of C4HF9, COF2, and C3F7COF, with 1-
of n-C3F7 and n-C4F7 radicals) decompose to produce ꞉CF2 molecules. C4F8, 2-C4F8 and C3HF7 persisting to 1500 K. These predictions have
The addition of HF to ꞉CF2 to generate CHF3 consumes 12.7% of great significance to the remediation of PFAS in incinerators, but we
difluorocarbene. Fragmentation of all CnFm species at high temperatures wish to stress that our study has not considered the role of flyash in these
produces ꞉CF2 as the major fluorine carrier, beyond 1400 K. systems, and did not include gas composition (moisture, oxygen, fuel)
Self-addition of CF3 radicals to form C2F6 represents a minor contribu­ typically encountered in incinerators. For instance, municipal waste
tion to the flux of fluorine between 1100 and 1300 K. combustors display 10% O2 and 15% moisture at the exhaust (Taylor
Due to the negligible formation of H/OH radicals in the current et al., 2014). In addition, pyrolysing or oxidising waste itself provides a
model, their contribution in decomposition of the perfluoropentanoic hydrogen source to mineralise fluorine to HF. We expect these consid­
acid remains limited. However, in an environment rich with O/H radical erations to improve the mineralisation of PFAS in real combustion sys­
pool (as in co-oxidation of perfluorinated carboxylic acids with hydro­ tems. Thus, our model needs to be executed by future users under their
carbons and other fuels), abstraction of the hydroxyl group leading to target conditions. Short residence time of 2.0 s (a value that typically
CO2 and n-CnF2n+1 radical will most likely constitute a dominant chan­ applies in waste incinerators) slightly alters the predicted yield of major
nel. Such reactions may prevail under real scenarios; i.e., when, per­ fluorine-bearing species at elevated temperatures and delays the com­
fluorinated carboxylic acids (and PFAS in general) are treated in thermal plete elimination of PFPA to around 1020 K, as illustrated in Fig. 7a.
destruction systems. For this reason, it is of a practical importance to Figure S3 contrasts mole fractions of ꞉CF2, C2F4, C2F6 and HF attained
investigate transformation chemistry that dictates interaction of small with 2 and 25 residence times. A lower yield of HF in case of a longer
perfluorinated species with commonly deployed catalysts in the capture residence time coincides with its consumption through the illustrated
and treatments of PFAS streams, i.e., alumina and Ca(OH)2, for PFAS bimolecular addition reactions.
decomposition and fluorine mineralisation in cement kilns or with other We conclude our discussion with two comments. Firstly, we note that
metal oxides that predominate in fly ash in the incinerators. while PFPA (or any PFCA in general) disappears by 1020 K, C–C bonds in
Formation of 1H-perfluropropane from hydrofluorination of per­ the product molecules start to show significant fission only at temper­
fluoropropene consumes only 2.0% of C3F6; not shown in the figures. To atures above 1200 K, with the transitional species disappearing by 1500
further address the hydrofluorination capacity of perfluorinated al­ K. These are the theoretical requirements for handling PFAS in thermal
kenes, in Fig. 10, we plot the temperature-dependent mole fraction of systems, based solely on the gas-phase reactions, in the absence of
C2HF5 that results from the reaction of an equimolar mixture of HF and moisture and oxygen in the gas phase, and heterogenous processes in the
C2F4. Formation of C2HF5 appears to be minimal, even when allowing a modelling. We expect that these theoretical limits can be further refined
longer residence time of 60 s for the occurrence of this bimolecular (i.e., decreased) in subsequent calculations, in which our model will
serve as a submodel in comprehensive computations. Secondly, the ef­
ficiency of practical incineration systems is defined by temperature,
residence time and turbulence (mixing). It is well known that, highly
turbulent conditions in incinerators provide uniform temperature and
minimise the emissions of products of incomplete combustion (e.g.
(Winchell et al., 2021),). The plug flow reactor model deployed in this
study assumes perfect mixing but only in the direction perpedicular to
flow, and no mixing in the axial direction. From this perspective, future
investigations should also include calculations of PFAS destruction and
mineralisation in stirred reactors.

4. Conclusions

This contribution reports thermokinetic parameters for a wide range


of reactions that govern the initial decomposition of the per­
fluoropentanoic acid, a model PFAS species, and the subsequent evo­
lution of the composition of reaction gases as function of residence time.
Modelled profiles of product species and the analysis of reaction path­
ways reveal the competitive importance of two routes; namely, the
formation of heptafluorobutanoyl fluoride + HF + CO (at low and in­
termediates temperatures) and production of perfluro-1-butene + CO2
+ HF (at high temperatures). Decomposition of HF over the C– – O bond
Fig. 10. Effect of residence time on the formation of C2HF5 from the reaction in heptafluorobutanoyl fluoride initiates an important pathway that
HF + C2F4 ⇆ C2HF5 starting from an equimolar concentration of the transforms C3F7COF into 1H-perfluoropropane. Fission of C–COF bond
two reactants. in the heptafluorobutanoyl fluoride affords the generation of ꞉CF2

13
M. Altarawneh et al. Chemosphere 286 (2022) 131685

molecules at high temperatures. Hydrofluorination of perfluorinated Acknowledgments


alkenes into their respective hydrofluoroalkanes constitutes a minor
corridor at all investigated temperatures. Formation of cyclic per­ This study has been supported by a UPAR grant from the United Arab
fluorinated compounds via ring-closure reactions of their corresponding Emirates University (grant number: 31N451) and a computational time
diradical adducts arises via accessible activation enthalpies. The tem­ allocation from Compute Canada.
perature window for the formation of C2F4 favours its degradation into
꞉CF2 molecules, rather than its self-addition to produce cyc-C4F8. The Appendix A. Supplementary data
presence of H/O/F radical in combustion medium is anticipated to shift
the product profiles toward formation of n-CnF2n+1 radicals, that readily Supplementary data to this article can be found online at https://doi.
degrade into F, ꞉CF2, and C2F4, and eventually to HF in the presence of a org/10.1016/j.chemosphere.2021.131685.
hydrogen source.
The analysis of reaction pathways provided an important insight into References
the relative importance of plausible initial decomposition channels. The
formation of a fluorine atom via the radical reactions such as CF3 ⇆ F + Afeefy, H.Y., Liebman, J.F., Stein, S.E., Linstrom, P.J., Mallard, W.G., 2005. NIST
Chemistry WebBook, NIST Standard Reference Database Number 69.
CF2 drives a facile degradation route of the parent perfluoropentanoic Ahubelem, N., Shah, K., Moghtaderi, B., Altarawneh, M., Dlugogorski, B.Z., Page, A.J.,
acid that results in the generation of short perfluorinated species. As 2015. Formation of chlorobenzenes by oxidative thermal decomposition of 1,3-
PFAS are generally deployed in formulations with hydrocarbon and dichloropropene. Combust. Flame 162, 2414–2421.
Ainagos, A.F., 1991. Mechanism and kinetics of pyrolysis of perfluorohexane. Kinet.
carbohydrate materials (e.g., hydrocarbon surfactants and gums present Catal. 32, 720–725.
in concentrates of PFAS foams) future work should focus on developing Alm, R.R., Stern, R.M., 1992. Aqueous film-forming foamable solution useful as fire
reaction mechanisms and chemical kinetic models for the co-pyrolysis/ extinguishing concentrate. In: 3M Co, US. US Patent 5,085,786.
Altarawneh, M., 2012. A theoretical study on the pyrolysis of perfluorobutanoic acid as a
oxidation of PFAS with these materials or with typical fuels. model compound for perfluoroalkyl acids. Tetrahedron Lett. 53, 4070–4073.
While this study considered H and OH to arise only from the Altarawneh, M., 2021. A chemical kinetic model for the decomposition of perfluorinated
decomposition of perfluoropentanoic acid itself, in incinerators and sulfonic acids. Chemosphere 263, 128256.
Altarawneh, M., Altarawneh, K., 2012. A theoretical study on the bimolecular reactions
other practical combustion systems, these radical pools will increase in
encountered in the pyrolysis of acetamide. J. Phys. Org. Chem. 25, 431–436.
size, prompting the formation of F, CF2 and C2F4 at lower temperatures Altarawneh, M., Dlugogorski, B.Z., 2014a. Mechanism of thermal decomposition of
and shorter residence time, and enhanced mineralisation of fluorine. Tetrabromobisphenol A (TBBA). J. Phys. Chem. 118, 9338–9346.
Our work provides a suitable submodel to be incorporated into future Altarawneh, M., Dlugogorski, B.Z., 2014b. Thermal decomposition of 1,2-Bis(2,4,6-tri­
bromophenoxy)ethane (BTBPE), a novel brominated flame retardant. Environ. Sci.
comprehensive mechanisms that will also include heterogenous pro­ Technol. 48, 14335–14343.
cesses. Such models should also comprise coupling between fluorine Altarawneh, M., Dlugogorski, B.Z., Kennedy, E.M., Mackie, J.C., 2009. Mechanisms for
reservoirs in the gas and solid phases. Evidently, attempts to destroy formation, chlorination, dechlorination and destruction of polychlorinated dibenzo-
p-dioxins and dibenzofurans (PCDD/Fs). Prog. Energy Combust. Sci. 35, 245–274.
PFAS in systems, which are operated at too low temperature or at too Altarawneh, M., Saeed, A., Al-Harahsheh, M., Dlugogorski, B.Z., 2019. Thermal
low residence time and do not include a means to bind the fluorine, may decomposition of brominated flame retardants (BFRs): products and mechanisms.
convert PFAS to short-chain perfluoro- and hydrofluorocarbons, and Prog. Energy Combust. Sci. 70, 212–259.
Assaf, E., Schoemaecker, C., Vereecken, L., Fittschen, C., 2018. The reaction of fluorine
other species such as COF2, known for its toxicity. From this perspective, atoms with methanol: yield of CH3O/CH2OH and rate constant of the reactions CH3O
further calculations are needed to gain understanding of the role of + CH3O and CH3O + HO2. Phys. Chem. Chem. Phys. 20, 10660–10670.
oxygen, moisture and fuel (as source of hydrogen) on the mineralisation Backe, W.J., Day, T.C., Field, J.A., 2013. Zwitterionic, cationic, and anionic fluorinated
chemicals in aqueous Film forming foam formulations and groundwater from U.S.
of fluorine to HF in gas phase reactions and on the destruction of tran­ military bases by nonaqueous large-volume injection HPLC-MS/MS. Environ. Sci.
sitional species. We have demonstrated that the gas phase conditions, in Technol. 47, 5226–5234.
the absence of reactions with solid phase species (such as clinker), in Barzen-Hanson, K.A., Roberts, S.C., Choyke, S., Oetjen, K., McAlees, A., Riddell, N.,
McCrindle, R., Ferguson, P.L., Higgins, C.P., Field, J.A., 2017. Discovery of 40 classes
both incinerators and cement kilns, lead to similar product distribution,
of per- and polyfluoroalkyl substances in historical aqueous film-forming foams
provided that the gas phase temperature exceeds 1200 K. Note that we (AFFFs) and AFFF-impacted groundwater. Environ. Sci. Technol. 51, 2047–2057.
also predict the emission of the toxic COF2 for the operating temperature Bentel, M.J., Yu, Y., Xu, L., Li, Z., Wong, B.M., Men, Y., Liu, J., 2019. Defluorination of
of up to 1400 K, as a minor product, and a range C3–C4 fluorocarbons up per- and polyfluoroalkyl substances (PFASs) with hydrated electrons: structural
dependence and implications toPFAS remediation and management. Environ. Sci.
to 1500 K. Thus, our model predicts 1020 K and 1500 K as the theo­ Technol. 53, 3718–3728.
retical temperatures for destruction of PFAS species and evolution of HF, Blake, P.G., Tomlinson, A.D., 1971. Thermal decomposition of fluoroacetic acid.
respectively, in incinerators operated at the residence time of 2 s. J. Chem. Soc. B 1596–1597.
Brahms, D.L.S., Dailey, W.P., 1996. Fluorinated carbenes. Chem. Rev. 96, 1585–1632.
Careful field studies are also needed to investigate the role of flyash in Canneaux, S., Bohr, F., Henon, E., 2014. KiSThelP: a program to predict thermodynamic
mineralisation of fluorine from PFAS in incinerators, in conjunction with properties and rate constants from quantum chemistry results. J. Comput. Chem. 35,
measurements of emissions of PFC and HFC, to determine whether such 82–93.
Clark, J.M., Nimlos, M.R., Robichaud, D.J., 2015. Bimolecular decomposition pathways
systems do indeed discharge these species. It is important to emphasise for carboxylic acids of relevance to biofuels. J. Phys. Chem. 119, 501–516.
that we predict no formation of the highly toxic perfluoroisobutene, Cobos, C.J., Croce, A.E., Castellano, E., 1995. Kinetics of the recombination reaction
even as a minor species, under all temperatures and residence times between F atoms and FCO radicals. Chem. Phys. Lett. 239, 320–325.
Cobos, C.J., Croce, A.E., Luther, K., Troe, J., 2010. Shock wave study of the thermal
considered in this study. decomposition of CF3 and CF2 radicals. J. Phys. Chem. 114, 4755–4761.
Darwin, R.I., 2004. Estimated quantities of aqueous film forming foam (AFFF) in the
Credit author statement United States. In: Associates, H. (Ed.), Report Prepared for the Fire Fighting Foam
Coalition.
David, J., Philip, H., 1997. An in situ IR study of the thermal decomposition of
Mohammednoor Altarawneh: Conceptualisation, Methodology, trifluoroacetic acid. J. Chem. Soc., Perkin Trans. 2, 1571–1576.
Writing - Original Draft, Formal Analysis, Mansour H. Almatarneh: Dellinger, B., Taylor, P.H., Tirey, D.A., Lee, C.C., 1989. Pathways of formation of
Formal analysis, Bogdan Z. Dlugogorski: Conceptualisation, Writing – chlorinated PICs from the thermal degradation of simple chlorinated hydrocarbons.
J. Hazard Mater. 22, 175–185.
Review & Editing, Data Curation. Dlugogorski, B.Z., Schaefer, T.H., 2021. Compatibility of aqueous film-forming foams
(AFFF) with sea water. Fire Saf. J. 120, 103288.
Declaration of competing interest Dóbé, S., Bérces, T., Temps, F., Wagner, H.G., Ziemer, H., 1994. Formation of methoxy
and hydroxymethyl free radicals in selected elementary reactions. Symp. Combust.
Proc. 25, 775–781.
Authors declare no conflict of interest. Dubocq, F., Wang, T., Yeung, L.W.Y., Sjöberg, V., Kärrman, A., 2020. Characterization of
the chemical contents of fluorinated and fluorine-Free firefighting foams using a

14
M. Altarawneh et al. Chemosphere 286 (2022) 131685

novel workflow combining nontarget screening and total fluorine nalysis. Environ. Mięsowicz, H., 1987. Depolimeryzacja politetrafluoroetylenu. Przem. Chem. 66,
Sci. Technol. 54, 245–254. 333–337.
Ellis, D.A., Mabury, S.A., Martin, J.W., Muir, D.C.G., 2001. Thermolysis of Mokrushin, V.B., Bedanov, V., Tsang, W., Zachariah, M., Knyazev, V., 2002. ChemRate.
fluoropolymers as a potential source of halogenated organic acids in the NIST, Gaithersburg, MD.
environment. Nature 412, 321–324. Montgomery, J.A., Frisch, M.J., Ochterski, J.W., Petersson, G.A., 2000. A complete basis
Fang, X.-X., Wang, Y., Jia, W.-Z., Song, J.-D., Wang, Y.-J., Luo, M.-F., Lu, J.-Q., 2019. set model chemistry. VII. Use of the minimum population localization method.
Dehydrofluorination of 1, 1, 1, 3, 3-pentafluoropropane over C-AlF3 composite J. Chem. Phys. 112, 6532–6542.
catalysts: improved catalyst stability by the presence of pre-deposited carbon. Appl. Nguyen, M.T., Sengupta, D., Raspoet, G., Vanquickenborne, L.G., 1995. Theoretical study
Catal. Gen. 576, 39–46. of the thermal decomposition of acetic acid: decarboxylation versus dehydration.
Frisch, M.J.T., Schlegel, H.B., Scuseria, G.E., Robb, M.A., Cheeseman, J.R., Scalmani, G., J. Phys. Chem. 99, 11883–11888.
Barone, V., Mennucci, B., Petersson, G.A., Nakatsuji, H., Caricato, M., Li, X., Ninomiya, Y., Goto, M., Hashimoto, S., Kawasaki, M., Wallington, T.J., 2001. Cavity ring-
Hratchian, H.P., Izmaylov, A.F., Bloino, J., Zheng, G., Sonnenberg, J.L., Hada, M., down spectroscopic study of the kinetics of the reactions of FCO radicals with O2 and
Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., NO at 295 K. Int. J. Chem. Kinet. 33, 130–135.
Kitao, O., Nakai, H., Vreven, T., Montgomery, J.A., Peralta, J.E., Ogliaro, F., Reaction Design, 2013. CHEMKIN-PRO.
Bearpark, M., Heyd, J.J., Brothers, E., Kudin, K.N., Staroverov, V.N., Kobayashi, R., Richter, H., Howard, J.B., 2000. Formation of polycyclic aromatic hydrocarbons and
Normand, J., Raghavachari, K., Rendell, A., Burant, J.C., Iyengar, S.S., Tomasi, J., their growth to soot—a review of chemical reaction pathways. Prog. Energy
Cossi, M., Rega, N., Millam, J.M., Klene, M., Knox, J.E., Cross, J.B., Bakken, V., Combust. Sci. 26, 565–608.
Adamo, C., Jaramillo, J., Gomperts, R., Stratmann, R.E., Yazyev, O., Austin, A.J., Riedel, T.P., Wallace, M.A.G., Shields, E.P., Ryan, J.V., Lee, C.W., Linak, W.P., 2021. Low
Cammi, R., Pomelli, C., Ochterski, J.W., Martin, R.L., Morokuma, K., Zakrzewski, V. temperature thermal treatment of gas-phase fluorotelomer alcohols by calcium
G., Voth, G.A., Salvador, P., Dannenberg, J.J., Dapprich, S., Daniels, A.D., Farkas, Ö., oxide. Chemosphere 272, 129859.
Foresman, J.B., Ortiz, J.V., Cioslowski, J., Fox, D.J., 2009. Gaussian 09. Gaussian, Schroeder, T., Bond, D., Foley, J., 2021. PFAS soil and groundwater contamination via
Inc, Wallingford CT. industrial airborne emission and land deposition in SW Vermont and Eastern New
Gao, X., Chorover, J., 2012. Adsorption of perfluorooctanoic acid and York State, USA. Environ. Sci.: Processes Impacts 23, 291–301.
perfluorooctanesulfonic acid to iron oxide surfaces as studied by flow-through ATR- Schug, K.P., Wagner, H.G., Zabel, F., 1979. Gas phase α elimination of hydrogen halides
FTIR spectroscopy. Environ. Chem. 9, 148–157. from halomethanes. I. Thermal decomposition of chlorodifluoromethane,
Garg, S., Wang, J., Kumar, P., Mishra, V., Arafat, H., Sharma, R.S., Dumée, L.F., 2021. trifluoromethane, and trichloromethane behind shock. Ber. Bunsenges. Phys. Chem.
Remediation of water from per-/poly-fluoroalkyl substances (PFAS) – challenges and 83, 167–175.
perspectives. J. Env. Chem. Eng. 9, 105784. Simon, C.M., Kaminsky, W., 1998. Chemical recycling of polytetrafluoroethylene by
Garrett, B.C., Truhlar, D.G., 2005. Chapter 5 - Variational transition state theory. In: pyrolysis. Polym. Degrad. Stabil. 62, 1–7.
Dykstra, C.E., Frenking, G., Kim, K.S., Scuseria, G.E. (Eds.), Theory and Applications Singh, R.K., Fernando, S., Baygi, S.F., Multari, N., Thagard, S.M., Holsen, T.M., 2019.
of Computational Chemistry. Elsevier, Amsterdam, pp. 67–87. Breakdown products from Perfluorinated alkyl substances (PFAS) degradation in a
Han, W., Kennedy, E.M., Kundu, S.K., Mackie, J.C., Adesina, A.A., Dlugogorski, B.Z., plasma-based water treatment process. Environ. Sci. Technol. 53, 2731–2738.
2010. Experimental and chemical kinetic study of the pyrolysis of trifluoroethane Snow, A.W., Hinnant, K.M., Farley, J.P., Ananth, R., 2017. Quantification of Fluorine
and the reaction of trifluoromethane with methane. J. Fluor. Chem. 131, 751–760. Content in AFFF Concentrates. NRL/MR/6120—17-9752.
Hidaka, Y., Nakamura, T., Kawano, H., 1991. High temperature pyrolysis of CF3H in Starnes, W.H., 2002. Structural and mechanistic aspects of the thermal degradation of
shock waves. Chem. Phys. Lett. 187, 40–44. poly(vinyl chloride). Prog. Polym. Sci. 27, 2133–2170.
Holmes, N., Klein, R.A., 2020. PFAS Waste Management. Queensland Government and Stockenhuber, S., Weber, N., Dixon, L., Lucas, J., Grimison, C., Stockenhuber, M.,
Green Science Policy Institute. Mackie, J.C., Kennedy, E.M., 2019. Thermal degradation of perfluorooctanonic acid
Hori, H., Ushio, T., Asai, T., Honma, R., Eid, N., Ameduri, B., 2021. Efficient (PFOA). 16th Internationl Conferenc on Environmental Science and Technololgy
mineralization of a novel fluorotelomer surfactant, 2H,3H,3H,5H,5H,6H,6H-4-thia- Rhodes, Greece.
perfluoro(2-methyl)-1-dodecanoic acid, in superheated water induced by a Taylor, P.H., Yamada, T., Striebich, R.C., Graham, J.L., Giraud, R.J., 2014. Investigation
combination of potassium permanganate and dioxygen. Chem. Eng. J. 405, 127006. of waste incineration of fluorotelomer-based polymers as a potential source of PFOA
Horst, J., Quinnan, J., McDonough, J., Lang, J., Storch, P., Burdick, J., Theriault, C., in the environment. Chemosphere 110, 17–22.
2021. Transitioning per- and polyfluoroalkyl substance containing fire fighting Wang, F., Liu, C., Shih, K., 2012. Adsorption behavior of perfluorooctanesulfonate
foams to new alternatives: evolving methods and best practices to protect the (PFOS) and perfluorooctanoate (PFOA) on boehmite. Chemosphere 89, 1009–1014.
environment. Groundwater Monit R 41, 19–26. Wang, F., Lu, X., Li, X.-y., Shih, K., 2015. Effectiveness and mechanisms of defluorination
Hou, S., Mackie, J.C., Kennedy, E.M., Dlugogorski, B.Z., 2013. Comparative study on the of perfluorinated alkyl substances by calcium compounds during waste thermal
formation of toxic species from 4-chlorobiphenyl in fires: effect of catalytic surfaces. treatment. Environ. Sci. Technol. 49, 5672–5680.
Procedia Eng 62, 350–358. Wang, N., Lv, H., Zhou, Y., Zhu, L., Hu, Y., Majima, T., Tang, H., 2019. Complete
Hratchian, H.P., Schlegel, H.B., 2005. Using hessian updating to increase the ifficiency of defluorination and mineralization of perfluorooctanoic acid by a mechanochemical
a Hessian based predictor-corrector reaction path following method. J. Chem. Theor. method using alumina and persulfate. Environ. Sci. Technol. Lett. 53, 8302–8313.
Comput. 1, 61–69. Weber, N.H., Stockenhuber, S.P., Delva, C.S., Abu Fara, A., Grimison, C.C., Lucas, J.A.,
Khan, M.Y., So, S., da Silva, G., 2020. Decomposition kinetics of perfluorinated sulfonic Mackie, J.C., Stockenhuber, M., Kennedy, E.M., 2021. Kinetics of decomposition of
acids. Chemosphere 238, 124615. PFOS relevant to thermal desorption remediation of soils. Ind. Eng. Chem. Res. 60,
Knyazev, V.D., Bencsura, Á., Slagle, I.R., 1997. Unimolecular decomposition of the FCO 9080–9087.
radical. J. Phys. Chem. 101, 849–852. Winchell, L.J., Ross, J.J., Wells, M.J.M., Fonoll, X., Norton Jr., J.W., Bell, K.Y., 2021. Per-
Krusic, P.J., Marchione, A.A., Roe, D.C., 2005. Gas-phase NMR studies of the thermolysis and polyfluoroalkyl substances thermal destruction at water resource recovery
of perfluorooctanoic acid. J. Fluor. Chem. 126, 1510–1516. facilities: a state of the science review. Water Environ. Res. 93, 826–843.
Lee, Y.-M., Lee, J.-Y., Kim, M.-K., Yang, H., Lee, J.-E., Son, Y., Kho, Y., Choi, K., Zoh, K.- Xiao, F., Sasi, P.C., Yao, B., Kubátová, A., Golovko, S.A., Golovko, M.Y., Soli, D., 2020.
D., 2020. Concentration and distribution of per- and polyfluoroalkyl substances Thermal stability and decomposition of perfluoroalkyl substances on spent granular
(PFAS) in the Asan Lake area of South Korea. J. Hazard Mater. 381, 120909. activated carbon. Environ. Sci. Technol. Lett. 7, 343–350.
Lewis, E.E., Naylor, M.A., 1947. Pyrolysis of polytetrafluoroethylene. J. Am. Chem. Soc. Yu, H., Kennedy, E.M., Ong, W.-H., Mackie, J.C., Han, W., Dlugogorski, B.Z., 2008.
69, 1968–1970. Experimental and kinetic studies of gas-phase pyrolysis of n-C4F10. Ind. Eng. Chem.
Liu, J., Qu, R., Wang, Z., Mendoza-Sanchez, I., Sharma, V.K., 2017. Thermal- and photo- Res. 47, 2579–2584.
induced degradation of perfluorinated carboxylic acids: kinetics and mechanism. Zhao, Y., Truhlar, D., 2008. The M06 suite of density functionals for main group
Water Res. 126, 12–18. thermochemistry, thermochemical kinetics, noncovalent interactions, excited states,
Meystre, J.A., Silva, R.J., 2015. Comparison of Gas Residence Time in Different Cement and transition elements: two new functionals and systematic testing of four M06-
Production Systems for the Co-processing of Municipal Solid Waste. 23rd ABCM class functionals and 12 other functionals. Theor. Chem. Acc. 120, 215–241.
International Congress of Mechanical Engineering. Brazilian Society of Mechanical
Sciences and Engineering, Rio de Janeiro, RJ, Brazil.

15

You might also like