Diurnal Photodegradation of Fluorinated Diketones (FDKS) by OH Radicals

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Chemosphere 286 (2022) 131562

Contents lists available at ScienceDirect

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

Diurnal photodegradation of fluorinated diketones (FDKs) by OH radicals


using different atmospheric simulation chambers: Role of keto-enol
tautomerization on reactivity
Pedro L. Lugo a, V.G. Straccia a, Cynthia B. Rivela a, Iulia Patroescu-Klotz b, Niklas Illmann b,
Mariano A. Teruel a, Peter Wiesen b, Maria B. Blanco a, *
a
(L.U.Q.C.A), Laboratorio Universitario de Química y Contaminación del Aire. Instituto de Investigaciones en Fisicoquímica de Córdoba (I.N.F.I.Q.C.), Dpto. de
Fisicoquímica. Facultad de Ciencias Químicas, Universidad Nacional de Córdoba. Ciudad Universitaria, 5000, Córdoba, Argentina
b
Institute for Atmospheric and Environmental Research, Bergische Universität Wuppertal, 42097, Wuppertal, Germany

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

• Predominance of enolic tautomers than


the keto form on the reactivity of fluo­
rinated diketones.
• Predominant addition of OH to the
double bond in the enol form of fluori­
nated diketones.
• Significant tropospheric ozone forma­
tion (POCP) for the fluorinated dike­
tones (23–34).
• Relative high acidification potentials
(AP) for the fluorinated diketones
(0.5–0.6).
• Short tropospheric lifetimes, about
hours, with negligible contribution to
global warming.

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

Handling Editor: R Ebinghaus Rate coefficients for the gas-phase reactions of OH radicals with a series of fluorinated diketones have been
determined for the first time at (298 ± 3) K and atmospheric pressure using the relative method and FTIR
Keywords: spectroscopy and GC-FID to monitor both reactants and references. The following values, in 10− 11 cm3 mole­
Diketones cule− 1 s− 1, were obtained for 1,1,1-trifluoro-2,4-pentanedione (TFP), 1,1,1-trifluoro-2,4-hexanedione (TFH) and
Fluorinated carbonyls
1,1,1-trifluoro-5-methyl-2,4-hexanedione (TFMH), respectively: k1(TFP + OH) = (1.3 ± 0.4), k2(TFH + OH) =
Keto-enolic tautomerization
(2.2 ± 0.8), k3(TFMH + OH) = (3.3 ± 1.0). The results are discussed with respect to the keto-enolic tautome­
In situ FTIR
SPME-GC-FID rization specific for β-diketones.
Reactivity Based on the present results, the tropospheric lifetimes of TFP, TFH and TFMH upon degradation by OH
radicals were calculated as 21, 13 and 8 h, respectively indicating that transport might play a role in the at­
mospheric fate of the studied compounds. Photochemical ozone creation potentials were estimated for TFP, TFH
and TFMH to be: 23, 29 and 34, respectively.

* Corresponding author.
E-mail address: m.belen.blanco@unc.edu.ar (M.B. Blanco).

https://doi.org/10.1016/j.chemosphere.2021.131562
Received 1 April 2021; Received in revised form 5 July 2021; Accepted 13 July 2021
Available online 19 July 2021
0045-6535/© 2021 Elsevier Ltd. All rights reserved.
P.L. Lugo et al. Chemosphere 286 (2022) 131562

1. Introduction 2.1. BUW chamber

Fluorinated diketones (FDKs) are extensively used in organic syn­ The BUW chamber consists of a 3 m long borosilicate glass cylinder
thesis of drugs, special dyes and coatings (De Rosa et al., 2015; Ilmi (45 cm inner diameter) closed at both ends by aluminium flanges, on
et al., 2019; Isakova et al., 2010) and metallurgy, where they act as which various ports for gas and reactants supply and sampling and in­
chelating agents (Babain et al., 2002). Some of these applications im­ struments monitoring physical parameters inside the chamber are
plies metal-organic chemical vapor deposition techniques that is a po­ mounted. The reactor, with a total volume of 480 L, is surrounded by 32
tential source of air borne diketones (Bareño et al., 2020). Due to their evenly spaced fluorescent lamps (Philips TLA 40 W, 300 ≤ λ ≤ 450 nm,
structure, possessing both C–F bonds and carbonyl moieties, they might λmax = 360 nm). The whole construction is encased in reflective steel
bear a significant global warming potential (GWP (Hodnebrog et al., sheets and cooled with air. The homogeneous mixing of the reactants is
2013)). While a series of hydrofluoroolefines, -ethers, -alcohols and ensured by a magnetically coupled Teflon fan placed on the front flange,
-ketones have been studied with respect to their atmospheric fate (ki­ inside the chamber. Between experiments the reactor is evacuated by a
netic and mechanistic information) and GWP values (Jiménez et al., double stage pumping system, a fore pump and a roots vacuum booster,
2016; Mellouki et al., 2015; Paul et al., 2020; Wallington et al, 2004, to an end vacuum up to 10− 3 Torr. Cleanliness is accomplished by filling
2015) for FDKs no data were yet reported. the reactor up to 200 mbar and evacuate to vacuum and repeating this
Prediction of rate constants through structure activity relationships sequence until no absorption bands belonging to reactants or products
(SAR) is a proficient tool for many substances those gas-phase reactions are present in the IR spectra.
with atmospheric oxidants cannot be investigated directly. The SAR The concentration-time profile of reactants and products are moni­
model developed by the Atkinson group in US (Atkinson, 1987; Kwok tored in situ using a multi-reflection White mirror system (2.80 ± 0.01 m
and Atkinson, 1995) works quite well for some class of substances but base length) mounted in the reactor and coupled to a Nicolet 6700 FTIR
there are many exceptions, halogenated compounds being among them. spectrometer (MCT detector). The White system is operated at 18 tra­
The inconsistencies result from missing an extensive kinetic data pool. verses which yields a 50.4 m path length. The IR spectra were recorded
Therefore, the latest SAR development excludes estimation of rates for in the 700-4000 cm− 1 spectral range with 1 cm− 1 resolution by co-
the reaction of OH radicals with halogen or sulfur containing com­ adding 30 interferograms over (30 ± 2) s. Typically, during an experi­
pounds for which the information is relatively scarce (Jenkin et al., ment were recorded 15 spectra, the first 5 without irradiation.
2018). Works as the present study contribute to refine and complete our In the BUW chamber the OH radicals were produced using the
understanding of SAR models. photolysis of methyl nitrite (CH3ONO) in the presence of NO (Atkinson
Beside the importance for the climate impact modelling and the et al., 1995):
contribution to a comprehensive gas-phase kinetics data base, fluori­
CH3 ONO + hv→CH3 O + NO (4)
nated β-diketones are of particular interest because of the keto-enolic
tautomerization specific for this kind of compounds. The existence of
CH3 O + O2 →CH2 O + HO2 (5)
this equilibrium is well known for 1,3-diketones and was described in
the liquid (Burdett and Rogers, 1964) as well as in the gas phase HO2 + NO→NO2 + OH (6)
(Nakanishi et al., 1977) while few reports exist on the influence of
electron-withdrawing substituents (halogen atoms) favouring the enolic
tautomer in solution (Bassetti et al., 1988; Lintvedt and Holtzclaw, 2.2. LUQCA chamber
1966). Similarly, the OH initiated gas-phase oxidation of diketones was
extensively studied (Bell et al., 2006; Holloway et al., 2005; Messaadia The LUQCA chamber, installed at the National University of Córdoba
et al., 2015; Zhou et al., 2008), but, by best of our knowledge, no similar has a similar design to the BUW chamber, being a borosilicate cylinder
information was ever reported for the fluorinated β-diketones. sealed at both ends by aluminum flanges, but the dimensions and some
In this work, we report for the first time the rate coefficients for the construction details differs. The cylinder, of 1.5 m length and 0.6 m
reactions of OH radicals with a series of FDKs: 1,1,1-trifluoro-2,4-penta­ inner diameter, has a total volume of 405 L. Homogeneity of the gas
nedione (TFP), 1,1,1-trifluoro-2,4-hexanedione (TFH) and 1,1,1-tri­ mixtures is ensured similar to the BUW chamber.
fluoro-5-methyl-2,4-hexanedione (TFMH) as follows: The glass cylinder includes, placed eccentrically at 5 cm of the bot­
CF3C(O)CH2C(O)CH3 + OH → Products k1 (1) tom and sealed tight between the aluminum flanges, a quartz made
CF3C(O)CH2C(O)CH2CH3 + OH → Products k2 (2) cylinder of 1.5 m length and 0.1 m inner diameter that houses three low-
CF3C(O)CH2C(O)CH(CH3)2 OH Products k3 (3)
+ →
pressure mercury vapor lamps (Philips, TUV 40 W, λmax = 254 nm). The
quartz cylinder is open to ambient air allowing the lamps to be cooled by
a fan placed at one end.
The investigations were performed in two similarly built atmo­
Between experiments the reactor is evacuated by a fore pump.
spheric simulation chambers coupled with different analytical tech­
Cleaning is achieved by filling the reactor up to 900 mbar and evacu­
niques, in situ FTIR spectroscopy in BUW chamber (Wuppertal) and GC-
ating under vacuum and repeating this sequence until no signals of the
FID in LUQCA chamber (Córdoba) using two different photolytic sources
compounds to be studied are present in the GC-FID.
of OH radicals.
Organic species were monitored by a gas chromatograph equipped
The kinetic data are used to infer the environmental implications of
with flame ionization detection (GC-FID) Shimadzu GC-2014B, using an
the OH radicals-initiated oxidation of FDKs such as tropospheric lifetime
Rtx-5 capillary column (fused silica G27, 30 m × 0.25 mm 0.25 μm). Gas
and photochemical ozone creation potential (POCP).
samples were collected from the 405 L reactor, using one of the sampling
ports on the front flanges, by solid phase microextraction (SPME). The
2. Experimental section
SPME technique involves extraction of analytes from the sample matrix
using a silica fiber covered with an absorbent polymer, followed by
Kinetic experiments using the relative-rate method were conducted
thermal desorption of analytes in the chromatograph’s injection port.
in both the BUW chamber and the LUQCA chamber at (990 ± 15) mbar
The fiber composition was chosen according to the compounds studied,
of air and a temperature of (298 ± 3) K. Typical duration of the exper­
namely Divinylbenzene/Carboxen/polydimethylsiloxane (DVB/CAR/
iments ranged from 10 to 35 min, with a 2 min period of homogenisation
PDMS), provided by Supelco, (Bellefonte, PA, USA). The methodology
before irradiation.
used was: irradiation of the mixture target + reference compounds for 1
min, sampling via SPME (exposing the fiber inside the reaction

2
P.L. Lugo et al. Chemosphere 286 (2022) 131562

Table 1 2.4. Materials


Initial mixing ratios in [ppm] (1 ppm = 2.46 × 1013 molecule cm− 3) of each
species used in the experiments. The chemicals used in the experiments had the following purities, as
Compound BUW chamber LUQCA chamber given by the manufacturer, and were used as supplied: synthetic air (Air
[ppm] [ppm] Liquide, 99.999%), 1,1,1-Trifluoro-2,4-pentanedione (Aldrich, 99%),
1,1,1-Trifluoro-2,4-pentanedione 8.4–10.5 5.0–7.5 1,1,1-Trifluoro-2,4-hexanedione (Aldrich, 99%), 1,1,1-Trifluoro-5-
1,1,1-Trifluoro-2,4-hexanedione 7.4–9.2 4.4–6.6 methyl-2,4-hexanedione (Aldrich, 99%), Isobutene (Messer Griesheim,
1,1,1-Trifluoro-5-methyl-2,4- 6.8–8.5 4.4–6.6 99%), 1,3-butadiene (Aldrich 99%), Propene (Messer Griesheim
hexanedione
99.95%), Diethyl ether (Fluka, 99%), NO (Messer Griesheim, 99%) and
Isobutene 20.9–31.3 –
1,3-Butadiene 20.9–31.3 – H2O2 (Interox, 85% w/w). Methyl nitrite was synthesized by the drop­
Propene 20.9–31.3 – wise addition of 50% H2SO4 to a saturated solution of sodium nitrite in
Diethyl ether – 40.7–43.6 methanol and collected and stored in a cooling trap at – 78 ◦ C. IR spectra
NO 20.9–31.3 – of a gaseous methyl nitrite sample shown no impurities were present.
H2O2 13–15
The initial concentrations of each species are summarized in Table 1.

Methyl nitrite 20.9–31.3 –

3. Results and discussion


chamber) for 8 min and thermal desorption of the sample over 2 min in
the injection port. This sequence was repeated until the desired amount The rate coefficients for the reaction between OH radicals and 1,1,1-
of irradiation time was achieved. All determinations were carried out trifluoropentane-2,4-dione (TFP), 1,1,1-trifluorohexane-2,4-dione
under atmospheric conditions. (TFH) or 1,1,1-trifluoro-5-methylhexane-2,4-dione (TFMH) were
In the LUQCA chamber OH radicals were generated by the photolysis determined relative to the consumption of three reference compounds at
of hydrogen peroxide using the mercury lamps: (298 ± 3) K and (990 ± 15) mbar. For the experiments evaluated via
FTIR results, the absorption features centred at 1112 cm− 1, assigned to
H2 O2 + hv→2 OH (7)
ν(CF3), were used to determine the consumption of all FDKs. For refer­
ences, the absorption bands centred at 889, 913 and 908 cm− 1 were used
2.3. Relative-rate method for the quantification of isobutene, propene and 1,3-butadiene, respec­
tively. Using GC-FID, for each FDK experiment no signal overlap was
The relative rate technique relies on the assumption that both the detected and, in the chromatograms, only one peak was observed for
fluorinated diketones (FDK) and the reference compounds are removed each FDK (Fig. 1). The relative ratios kFDK/kref. of each experiment were
solely by reaction with OH radicals: obtained from linear regression analysis where the intercepts of each
regression line were found to be nearly zero. Table 2 summarises the
OH + FDK→Products (8)
experimental results obtained within this study. The relative-rate plots
OH + Reference→Products (9) according to equation (10) are shown in Figs. 2–4. No differences in
kFDK/kref were observed for different initial concentrations or a variation
The direct interaction of the compounds with the OH source chem­ of the total consumption during the experiments (21–40% for TFP,
icals and heterogeneous loss was observed for 5 min previous each 26–56% for TFH and 30–63% for TFMH, respectively).
irradiation and found negligible. Photolysis of the fluorinated diketones The following rate coefficients were determined by averaging the
and reference compounds in the absence of the OH sources was likewise values of the individual experiments:
found to be negligible. Therefore, the relationship between the rate
constants of the reference compound and the FDK can be expressed as: k1-FTIR = (1.3 ± 0.5) × 10− 11 cm3 molecule− 1 s− 1
{ }
kFDK
{ } k1-GC-FID = (1.3 ± 0.2) × 10− 11 cm3 molecule− 1 s− 1
[FDK]0 [Ref ]0
Ln = Ln (10) k2-FTIR = (2.3 ± 0.8) × 10− 11 cm3 molecule− 1 s− 1
[FDK]t kRef [Ref ]t
k2-GC-FID = (2.1 ± 0.8) × 10− 11 cm3 molecule− 1 s− 1

where, [FDK]0, [Ref]0, [FDK]t and [Ref]t are the concentrations of the k3-FTIR = (3.2 ± 1.1) × 10− 11 cm3 molecule− 1 s− 1
fluorinated diketones and reference compound at times t = 0 and t, k3-GC-FID = (3.4 ± 0.8) × 10− 11 cm3 molecule− 1 s− 1

respectively and kFDK and kRef are the rate coefficients of the reactions
(8) and (9), respectively.

Fig. 1. Chromatographic separation of (a) 1,1,1-Trifluoropentane-2,4-dione (TFP) and (b) diethyl ether (reference) under the experimental conditions of the pre­
sent study.

3
P.L. Lugo et al. Chemosphere 286 (2022) 131562

Table 2 in the enolic tautomers. For the TFP-analogue structure pentane-2,4-


Rate coefficients for the reaction of OH radicals with TFP, TFH TFMH experi­ dione (acetyl acetone, Acac) is has been shown that in the gas-phase
mentally determined in this study. more than 90% of the compound are present in the enolic form at
FDKs Technique Reference kFDK/ kOH × 1011 (cm3 room temperature (Nakanishi et al., 1977). The enolic form is stabilized
kref molecule− 1s− 1) due to intramolecular H bonding between the OH moiety and the
TFP Isobutene a 0.29 1.5 ± 0.3 carbonylic oxygen. Zhou et al. (2008) investigated the OH radical
1,1,1-Trifluoro- ± 0.01 initiated oxidation of Acac in the gas-phase and showed that the iden­
2,4-pentanedione FTIR 0.29 1.5 ± 0.3 tified main products (acetic acid, methyl glyoxal and pentane-2,3,
± 0.01
b 4-trione) can be well explained by addition of OH to the C– – C double
Propene 0.44 1.1 ± 0.2
± 0.02 bond in the enol form. Burdett and Rogers (1964) determined the frac­
0.48 1.3 ± 0.2 tion of enol for a series of β-diketones via NMR spectroscopy in the liquid
± 0.02 phase. According to their results the fraction of enol is 81% for Acac and
Average 1.3 ± 0.5 97% for TFP, respectively. However, they do not differentiate between
GC-FID Diethyl 0.98 1.3 ± 0.3
ether c ± 0.04 1.3 ± 0.4
the various enol tautomers possible in the case of asymmetric
0.96 β-diketones.
± 0.03 Fig. 6 presents the region between 1000 cm− 1 and 2000 cm− 1 of the
Average 1.3 ± 0.3 FTIR spectra recorded for the title FDKs in the BUW chamber. In all cases
TFH Isobutene a 0.41 2.1 ± 0.3
it is obvious that the carbonyl absorption is dominated by absorption
1,1,1-Trifluoro- ± 0.01
2,4-hexanedione 0.42 2.1 ± 0.4 features centred on the range 1609–1616 cm− 1 and 1659 - 1664 cm− 1.
± 0.01 Only for TFH and TFMH small absorptions are visible as well around
FTIR Propene b
0.98 2.6 ± 0.4 1840 cm− 1. Based on text-book knowledge the absorption of the C– –O
± 0.02 stretching vibration in saturated, acyclic ketones should be located in
0.88 2.3 ± 0.4
the spectral region between 1705 and 1725 cm− 1. However, intra-
± 0.03
Average 2.3 ± 0.8 molecular H bonding and resonance effects in the enolic tautomers
GC-FID Diethyl 1.70 2.2 ± 0.4 would decrease the bond strength of the C– – O moiety and subsequently
ether c ± 0.07 1.9 ± 0.3 lower the frequency of its absorption band. Therefore, the infrared
1.45
spectra prove the predominance of the enolic tautomers for all investi­
± 0.04
Average 2.1 ± 0.8 gated species in the gas-phase. The features around 1840 cm− 1 suggests
TFMH Propene b
1.08 2.8 ± 0.6 that in the case of TFH and TFMH a small fraction of these compounds is
1,1,1-Trifluoro-5- ± 0.05 present as keto tautomer in the experimental system.
methyl-2,4- FTIR 1.18 3.1 ± 0.8 On the other hand, carbonyl absorption features are shifted to higher
hexanedione ± 0.13
frequencies when halogen atoms are in the α-position. For example, the
1,3- 0.45 3.2 ± 0.8
Butadiene ± 0.05 absorption band of the C– – O stretching vibration is shifted around 40
d
cm− 1 towards higher frequencies in the case of trifluoroacetic acid (TFA)
0.51 3.6 ± 0.7 compared to acetic acid. Thus, the twin-peak profile of the FDK’s
± 0.03
carbonyl absorptions likely results from the coexistence of both enolic
Average 3.2 ± 1.1
GC-FID Diethyl 2.53 3.3 ± 0.7
tautomers in the gas-phase with the band below 1700 cm− 1 being
ether c ± 0.12 3.5 ± 0.6 attributed to enol (b) (Fig. 5). This is also supported by the small shift of
2.67 the twin-peak features toward lower wavenumbers observed when
± 0.10 comparing the TFP spectrum to TFH and TFMH, which is consistent with
Average 3.4 ± 0.8
the increased inductive effect of the alkyl rest R. By assuming the ab­
a − 11
kisobutene+OH = (5.07 ± 0.51) × 10 cm molecule s− 1(Atkinson et al.,
3 − 1
sorption cross sections of C– – O bands in both enols to be similar, the
1997). ratio of their IBIs (= integrated band intensity) indicates rather a pref­
b
kpropene+OH = (2.63 ± 0.20) × 10− 11 cm3 molecule− 1 s− 1(Atkinson et al., erence for enol (a) for all three FDKs. This is in agreement with a study of
1997). Zahedi-Tabrizi and co-workers (Zahedi-Tabrizi et al., 2006) who
c
kdiethyl ether+OH = (1.32 ± 0.20) × 10− 11 cm3 molecule− 1 s− 1(Mellouki et al.,
concluded that, based on DFT calculations, enol (a) (according to Fig. 5)
2015).
d is slightly more stable than enol (b) in the case of TFP. However, the
k1,3-butadiene+OH = (7.00 ± 0.30) × 10− 11 cm3 molecule− 1 s− 1(Ghosh et al.,
2010). authors compared experimental and calculated IR spectra of gaseous
TFP and assigned the bands centred on around 1616 cm− 1 and 1661
cm− 1 to the symmetric and asymmetric C– – C–C–– O stretching rather
where the error represents a combination of the 2σ statistical error of the
than to the C– – O stretching of both enol tautomers. A mechanistic
mean and an additional 15% relative error to cover the uncertainties
investigation of the FDK’s gas-phase oxidation is therefore needed to
derived from the evaluation procedure. To the best of our knowledge
estimate the fraction of both enol tautomers.
there are no previous kinetic studies on FDKs and thus a direct com­
Table 3 compares the values calculated for the reaction of the
parison of the experimentally determined rate coefficients is not
different tautomers of the studied FDKs with OH using the AOPWIN™
possible. However, an excellent agreement is observed between the two
software (EPI Suite™, https://www.epa.gov/tsca-screening-tools/epi-s
detection methods for each of the investigated species.
uitetm-estimation-program-interface) to the experimentally deter­
The FDKs are, as stated in the introduction, β-diketones. Typical for
mined values. This employs the SAR premises elaborated by Kwok and
these compounds is the ability to rearrange the position of double bonds
Atkinson (1995). In the keto form, the reaction of the studied com­
in their molecules, exchanging one carbonyl moiety for an enol group.
pounds with OH radicals might occur via hydrogen atom abstraction at
This particular case of isomerism is called keto-enol tautomerism and
the –CH2– group enclosed between both carbonyl moieties and either
the resulting structures are named tautomers. Fig. 5 shows the possible
the terminal methyl, ethyl or iso-propyl group. The rate coefficients (at
tautomers formed in the keto-enolic equilibrium of FDKs in the gas
room temperature) for the keto form of TFP, TFH and TFMH estimated
phase, taking into account their asymmetric structure. The OH reaction
by AOPWIN™ are 0.18 × 10− 12 cm3 molecule− 1 s− 1, 1.31 × 10− 12 cm3
with these species may thus proceed either via H-abstraction in the keto
molecule− 1 s− 1 and 2.59 × 10− 12 cm3 molecule− 1 s− 1, respectively. The
form or addition of the OH radical to the double bond and H-abstraction
contribution of the addition of OH to the C– – C double bond to the

4
P.L. Lugo et al. Chemosphere 286 (2022) 131562

Fig. 2. Relative-rate plot of 1,1,1-trifluoropentane-2,4-dione.; error bars derived from experimental and evaluation errors.

Fig. 3. Relative-rate plot of 1,1,1-trifluorohexane-2,4-dione; error bars derived from experimental and evaluation errors.

overall rate constant is calculated similarly as 5.94 × 10− 11 cm3 mole­ methyl nitrite photolysis. Therefore, the rate coefficients cannot be
cule− 1 s− 1 for enol (a) and 7.82 × 10− 11 cm3 molecule− 1 s− 1 for enol (b), assigned to one of the enol tautomers and reflect the overall enol reac­
same for all three FDKs. By considering the magnitude of the experi­ tivity for each investigated system.
mentally determined rate coefficients together with the gas-phase IR The experimentally determined rate coefficients are up to a factor of
spectra, there is a clear evidence that the OH initiated oxidation of the 6 smaller than predicted by AOPWIN™. In Table 3 we added, for com­
FDKs is dominated by addition to the C– – C double bond of the enol parison, the calculated and experimental data for pentane-2,4-dione
tautomers. Re-equilibrium between the tautomers is not expected to (Acac) as reported in former studies (Bell et al., 2006; Holloway et al.,
play a significant role due to the much higher reactivity of the C– –C 2005; Messaadia et al., 2015; Zhou et al., 2008). Given that the calcu­
double bond compared to the H atom abstraction in the diketone forms lated kOH value for the enol tautomer of Acac is consistent with the
and the large fraction of enol present in the gas-phase for all investigated experimentally determined rate coefficients, the SAR approach seems to
FDKs. Besides, re-equilibrium processes would result in non-linear reproduce generally the reactivity of the enol’s substitution pattern. This
relative rate plots. Since this behaviour could not be observed for any suggests that the differences between the SAR calculation and our
of the target species, irrespective of the level of consumption and the experimental data are correctly attributed to the inaccuracy of the
detection method, this is a further evidence that re-equilibrium is not estimated contribution due to the presence of halogen atoms, as
important in the experimental system. However, one should note that mentioned in the introduction.
the tentative discrimination of both enol tautomers would be possible According to literature data the bimolecular rate coefficient is in the
solely in the carbonyl absorption region which cannot be used for the range (7.9–9.1) × 10− 11 cm3 molecule− 1 s− 1 for the reaction of OH
evaluation due to the formation of high amounts of NO2 during the radicals with Acac (Bell et al., 2006; Holloway et al., 2005; Messaadia

5
P.L. Lugo et al. Chemosphere 286 (2022) 131562

Fig. 4. Relative-rate plot of 1,1,1-trifluoro-5-methylhexane-2,4-dione; error bars derived from experimental and evaluation errors.

Fig. 5. Keto-enol tautomerism of the fluorinated diketones (R = –CH3 for TFP, R = –CH2CH3 for TFH and R = –CH(CH3)2 for TFMH).

Fig. 6. Infrared spectra of the fluorinated diketones and acetyl acetone in the gas-phase.

et al., 2015; Zhou et al., 2008). Thus, our results show TFP to be a factor OH radicals with 3,3,3-trifluoropropene which is thus about 20 times
of 6–7 less reactive than its hydrogenated analogue structure. This can less reactive than propene (Daranlot et al., 2010). Since the OH reaction
be rationalized in terms of the strong electron-withdrawing effect of the of this species proceeds solely through the addition to the C– – C double
–CF3 group yielding a decrease of electron density associated with the bond one can assign the reduction of reactivity to the strong negative
olefinic bond. Consequently, the enol is expected to be less reactive to­ inductive effect of the adjacent –CF3 substituent upon the C– – C double
wards electrophiles. Given that inductive effects are attenuated with bond. On the other hand, 2,2,2-trifluoroethyl methacrylate (Tovar and
distance from the reactive centre of the molecule, one would intuitively Teruel, 2014) is less reactive towards OH radicals by a factor of about 2
expect enol (a) to be less reactive than enol (b) which is in agreement compared to ethyl methacrylate (Blanco et al., 2006; Ren et al., 2019).
with the SAR calculations. A rate coefficient of 1.3 × 10− 12 cm3 mole­ Given that rate coefficients are only slightly higher for ethyl methac­
cule− 1 s− 1 (González et al., 2015) has been reported for the reaction of rylate than for methyl methacrylate the influence of H atom abstraction

6
P.L. Lugo et al. Chemosphere 286 (2022) 131562

Table 3 Table 5
Comparison of calculated and experimentally determined rate coefficients POCP values calculated in the present work.
values in cm3 molecule− 1 s− 1 for the reaction of OH radicals with FDKs. VOCs εPOCP
Compound Tautomer kH- kaddition koverall kexperimental
Ethene (reference value) CH2=CH2 100a
× × 1011 × 1011 × 1011
abstraction
2 Acac CH3C(O)CH2C(O)CH3 62b
10
TFP CF3C(O)CH2C(O)CH3 23b
TFP keto form 0.179 0.000 0.018 1.3 ± 0.4b TFH CF3C(O)CH2C(O)CH2CH3 29b
enol (a) 0.242 5.944 5.968 TFMH CF3C(O)CH2C(O)CH(CH3)2 34b
enol (b) 0.276 7.821 7.849 a
(Derwent et al., 1998); and c this work.
TFH keto form 1.308 0.000 0.131 2.2 ± 0.8b
enol (a) 1.371 5.944 6.081
enol (b) 1.241 7.821 7.945 In order to estimate tropospheric lifetimes for reactive gases not only
TFMH keto form 2.593 0.000 0.259 3.3 ± 1.0b
the chemical transformations but also the meteorology and geographical
enol (a) 2.656 5.944 6.210
enol (b) 2.415 7.821 8.062
location of the emission sources should be considered. However, since it
Acac keto form 0.730 0.000 0.730 8.7 ± 0.7c is difficult to comprise these conditions at all times, for practical reasons
enol form 0.380 7.821 7.859 it is useful to calculate an approximate tropospheric value. Thus, for
a
TFP: 1,1,1-trifluoro-pentan-2,4-dione; TFH: 1,1,1-trifluoro-hexan-2,4-dione; reactive species those lifetimes are under 1 day (calculated as above) an
TFMH: 1,1,1-trifluoro-hexan-5-methyl-2,4-dione; Acac: Pentane-2,4-dione. average daytime [OH] of 2.5 × 106 molecule cm− 3 is used together with
b
this work, errors as above: 2σ statistical error of the mean and an additional the rate coefficient at 298 K (Calvert et al., 2011). This suggest a
15% relative error; c average between reference data (Holloway et al., 2005), non-uniform mixing for these fluorinated diketones in the troposphere
(Zhou et al., 2008), (Messaadia et al., 2015), error propagation method. and hence no significant contribution to global warming is expected
(Hodnebrog et al., 2013). However, if in the atmospheric oxidation of
the fluorodiketones products conserving CF3 moieties are formed, they
Table 4 might be less reactive possessing a higher GWP than the parent
Atmospheric lifetimes of the FDKs studied
compounds.
regarding to the reaction with OH radicals.
Also of importance for the environmental impact is the acidification
FDKs ƮOH (Hours) potential (AP) every airborne compound containing heterogeneous
TFP 21 atoms possesses (de Leeuw, 1993). Accordingly, values of 0.62, 0.57 and
TFH 13 0.53 are calculated for TFP, TFH and TFMH, respectively. The reference
TFHM 8
value is 1.00 for SO2, hence these results appear quite high. This in­
dicates that these fluorinated carbonyl compounds (FDKs) and their
on the overall rate coefficient is negligible. Hence, a significant reduc­ atmospheric degradation products could be involved in "acid rain"
tion of the C–– C double bond’s reactivity was already observed due to a events with the known consequences for deteriorating water and soil
more distant –CF3 group and one would expect enol (b) of the TFP quality and subsequently affecting the biota and human health (Bouw­
system to be much less reactive than Acac as well. However, besides that man et al., 2002; Granados Sánchez et al., 2010; Huang, 1992). How­
the (-I)-effect of –CF3 substituent reduces the reactivity of the TFP sys­ ever, since the time scale between emission and wash out as acidic
tem comparative to acetyl acetone no further statement can be gained species is not known, these values should be regarded as upper limits for
from the experimentally determined rate coefficient without discrimi­ the acidification potential of the studied compounds.
nation between the individual contribution of both enol tautomers.
No similar information is available for non-halogenated analogues of 3.1.2. Estimated photochemical ozone creation potentials (POCP)
TFH and TFMH. The present work indicates that tautomerism has an Ozone formation in the troposphere is related to health problems
impact on the reactivity of the studied fluorinated β-diketones and hence (Vinikoor-Imler et al., 2014) and damage to vegetation (Ghude et al.,
further investigation are necessary to decipher the ratio between the 2014). As part of this study, the photochemical ozone creation potentials
individual tautomers in the gas phase. were estimated for the three fluorinated diketones, employing a method
designed by Derwent et al. (1998) and Jenkin (1998). The method
correlates the ozone creation potential to the FDKs OH reactivity relative
3.1. Atmospheric implications to ethene by the following equation:

3.1.1. Tropospheric lifetimes, global warming potential and acidification εPOCP = α1 × γs × γβR (1 − α2 × nC ) (11)
potential In equation (11) ε POCP
is the estimated photochemical ozone creation
Calculated tropospheric lifetimes of the FDKs studied in this work potential, α1, α2, β, γS and γR are parameters we calculated combining
with respect to reaction with OH radicals are given in Table 4, consid­ the rate coefficients determined in this work with the information given
ering a 24-h average OH concentration of 1 × 106 molecules cm− 3 in Jenkin (1998) and the last IUPAC recommendation for the rate co­
(Atkinson et al., 1997). efficient for the reaction of C2H4 with OH. Table 5 shows the calculated
These values indicate that atmospheric transport of the fluorinated values for TFP, TFH, TFMH and Acac, for comparison.
carbonyls cannot be excluded, although the area of impact and how far In the case of TFP and Acac the estimated POCPs reflect the
away from the emission sources they travel cannot be estimated solely decreased reactivity toward OH radicals induced by the presence of
from the actual data. On the other hand, in coastal regions the con­ fluorine. In order to make a sound evaluation of the contribution of FDKs
centration of Cl atoms can be higher than 1 × 104 atoms/cm3 (Wing­ to the formation of tropospheric ozone a more thorough analysis is
enter et al., 1996) and also anthropogenic sources such as brick and needed to include their reactivity toward ozone as a consequence of the
ceramic industries, water treatment plants, etc. (Wang et al., 2009, enol tautomerism.
2019) increases locally the level of chlorine atoms which could
contribute to the degradation of these compounds. More, the enolic 4. Conclusions
structure might also favour the reaction with ozone. Further investiga­
tion on the gas-phase reactions of TFP, TFH and TFMH with other at­ This work reports the first determination of the rate constant co­
mospheric oxidants ozone, chlorine atoms and NO3 radicals are under efficients for the gas-phase reaction of OH radicals with 1,1,1-trifluoro-
development.

7
P.L. Lugo et al. Chemosphere 286 (2022) 131562

2,4-pentanedione (TFP), 1,1,1-trifluoro-2,4-hexanedione (TFH) and compounds. Tetrahedron 44, 2997–3004. https://doi.org/10.1016/S0040-4020(88)
90039-7.
1,1,1-trifluoro-5-methyl-2,4-hexanedione at room temperature and at­
Bell, P., Nicovich, J.M., Wine, P.H., 2006. Acetylacetone Photolysis at 248 Nm: Hydroxyl
mospheric pressure. The measurements performed using IR spectros­ Radical Yield and Temperature-dependent Rate Coefficients for the OH +
copy indicated that the keto-enolic tautomerization takes place for all Acetylacetone Reaction.
studied compounds in the gas-phase, the enolic forms where the OH Blanco, M.B., Taccone, R.A., Lane, S.I., Teruel, M.A., 2006. On the OH-initiated
degradation of methacrylates in the troposphere: gas-phase kinetics and formation of
moieties are vicinal to the –CF3 group being probably predominant. pyruvates. Chem. Phys. Lett. 429, 389–394. https://doi.org/10.1016/j.
Under the experimental conditions of this study was not possible to cplett.2006.08.088.
discern between the rate coefficients of the individual tautomers. Bouwman, A.F., Van Vuuren, D.P., Derwent, R.G., Posch, M., 2002. A global analysis of
acidification and eutrophication of terrestrial ecosystems. Water, Air, Soil Pollut.
The lifetime durations calculated upon the present results indicate 141, 349–382. https://doi.org/10.1023/A:1021398008726.
that transport might play a role in the atmospheric fate of the studied Burdett, J.L., Rogers, M.T., 1964. Keto-enol tautomerism in β-dicarbonyls studied by
compound, but no major contribution to global warming is expected. nuclear magnetic resonance Spectroscopy.1 I. Proton chemical shifts and
equilibrium constants of pure compounds. J. Am. Chem. Soc. 86, 2105–2109.
The relatively high POCP values estimated for FDKs, suggest that they https://doi.org/10.1021/ja01065a003.
might contribute to the formation of tropospheric ozone or smog. The Calvert, J.G., Mellouki, A., Orlando, J.J., Pilling, M.J., Wallington, T.J., 2011.
upper limits estimates calculated for their acidification potential suggest Mechanisms of Atmospheric Oxidation of the Oxygenates. Oxford University Press,
Oxford, UK.
that FDKs are involved in acid rain events. Daranlot, J., Bergeat, A., Caralp, F., Caubet, P., Costes, M., Forst, W., Loison, J.-C.,
Further investigations to include the products formed in the reaction Hickson, K.M., 2010. Gas-phase kinetics of hydroxyl radical reactions with alkenes:
with OH radicals and comprehensive analysis of the gas-phase reactions experiment and theory. ChemPhysChem 11, 4002–4010. https://doi.org/10.1002/
cphc.201000467.
with other oxidants as ozone, chlorine atoms and NO3 radicals are still
de Leeuw, F.A.A.M., 1993. Assessment of the atmospheric hazards and risks of new
needed in order to understand the implications related to the presence of chemicals: procedures to estimate “hazard potentials. Chemosphere 27, 1313–1328.
FDKs in the atmosphere. https://doi.org/10.1016/0045-6535(93)90226-U.
De Rosa, M., Arnold, D., Hartline, D., Truong, L., Verner, R., Wang, T., Westin, C., 2015.
Effect of bronsted acids and bases, and lewis acid (Sn2+) on the regiochemistry of
Author contribution the reaction of amines with trifluoromethyl-β-diketones: reaction of 3-aminopyrrole
to selectively produce regioisomeric 1H-Pyrrolo[3,2-b]pyridines. J. Org. Chem. 80,
12288–12299. https://doi.org/10.1021/acs.joc.5b02192.
María Belén Blanco and Mariano Teruel: Conceptualization. Niklas Derwent, R.G., Jenkin, M.E., Saunders, S.M., Pilling, M.J., 1998. Photochemical ozone
Illmann and Iulia Patroescu-Klotz: Methodology and Validation. Pedro creation potentials for organic compounds in northwest Europe calculated with a
Lugo, Vianni Straccia and Cynthia Rivela: Investigation. Pedro Lugo, master chemical mechanism. Atmos. Environ. 32, 2429–2441. https://doi.org/
10.1016/S1352-2310(98)00053-3.
María Belén Blanco, Niklas Illmann and Iulia Patroescu-Klotz: Writing-
Ghosh, B., Park, J., Anderson, K.C., North, S.W., 2010. OH initiated oxidation of 1,3-
Original draft preparation. Peter Wiesen and María Belén Blanco: butadiene in the presence of O2 and NO. Chem. Phys. Lett. 494, 8–13. https://doi.
Supervision. org/10.1016/j.cplett.2010.05.056.
Ghude, S.D., Jena, C., Chate, D.M., Beig, G., Pfister, G.G., Kumar, R., Ramanathan, V.,
2014. Reductions in India’s crop yield due to ozone. Geophys. Res. Lett. 41,
5685–5691. https://doi.org/10.1002/2014GL060930.
Declaration of competing interest
González, S., Jiménez, E., Ballesteros, B., Martínez, E., Albaladejo, J., 2015. Hydroxyl
radical reaction rate coefficients as a function of temperature and IR absorption cross
The authors declare that they have no known competing financial sections for CF3CH=CH2 (HFO-1243zf), potential replacement of CF3CH2F (HFC-
134a). Environ. Sci. Pollut. Res. Int. 22, 4793–4805. https://doi.org/10.1007/
interests or personal relationships that could have appeared to influence
s11356-014-3426-2.
the work reported in this paper. Granados Sánchez, D., López Ríos, G.F., Hernández García, M.Á., 2010. La lluvia ácida y
los ecosistemas forestales. Rev. Chapingo Ser. Cienc. For. Ambiente 16, 187–206.
https://doi.org/10.5154/r.rchscfa.2010.04.022.
Acknowledgments Hodnebrog, Ø., Etminan, M., Fuglestvedt, J.S., Marston, G., Myhre, G., Nielsen, C.J.,
Shine, K.P., Wallington, T.J., 2013. Global warming potentials and radiative
The authors wish to acknowledge to EUROCHAMP 2020, FonCyT, efficiencies of halocarbons and related compounds: a comprehensive review. Rev.
Geophys. 51, 300–378. https://doi.org/10.1002/rog.20013.
CONICET and SECyT, UNC, Argentina. P. L. L. G and V. S. C wish to
Holloway, A.-L., Treacy, J., Sidebottom, H., Mellouki, A., Daële, V., Bras, G.L., Barnes, I.,
acknowledge to CONICET for a doctoral fellowship and support. M. B. B. 2005. Rate coefficients for the reactions of OH radicals with the keto/enol tautomers
and C. B. R. wish to acknowledge to Alexander von Humboldt Founda­ of 2,4-pentanedione and 3-methyl-2,4-pentanedione, allyl alcohol and methyl vinyl
ketone using the enols and methyl nitrite as photolytic sources of OH. J. Photochem.
tion for the support.
Photobiol. Chem., In Honour of Professor Richard P. Wayne 176, 183–190. https://
doi.org/10.1016/j.jphotochem.2005.08.031.
References Huang, H., 1992. Studies of acid rain in the Eastern United States: a review. Int. J.
Environ. Stud. 41, 267–275. https://doi.org/10.1080/00207239208710766.
Ilmi, R., Haque, A., Al-Busaidi, I.J., Al Rasbi, N.K., Khan, M.S., 2019. Synthesis and
Atkinson, R., 1987. A structure-activity relationship for the estimation of rate constants
photophysical properties of hetero trinuclear complexes of tris β-diketonate
for the gas-phase reactions of OH radicals with organic compounds. Int. J. Chem.
Europium with organoplatinum chromophore. Dyes Pigments 162, 59–66. https://
Kinet. 19, 799–828. https://doi.org/10.1002/kin.550190903.
doi.org/10.1016/j.dyepig.2018.10.011.
Atkinson, R., Arey, J., Aschman, S.M., Corchnoy, S.B., Shu, Y., 1995. Rate constants for
Isakova, V., Khlebnicova, T., Lakhvich, F., 2010. Chemistry of fluoro-substituted beta-
the gas-phase reactions of cis-3-hexen-1-ol, cis-3-hexenylacetate, trans-2- hexenal,
diketones and their derivatives. Russ. Chem. Rev. - RUSS CHEM REV-ENGL TR 79,
and linalool with OH and NO3 radicals and O3 at 296 ± 2 K, and OH radical
849–879. https://doi.org/10.1070/RC2010v079n10ABEH004123.
formation yields from the O3 reactions. Int. J. Chem. Kinet. 27, 941–955. https://doi.
Jenkin, M.E., 1998. Photochemical Ozone and PAN Creation Potentials: Rationalisation
org/10.1002/kin.550271002.
and Methods of Estimation. AEA Technology Plc, Report AEAT- 4182/20150/003,
Atkinson, R., Baulch, D.L., Cox, R.A., Hampson, R.F., Kerr, J.A., Rossi, M.J., Troe, J.,
AEA Technology Plc. National Environmental Technology Centre, Culham,
1997. Evaluated kinetic and photochemical data for atmospheric chemistry:
Oxfordshire OX14 3DB, UK.
supplement VI. IUPAC subcommittee on gas kinetic data evaluation for atmospheric
Jenkin, M.E., Valorso, R., Aumont, B., Rickard, A.R., Wallington, T.J., 2018. Estimation
chemistry. J. Phys. Chem. Ref. Data 26, 1329–1499. https://doi.org/10.1063/
of rate coefficients and branching ratios for gas-phase reactions of OH with aliphatic
1.556010.
organic compounds for use in automated mechanism construction. Atmos. Chem.
Babain, V., Romanovskii, V., Starchenko, V., Shadrin, A., Kudinov, G., Podoinitsyn, S.,
Phys. 18, 9297–9328. https://doi.org/10.5194/acp-18-9297-2018.
Revenko, Y., 2002. Extraction of actinide salts and oxides by β-diketones in
Jiménez, E., González, S., Cazaunau, M., Chen, H., Ballesteros, B., Daële, V.,
supercritical or liquid carbon dioxide as applied to decontamination. J. Nucl. Sci.
Albaladejo, J., Mellouki, A., 2016. Atmospheric degradation initiated by OH radicals
Technol. 39, 267–269. https://doi.org/10.1080/00223131.2002.10875459.
of the potential foam expansion agent, CF3(CF2)2CH═CH2 (HFC-1447fz): kinetics
Bareno, V.D.O., Santos, D.S., Frigo, L.M., Mello, D.L. de, Malavolta, J.L., Blanco, R.F.,
and formation of gaseous products and secondary organic aerosols. Environ. Sci.
Pizzuti, L., Flores, D.C., Flores, A.F.C., Bareño, V.D.O., Santos, D.S., Frigo, L.M.,
Technol. 50, 1234–1242. https://doi.org/10.1021/acs.est.5b04379.
Mello, D.L. de, Malavolta, J.L., Blanco, R.F., Pizzuti, L., Flores, D.C., Flores, A.F.C.,
Kwok, E.S.C., Atkinson, R., 1995. Estimation of hydroxyl radical reaction rate constants
2020. An acetal acylation methodology for producing diversity of trihalomethyl-1,3-
for gas-phase organic compounds using a structure-reactivity relationship: an
dielectrophiles and 1,2-azole derivatives. J. Braz. Chem. Soc. 31, 244–264. https://
update. Atmos. Environ. 29, 1685–1695. https://doi.org/10.1016/1352-2310(95)
doi.org/10.21577/0103-5053.20190160.
00069-B.
Bassetti, M., Cerichelli, G., Floris, B., 1988. Substituent effects in keto-enol tautomerism.
Part 3.1 influence of substitution on the equilibrium composition of of β-dicarbonyl

8
P.L. Lugo et al. Chemosphere 286 (2022) 131562

Lintvedt, R.L., Holtzclaw, H.F., 1966. Proton magnetic resonance spectra and electronic Wallington, T.J., Hurley, M.D., Nielsen, O.J., Sulbaek Andersen, M.P., 2004. Atmospheric
effects in substituted 1,3-diketones. J. Am. Chem. Soc. 88, 2713–2716. https://doi. chemistry of CF3CFHCF2OCF3 and CF3CFHCF2OCF2H: reaction with Cl atoms and
org/10.1021/ja00964a018. OH radicals, degradation mechanism, and global warming potentials. J. Phys. Chem.
Mellouki, A., Wallington, T.J., Chen, J., 2015. Atmospheric chemistry of oxygenated 108, 11333–11338. https://doi.org/10.1021/jp046454q.
volatile organic compounds: impacts on air quality and climate. Chem. Rev. 115, Wallington, T.J., Sulbaek Andersen, M.P., Nielsen, O.J., 2015. Atmospheric chemistry of
3984–4014. https://doi.org/10.1021/cr500549n. short-chain haloolefins: photochemical ozone creation potentials (POCPs), global
Messaadia, L., El Dib, G., Ferhati, A., Chakir, A., 2015. UV–visible spectra and gas-phase warming potentials (GWPs), and ozone depletion potentials (ODPs). Chemosphere
rate coefficients for the reaction of 2,3-pentanedione and 2,4-pentanedione with OH 129, 135–141. https://doi.org/10.1016/j.chemosphere.2014.06.092.
radicals. Chem. Phys. Lett. 626, 73–79. https://doi.org/10.1016/j. Wang, L., Ge, M., Wang, W., 2009. Kinetic study of the reaction of chlorine atoms with 3-
cplett.2015.02.032. methyl-3-buten-1-ol. Chin. Sci. Bull. 54, 3808–3812. https://doi.org/10.1007/
Nakanishi, H., Morita, H., Nagakura, S., 1977. Electronic structures and spectra of the s11434-009-0439-1.
keto and enol forms of acetylacetone. Bull. Chem. Soc. Jpn. 50, 2255–2261. https:// Wang, X., Jacob, D.J., Eastham, S.D., Sulprizio, M.P., Zhu, L., Chen, Q., Alexander, B.,
doi.org/10.1246/bcsj.50.2255. Sherwen, T., Evans, M.J., Lee, B.H., Haskins, J.D., Lopez-Hilfiker, F.D., Thornton, J.
Paul, S., Mishra, B.K., Baruah, S.D., Deka, R.C., Gour, N.K., 2020. Atmospheric oxidation A., Huey, G.L., Liao, H., 2019. The role of chlorine in global tropospheric chemistry.
of HFE-7300 [n-C2F5CF(OCH3)CF(CF3)2] initiated by •OH/Cl oxidants and Atmos. Chem. Phys. 19, 3981–4003. https://doi.org/10.5194/acp-19-3981-2019.
subsequent degradation of its product radical: a DFT approach. Environ. Sci. Pollut. Wingenter, O.W., Kubo, M.K., Blake, N.J., Smith, T.W., Blake, D.R., Rowland, F.S., 1996.
Res. Int. 27, 907–920. https://doi.org/10.1007/s11356-019-06975-1. Hydrocarbon and halocarbon measurements as photochemical and dynamical
Ren, Y., Cai, M., Daele, V., Mellouki, W., 2019. Rate coefficients for the reactions of OH indicators of atmospheric hydroxyl, atomic chlorine, and vertical mixing obtained
radical and ozone with a series of unsaturated esters. Atmos. Environ. 200, 243–253. during Lagrangian flights. J. Geophys. Res. Atmospheres 101, 4331–4340. https://
https://doi.org/10.1016/j.atmosenv.2018.12.017. doi.org/10.1029/95JD02457.
Tovar, C.M., Teruel, M.A., 2014. Gas-phase kinetics of OH radicals reaction with a series Zahedi-Tabrizi, M., Tayyari, F., Moosavi-Tekyeh, Z., Jalali, A., Tayyari, S.F., 2006.
of fluorinated acrylates and methacrylates at atmospheric pressure and 298 K. Structure and vibrational assignment of the enol form of 1,1,1-trifluoro-2,4-
Atmos. Environ. 94, 489–495. https://doi.org/10.1016/j.atmosenv.2014.05.057. pentanedione. Spectrochim. Acta. A. Mol. Biomol. Spectrosc. 65, 387–396. https://
Vinikoor-Imler, L.C., Owens, E.O., Nichols, J.L., Ross, M., Brown, J.S., Sacks, J.D., 2014. doi.org/10.1016/j.saa.2005.11.019.
Evaluating potential response-modifying factors for associations between ozone and Zhou, S., Barnes, I., Zhu, T., Bejan, I., Albu, M., Benter, T., 2008. Atmospheric chemistry
health outcomes: a weight-of-evidence approach. Environ. Health Perspect. 122, of acetylacetone. Environ. Sci. Technol. 42, 7905–7910. https://doi.org/10.1021/
1166–1176. https://doi.org/10.1289/ehp.1307541. es8010282.

You might also like