You are on page 1of 8

pubs.acs.

org/JPCA Article

Rate Coefficients for the Gas-Phase Reactions of Nitrate Radicals


with a Series of Furan Compounds
Fatima Al Ali, Cécile Coeur,* Nicolas Houzel, Houceine Bouya, Alexandre Tomas,
and Manolis N. Romanias
Cite This: https://doi.org/10.1021/acs.jpca.2c03828 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
Downloaded via UNIV DO ESTADO DO RIO DE JANEIRO on November 18, 2022 at 18:02:25 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: The atmospheric reaction of a series of furan compounds (furan (F), 2-


methylfuran (2-MF), 3-methylfuran (3-MF), 2,5-dimethylfuran (2,5-DMF), and 2,3,5-
trimethylfuran (2,3,5-TMF)) with nitrate radical (NO3) has been investigated using the
relative rate kinetic method in the CHamber for the Atmospheric Reactivity and the
Metrology of the Environment (CHARME) simulation chamber at the laboratoire de Physico-
Chimie de l’Atmosphere (LPCA) laboratory (Dunkerque, France). The experiments were
performed at (294 ± 2) K atmospheric pressure and under dry conditions (relative humidity,
RH < 2%) with proton transfer mass reaction−time of flight−mass spectrometer (PTR-ToF-
MS) for the chemical analysis. The following rate coefficients (in units cm3 molecule−1 s−1)
were determined: furan, k(F) = (1.51 ± 0.38) × 10−12, 2-methylfuran, k(2‑MF) = (1.91 ± 0.32) ×
10−11, 3-methylfuran, k(3‑MF) = (1.49 ± 0.33) × 10−11, 2,5-dimethylfuran, k(2,5‑DMF) = (5.82 ±
1.21) × 10−11, and 2,3,5-trimethylfuran, k(2,3,5‑TMF) = (1.66 ± 0.69) × 10−10. The uncertainty
on the measured rate coefficient (ΔkFC) includes both the uncertainty on the measurement
and that on the rate coefficient of the reference molecule. To our knowledge, this work represents the first determination for the rate
coefficient of the 2,3,5-TMF reaction with NO3. This work shows that the reaction between furan and methylated furan compounds
with nitrate radical (NO3) is the dominant removal pathway during the night with lifetimes between 0.5 and 55 min for the studied
molecules.

1. INTRODUCTION identified as major constituents of the volatile organic


Biomass burning (BB) is the burning of living and dead compounds (VOCs) emitted during wildfire burning released
vegetation, which includes the human-initiated burning for from the pyrolysis of cellulose,9−12 comprising about 30% of
land clearing and land-use change, natural fires, and domestic the total number of VOCs identified in plumes, with significant
heating. Due to the increasing population rate and the extra atmospheric concentration levels ranging between 68 and 118
demand for food, the rate of land clearing and land use ppbv relative to 1 ppmv of CO formation. The concentration
(directly related to wildfires and BB events)1,2 has increased of VOCs emitted from biomass burning differs depending on
dramatically during the last few decades,3 and increasing trends the type of biomass and fire conditions. However, Stockwell et
al. (2015) characterizing the emissions from cooking fires,
are projected over the near future.4 BB emissions play a key
peat, crop residue, and other fuels showed that for the different
role in the oxidative capacity of the atmosphere. During the
biomass fuels, the furan emission factor (≅0.1−2.3 g Kg−1)
daytime, BB plumes can have OH concentrations 5−10 times
emitted was higher than that of phenolic compounds (≅0.05−
higher than background air,5 while an increase of more than
0.8 g Kg−1) but smaller than those of aromatic hydrocarbons
120% of the total OH reactivity during BB events has been
(≅0.2−5 g Kg−1).13 Furans have been demonstrated to
reported.6 During the nighttime, it has been recently shown
contribute significantly (≅20%) to overall NO314 and OH15
that the mixing of smoke-derived O3 with NOx in a BB plume
reactivity of emissions from the burning of common wildfire
leads to the production of NO3 radicals (the dominant
and residential fuels. For comparison, Decker et al. (2021)
nighttime oxidant) with a rate of approximately 1 ppbv h−1.77
showed in the FIREX-AQ aircraft data that the contribution of
As a consequence, significant nighttime chemical trans-
formations potentially occur inside the BB plume that could
strongly impact the oxidative capacity of the atmosphere and Received: June 2, 2022
air quality.8 Yet, these pathways are poorly explored in the Revised: October 17, 2022
literature.
The impact of BB emissions on air quality and climate is
related to the complex mixture of gases and particles released
into the atmosphere. Quite recently, furans have been

© XXXX American Chemical Society https://doi.org/10.1021/acs.jpca.2c03828


A J. Phys. Chem. A XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry A pubs.acs.org/JPCA Article

furan compounds to OH reactivity (25%) is larger than the de l’Atmosphere (LPCA) in Dunkerque, France. CHARME is
contribution of phenolic compounds (16.4%) but smaller than a 9.2 m3 evacuable cylinder (diameter, ≈1.7 m; length, ≈4.0 m;
that of alkenes (33%). However, NO3 reactivity is mainly and surface to volume ratio, ≈3.5 m−1) made from stainless
governed by phenolic compounds (64.4%), which is much steel (304 L) and electropolished. Four fans (stainless steel;
larger than the contribution of furan compounds (19%) and diameter 50 cm) are located at the bottom allowing the fast
alkenes (5%).14 Other sources of furans include waste homogenization of the reaction mixture. The humidity and
incineration, fuel and biofuel combustion, temperature-based temperature within the chamber were monitored by a
industrial processes,16 and natural sediment and water combined T/RH probe (HMT334 series transmitters,
samples,17 and are toxic species and have direct adverse effects VAISALA). CHARME is coupled to a pump (Cobra
on human health.18 Moreover, being promising future NC0100-0300B) to vacuum the pressure down to 0.4 mbar
alternative biofuels, 2-methylfuran and 2,5-dimethylfuran are and to an air generator (Parker Zander KA-MT 1-8)
expected to highly increase in the upcoming years.19,20 They connected to a compressor (SLM-S 7.5�Renner SCROLL-
are highly reactive compounds rapidly removed from the Line) to fill the chamber with purified and dry air. The time
atmosphere due to their gas-phase oxidation. Interestingly, a required to evacuate and reach the atmospheric pressure in the
recent experimental and model study showed that during the chamber is around 1 h. CHARME is described in detail
daytime, furan degradation drives the chemistry inside a fresh elsewhere.31
biomass plume,15 forming secondary pollutants that will 2.2. Experimental Methods. In the present work, the
eventually lead to tropospheric ozone (O3) and secondary initial mixing ratios for reactants were in the range 500−1200
organic aerosol (SOA) formation. Up to now, the contribution ppbv for both furans and reference compounds. The injection
of furan degradation on the nighttime tropospheric chemistry of the organics into the chamber was performed using a glass
is poorly addressed. impinger in which known liquid volumes of the compounds
The atmospheric oxidation of furan compounds can be were under a low flow of gently heated purified air (2−10 L
governed by OH, NO3, O3, and Cl. There are several studies in min−1).
the literature reporting the tropospheric degradation of furan Nitrate radicals were formed in the chamber from the
and methylated furans by OH,21−23 chlorine,21,24 and thermal decomposition at (243 ± 5) K of dinitrogen pentoxide
NO321,25,26 radicals under atmospheric pressure and room (N2O5) under a flow of gaseous nitrogen (at ≅100 mL min−1)
temperature. All studies showed the formation of unsaturated according to reaction 1
dicarbonyl products as major primary products from the
N2O5 NO3 + NO2 (1)
oxidation of methylated furan compounds. According to the
presented studies on the degradation mechanism of furan N2O5 was first synthesized in an S-shaped tube (length = 2.5
compounds by NO3 radical, the major reaction pathway occurs m; diameter = 2 cm) from the gas-phase reaction of nitrogen
by NO3 addition to carbon C-2 or carbon C-5 on the ring due monoxide (NO) with ozone (O3) according to reaction 2
to the formation of resonance-stabilized adducts. As a result, followed by reactions 3 and 4
unsaturated dicarbonyls are formed from the loss of NO2,
rearrangement of the intermediate, and ring opening.25,26 NO + O3 NO2 + O2 (2)
The very scarce kinetic data existing in the literature indicate
that the NO3-initiated oxidation of this class of compounds is NO2 + O3 NO3 + O2 (3)
an extremely fast process and, consequently, their correspond-
ing role in nighttime chemistry is expected to be dominant and NO2 + NO3 N2O5 (4)
more significant than that during the daytime. Methylated Ozone was produced from O2 using a Corona discharge ozone
furan compounds (2-MF, 3-MF, 2,5-DMF, 2,3,5-TMF) are an generator (Air Tree Ozone Technology, C-Lasky C-010-DTI).
important part of this family released during these fire events.27 Dinitrogen pentoxide crystals were collected at 193 K in a glass
At the same time, unlike their daytime atmospheric reactivity cold trap introduced in a liquid nitrogen/ethanol bath and
with hydroxyl radical, their nighttime reactivity with NO3 stored at the same temperature in a freezer for few days before
radical is poorly studied in the literature. Moreover, the few being used. N2O5 was continuously injected into the chamber
studies highlighting it reveals fast reactions.28−30 Thus, it is until ≥98% consumption of both the furan and the reference
important to study these reactions to understand furan compounds. However, only until 50−80% consumption of 2-
chemistry and its impact on the atmosphere. MF was retained in the relative rate plots since the plots are
The aim of the present study is to determine the rate nonlinear beyond these limits (see Figure 1).
constants for the reactions of a series of furan compounds The VOC concentrations were continuously recorded
(furan (F), 2-methylfuran (2-MF), 3-methylfuran (3-MF), 2,5- (every 10 s) with a PTR-ToF-MS (proton transfer reaction−
dimethylfuran (2,5-DMF), and 2,3,5-trimethylfuran (2,3,5- time of flight−mass spectrometer 1000, IONICON Analytik,
TMF)) with NO3 radical. The rate coefficients were measured Innsbruck, Austria). The air from the chamber was sampled at
using the relative-rate method, and their atmospheric lifetimes 50 mL min−1 through an inlet polyether ether ketone (PEEK)
are determined. capillary line of 100 cm heated at 333 K to reduce the loss of
the VOCs during their sampling and then ionized in the drift
2. EXPERIMENTAL SECTION tube (E = 600 V cm−1; E/N = 136 Td) operating at 2.2 mbar
2.1. Description of the Chamber. The experiments were and 333 K. The organics were monitored from the intensity of
performed in the dark at room temperature (294 ± 2 K) in their parent ion [M + H]+: F (m/z 69.1), 2-MF and 3-MF (m/
zero air at atmospheric pressure (760 Torr) and dry conditions z 83.1), 2,5-DMF (m/z 97.1), 2,3,5-TMF (m/z 111.1), α-
(RH < 2%) in the CHamber for the Atmospheric Reactivity terpinene (α-T), γ-terpinene (γ-T), α-pinene (α-P) and β-
and the Metrology of the Environment (CHARME) indoor pinene (β-P) (m/z 137.2), 2-methyl-2-butene (2-MB) (m/z
simulation chamber located at Laboratoire de Physico-Chimie 71.1), and 2,3-dimethyl-2-butene (DMB) (m/z 85.1).
B https://doi.org/10.1021/acs.jpca.2c03828
J. Phys. Chem. A XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry A pubs.acs.org/JPCA Article

FC + NO3 products(kFC) (5)

reference + NO3 products(kRef ) (6)

Kinetic treatment of reactions 5 and 6 leads to eq I, that


represents the FCs and reference compounds changes as a
function of time
[FC]0 kFC [Ref]0
ln k wt = × ln
[FC]t kRef [Ref]t (I)

where kFC and kRef are the rate coefficients for the reactions of
the furan-based compound of interest and the reference with
NO3, respectively. [FC]0, [FC]t, [Ref]0, and [Ref]t are the
concentrations of the furan-based compound and of the
reference at time 0 and t, respectively. kw is the wall loss rate
coefficient, which was taken into account for furan (kw = (5−
7) × 10−5 s−1) while it was neglected for all other organic
Figure 1. Kinetic data plots for the reaction of 2-methylfuran with compounds (kw < 5 × 10−5 s−1). The dilution effect was
nitrate radicals at 294 ± 2 K using 2-MB (black squares), γ-T (red negligible as the signal of acetonitrile remained constant
triangles), and DMB (blue cross) as reference compounds.
throughout the experiment. Plots of the kinetic data of eq I
should yield a straight line with zero intercept and a slope (b)
The wall losses of organic compounds were determined in which is the rate coefficient ratio kFC/kRef.
the dark from the decay of their concentrations before N2O5 The references chosen fulfill several requirements as follows.
introduction. The cleanliness of the reaction chamber was The two compounds (studied and reference compounds)
evaluated in all of the experiments by measuring the should have comparable rate coefficients (0.2 < kFC/kRef < 5).
background levels of the monitored ions for both furan and For each studied furan compound, two to three reference
reference compounds for a time range of 10−15 min. compounds were used and at least three experiments were
Background concentrations were negligible compared to the performed relative to each reference at increasing concen-
initial experimental concentrations. Acetonitrile was injected in trations of the organic compounds. The references used in this
the chamber as a dilution tracer, and the signal was followed by study and their rate coefficients with nitrate radicals (in units
a PTR-MS-TOF.31 of cm3 molecule−1 s−1) were k(α‑terpinene) = (1.82 ± 1.04) ×
Preliminary tests were performed for each VOC to control 10−10,32k(γ‑terpinene) = (2.94 ± 0.81) × 10−11,32k(α‑pinene) = (6.20
the linearity of the PTR-MS signal with respect to the injected ± 1.42) × 10 − 1 2 , 3 3 k ( β ‑ p i n e n e ) = (2.50 ± 0.69) ×
VOC concentrations. Furthermore, the possible contribution 10−11,33k(2‑methyl‑2‑butene) = (9.38 ± 0.47) × 10−12,34 and
of degradation products to the ion peaks used to monitor the k(2,3‑dimethyl‑2‑butene) = (4.50 ± 0.40) × 10−11.35 Except 2,3-
furan or reference compounds was also evaluated and was dimethyl-2-butene, the rate coefficient values of reference
found to be negligible. In these experiments, each compound molecules were IUPAC recommended. In the case of 2,3-
of interest was individually introduced in the chamber and left dimethyl-2-butene, the rate coefficient was retrieved by the
to react with NO3 until its complete consumption. The work of Lancar et al. (1991).35 The agreement of the rate
products formed resulted in various mass peaks in the PTR- coefficient measured relative to the different references gives
MS, but in all cases, they did not interfere with the mass peaks confidence in the values of reference rate coefficients chosen.
selected to monitor the furans or reference compounds. In the following sections, the following abbreviations are used
2.3. Materials. The compounds used in this study, their to denote the reference compounds used: α-T for α-terpinene,
manufacturer, and stated purity were O2 (Praxair, 99.5%), β-T for β-terpinene, α-p for α-pinene, β-P for β-pinene, 2-MB
nitrogen monoxide (Praxair, 99%), furan (>99%, Sigma- for 2-methyl-2-butene, and DMB for 2,3-dimethyl-2-butene.
Aldrich), 2-methylfuran (99%, Sigma-Aldrich), 3-methylfuran 2.5. Error Analysis. In this section, a detailed explanation
(98%, Sigma-Aldrich), 2,5-dimethylfuran (99%, Sigma-Al- is given for the potential uncertainties obtained from our
drich), 2,3,5-trimethylfuran (99%, Sigma-Aldrich), α-terpinene kinetic measurements. The standard deviation arising from the
(>95%, Sigma-Aldrich), γ-terpinene (97%, Sigma-Aldrich), α- least squares linear regression of the data (σ) resembles the
pinene (98%, ACROS organics), β-pinene (>94%, TCI), 2- uncertainty of the slope. The uncertainty of the determined
methyl-2-butene (>99%, Sigma-Aldrich), 2,3-dimethyl-2-bu- rate coefficients included the error on the measurement and
tene (98%, Sigma-Aldrich), and acetonitrile (Sigma-Aldrich, ≥ the uncertainty on the value of the reference rate coefficient as
99.5%). follows. Taking into consideration the slope (b), the known
2.4. Relative Rate Kinetic Method. The rate coefficients kRef and their uncertainties (σ), and ΔkRef, the uncertainty on
for the gas-phase reaction of furan compounds with NO3 the measured rate coefficient (ΔkFC) is calculated according to
radicals were determined using the relative-rate kinetic eq II.36
method, i.e., comparing the decay rates of furan compounds
(FCs) and reference compounds. This method permits kFC = × kRef + b × kRef (II)
determining the unknown rate coefficient of a compound
relative to a compound with a known NO3 rate coefficient. The The final average rate coefficient kaverage and its uncertainty
furan of interest and reference compound react simultaneously Δkaverage (standard error of the mean) were calculated using
with NO3. eqs III and IV
C https://doi.org/10.1021/acs.jpca.2c03828
J. Phys. Chem. A XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry A pubs.acs.org/JPCA Article

Table 1. Summary of All of the Rate Coefficients for the Reactions of Furan Compounds with NO3

a
Indicated error is twice the standard deviation arising from the linear regression analysis. bExpressed in cm3 molecule−1 s−1. cIndicated error
calculated using eq II. dAverage rate coefficients obtained from all reference measurements. eAverage error according to eq IV.

n
i=1
ki the two or three references are in accordance (relative
kaverage = discrepancy lower than 10%) and lead to the following average
n (III)
values (in cm3 molecule−1 s−1): k(F) = (1.51 ± 0.38) × 10−12,
n
( ki)2 k(2‑MF) = (1.91 ± 0.32) × 10−11, k(3‑MF) = (1.49 ± 0.33) ×
i=1
kaverage = 10−11, k(2,5‑DMF) = (5.82 ± 1.21) × 10−11 and k(2,3,5‑TMF) = (1.66
n (IV)
± 0.69) × 10−10. These rate coefficients are compared to the
where ki and Δki represent the rate coefficient and the available literature data.
uncertainties obtained in each experiment and n represents the 3.1. Trends in Reactivity. Comparing the rate coefficient
number of experiments.36 obtained for furan with those of 2-MF and 3-MF, the addition
of a methyl group on the furan ring increases the rate
3. RESULTS AND DISCUSSION coefficient by ∼1 order of magnitude. This could be attributed
The experimental conditions and the results are given in Table to the positive inductive effect of the methyl group, favoring
1. Furan and reference concentrations ranged from ∼300 to the NO3 addition because of the increase of the ring’s electron
∼1400 ppbv. The relative rate plots in the form of eq I for the density. However, the position of the methyl group whether on
experiments of 2-MF are presented in Figure 1. The plots of C-2 or C-3 did not have an impact on the rate coefficient,
the experiments for the other studied molecules (furan, 3- suggesting a negligible effect of the methyl group position on
methylfuran, 2,5-dimethylfuran, and 2,3,5-trimethylfuran) are
the rate coefficient. Comparing single-alkylated furan rate
presented in Figures S1−S4 in the Supporting Information.
The obtained plots show good linearity with near zero coefficients (2-MF and 3-MF) with that of 2,5-dimethylfuran
intercept. Each figure includes the experimental data of all of shows an increase by a factor of ∼3 attributed again to the
the experiments. increase of the positive inductive effect on the ring. The same
The slopes (kFC/kRef) derived from the plots, the determined trend is observed when comparing the NO3 rate coefficients of
NO3 rate coefficients (kFC) for each experiment, and the kFC 2,5-dimethylfuran and 2,3,5-trimethylfuran. This was also
averages are displayed in Table 1.36 The rate coefficients using validated in another literature study.29
D https://doi.org/10.1021/acs.jpca.2c03828
J. Phys. Chem. A XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry A pubs.acs.org/JPCA Article

Table 2. Comparison of the Average Rate Coefficient Measured in This Study with Those in the Literaturea
compound kFC (average) (this work) cm3 molecule−1 s−1 kFC (literature) cm3 molecule−1 s−1 setup/technique
furan (1.51 ± 0.38) × 10 −12
(1.40 ± 0.20) × 10−12 RR-ASC/GC-FID28
(9.98 ± 0.62) × 10−13 RR-FT/GC-FID29
(1.3 ± 0.2) × 10−12 AR-DFT/LIF37
(1.50 ± 0.23) × 10−12 RR-ASC/FTIR30
2-methylfuran (1.91 ± 0.32) × 10−11 (2.57 ± 0.17) × 10−11 RR-FT/GC-FID29
(2.37 ± 0.55) × 10−11 RR-ASC/FTIR30
3-methylfuran (1.49 ± 0.33) × 10−11 (2.86 ± 0.06) × 10−11 RR-FT/GC-FID29
(1.31 ± 0.46) × 10−11 RR-ASC/GC-FID38
(1.26 ± 0.18) × 10−11 RR-ASC/GC-FID, FTIR21
2,5-dimethylfuran (5.82 ± 1.21) × 10−11 (5.78 ± 0.34) × 10−11 RR-FT/GC-FID29
(1.10 ± 0.33) × 10−10 RR-ASC/FTIR30
2,3,5-TMF (1.66 ± 0.69) × 10−10 b

a
Abbreviations used in the table: Relative rate (RR) method, absolute rate (AR) method, atmospheric simulation chamber�Teflon or glass made
(ASC), discharge flow tube (DFT), flow tube (FT), laser-induced fluorescence (LIF), gas chromatography-flame ionization detection (GC-FID),
Fourier transform infrared (FTIR) spectroscopy. bNo data.

Table 3. Atmospheric Lifetimes of Furan Compounds with Respect to Their Reaction with Nitrate Radicals (NO3), Hydroxyl
Radicals (OH), and Ozone (O3)
τ(NO3) (min)a,b τ(NO3) (min)c
compound (nighttime) (daytime) τ(OH) (min)d τ(O3) (min)e τ(Cl) (min)f
44 45
furan 55 1103 199 8547(∼6 days) 8160 (∼5 days)43
2-methylfuran 4 87 11422 4380 (∼3 days)43
3-methylfuran 6 112 9522 1084 (∼1 day)38 4200 (∼3 days)43
2,5-dimethylfuran 1 29 6622 5441 3480 (∼2 days)43
2,3,5-trimethylfuran 0.5 10
a
This work. b[NO3] = 2 × 108 molecules cm−3 (average nighttime concentration).39 c[NO3] = 1 × 107 molecules cm−3 (average daytime
concentration).39 d[OH] = 2 × 106 12 h daytime average.40 e[O3] = 7.5 × 1011 molecules cm−3.41 f[Cl] = 1 × 104 molecules cm−3.46

3.2. Comparison with Literature Data. Literature FID and Fourier transform infrared (FTIR) spectroscopy.22 A
studies measuring the rate coefficients of the reactions of study determined the rate coefficient for the reaction of furan,
furan compounds with NO3 are scarce. Table 2 summarizes the 2-MF, and 2,5-DMF with NO3.30 They also used the relative
rate coefficients kFC obtained in this study and in the literature rate method in a 7300 L indoor simulation chamber by in situ
with different techniques used in each study. Atkinson et al. FTIR spectroscopy. Several references were used to measure
performed relative rate room temperature kinetic measure- each rate coefficient (for furan: cyclohexene, α-P, camphene,
ments to study the reaction of furan with NO3 radicals inside a and α-angelicalactone; for 2-MF: 2-carene, 2,3-DMB, α-P,
4000 L atmospheric pressure simulation chamber coupled with pyrrole, and 2,5-DMF; for 2,5-DMF: 2-carene, 2,3-DMB, α-P,
a gas chromatography-flame ionization detector (GC-FID). pyrrole, and 2-MF).
Trans-2-butene was used as a reference compound for the For furan, all literature rate coefficients are in good
furan rate coefficient measurement.28 Kind et al. performed agreement with the value determined in the present study.
relative rate kinetic measurements inside a low-pressure flow Besides the value of Kind et al., and considering the
tube reactor, coupled with a GC-FID for the detection of experimental uncertainties, our measurement shows good
organic species.29 Trans-butene was used as a reference agreement with those provided in the literature, with
compound to determine the furan reaction rate coefficient, differences below 15%. The reasons for the lower value
while 2,3-dimethyl-2-butene was used as a reference for 2- obtained by Kind et al. are unclear. For the 2-MF + NO3
methylfuran, 3-methylfuran, and 2,5-dimethylfuran measure- reaction, the present work shows a satisfactory agreement with
ments. In the study of Cabanas et al., the rate coefficient was the only two studies in the literature with discrepancies below
measured as a function of temperature inside a low-pressure 25%. Regarding 3-MF, the present work is consistent with
(1.1−1.5 Torr) discharge flow tube reactor coupled with LIF Alvarado et al.38 and Tapia et al.,21 to within 15%, while the
for the detection of NO3 radicals. Experiments were carried out rate coefficient obtained by Kind et al. is about two times
under pseudo-first-order conditions, i.e., following nitrate higher than the present one. As for furan, the reasons for the
radical decay with time to determine the rate coefficient of disagreement are unclear. Regarding 2,5-DMF, there are only
the furan reaction.37 Alvarado et al. measured the 3-MF + NO3 two literature studies determining its rate coefficient: Kind et
rate coefficient in an atmospheric pressure simulation chamber al. report a measurement in excellent agreement with that
coupled with a GC-FID at room temperature using 2-methyl- presented in this work; however, Newland et al.30 gave a 100%
2-butene as a reference compound.38 This rate coefficient was higher rate coefficient. The reason for this discrepancy is
also measured by Tapia et al. using a 500 L atmospheric unknown, but for the common reference (2,3-dimethyl-2-
pressure and room temperature smog chamber. 2-Methyl-2- butene) between the two studies, the rate coefficient used by
butene and α-pinene were used as reference molecules, and the Newland et al. is ≅1.3 times higher than the value used by us.
analysis of the kinetic profiles was achieved by deploying GC- To the best of our knowledge, this work represents the first
E https://doi.org/10.1021/acs.jpca.2c03828
J. Phys. Chem. A XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry A pubs.acs.org/JPCA Article

work giving the rate coefficient for the reaction between the
2,3,5-trimethylfuran reaction with NO3.

*
ASSOCIATED CONTENT
sı Supporting Information
3.3. Atmospheric Implications. From the rate coef- The Supporting Information is available free of charge at
ficients measured in this study, the atmospheric lifetimes of https://pubs.acs.org/doi/10.1021/acs.jpca.2c03828.
furan and methylated furan compounds with respect to their
reaction with NO3 radicals can be determined. Table 3 Kinetic data plots for the reaction of furan with nitrate
includes lifetimes calculated in this work compared to lifetimes radicals at 294 ± 2 K using α-P (green circles) and β-P
due to reactions with hydroxyl radical, O3, and chlorine (Cl) (pink triangles) as reference compounds (Figure S1);
atoms. Assuming an average NO3 nighttime concentration of 2 kinetic data plots for the reaction of 3-methylfuran with
× 108 molecules cm−3,39 the calculated lifetimes of the nitrate radicals at 294 ± 2 K using 2-MB (black squares),
investigated compounds are in the range of 0.5−55 min. The γ-terpinene (red triangles) and α-P (green circles) as
lifetimes of these compounds due to their reactions with OH, reference compounds (Figure S2); kinetic data plots for
O3, or Cl were calculated using a 12 h daylight average OH the reaction of 2,5-dimethylfuran with nitrate radicals at
294 ± 2 K using γ-T (red triangles) and DMB (blue
concentration of 2 × 106 molecules cm−3,40 an ozone
crosses) as reference compounds (Figure S3) (PDF)
concentration of 7.5 × 1011 molecules cm−3,41 an average Cl
global concentration of 1 × 104 molecules cm−3, and the
corresponding OH, O3, and Cl rate coefficients. The lifetimes
were in the range of 66−199, 54−8547, and 3480−8160 min
■ AUTHOR INFORMATION
Corresponding Author
with OH, O3, and Cl, respectively. Therefore, the nighttime Cécile Coeur − Laboratoire de Physico-Chimie de
reaction with NO3 radicals is anticipated to be the principal l’Atmosphère, Université du Littoral Côte d’Opale,
sink for these compounds. The reaction with ozone is Dunkerque 59140, France; orcid.org/0000-0002-5643-
considered negligible for furan and 3-methylfuran, but it 5148; Phone: +33 328 23 76 42; Email: coeur@univ-
could be competitive with OH radical in the case of 2,5- littoral.fr
dimethylfuran. For the reaction of these furanoids with NO3
during the daytime, the lifetimes were calculated using the Authors
NO3 daytime concentration of 1 × 107 molecules cm−3.39 The Fatima Al Ali − Laboratoire de Physico-Chimie de
calculated lifetime was in the range 10−1103 min, and thus, l’Atmosphère, Université du Littoral Côte d’Opale,
the competition between the reaction with NO3 and OH Dunkerque 59140, France; Institut Mines Télécom Nord
during the daytime can be important for 2,5-DMF and 2,3,5- Europe, Univ. Lille, Center for Energy and Environment, F-
TMF. This aspect was also observed in the reaction of 59000 Lille, France
catechols (1,2-dihydroxybenzenes) with nitrate and hydroxyl Nicolas Houzel − Laboratoire de Physico-Chimie de
radicals.42 Regarding the reaction with chlorine atoms, it is not l’Atmosphère, Université du Littoral Côte d’Opale,
significant compared to other reactions; however, it will be Dunkerque 59140, France
much more important in coastal areas with higher Cl Houceine Bouya − Laboratoire de Physico-Chimie de
concentrations.43 Yet, it is worth emphasizing that, in fire l’Atmosphère, Université du Littoral Côte d’Opale,
plumes, where the concentration of nitrate radicals highly Dunkerque 59140, France
increases, which may be fast (>0.5 ppbv h−1),7 the nighttime Alexandre Tomas − Institut Mines Télécom Nord Europe,
Univ. Lille, Center for Energy and Environment, F-59000
reactivity of these compounds dominates with a lifetime in the
Lille, France; orcid.org/0000-0002-0125-581X
order of minutes. Moreover, temperature-dependent studies
Manolis N. Romanias − Institut Mines Télécom Nord Europe,
are needed since the temperature in the plume could be higher
Univ. Lille, Center for Energy and Environment, F-59000
than ambient temperature. The present results will help in Lille, France; orcid.org/0000-0002-9049-0319
better understanding the relevant chemistry in biomass
burning plumes. It is clear that methylated furans are expected Complete contact information is available at:
to be mainly present in fresh BB plumes; their tropospheric https://pubs.acs.org/10.1021/acs.jpca.2c03828
oxidation will result in the formation of products in the gas and
perhaps the particulate phase anticipated to be present in aged Notes
BB plumes. The authors declare no competing financial interest.

4. CONCLUSIONS ■ ACKNOWLEDGMENTS
This work was supported by the CaPPA project (Chemical and
In the present work, five heteroaromatic furan-based Physical Properties of the Atmosphere) funded by the French
compounds were investigated relative to their nighttime National Research Agency (ANR-11-LABX-0005-01) and the
atmospheric reactivity with NO3 radicals in the CHARME CLIMIBIO program supported by the Hauts-de-France
atmospheric simulation chamber. Rate coefficients are Regional Council, the French Ministry of Higher Education
determined for the reaction of the studied compounds at and Research, and the European Regional Development Fund.
294 K and 760 Torr based on the relative rate method. These
values highlight the effect of the methyl group on increasing
the rate coefficient due to the inductive effect induced on the
ring. Also, this work sheds light on the importance of the
■ REFERENCES
(1) Durigan, M. R.; Cherubin, M. R.; de Camargo, P. B.; Ferreira, J.
N.; Berenguer, E.; Gardner, T. A.; Barlow, J.; Dias, C. T.; dos, S.;
reaction with NO3 as a major atmospheric sink for this class of Signor, D.; de Oliveira, R. C.; et al. Soil organic matter responses to
compounds. This work is the first study reporting the rate anthropogenic forest disturbance and land use change in the eastern
coefficient measured for the reaction of 2,3,5-trimethylfuran. Brazilian Amazon. Sustainability 2017, 9, No. 379.

F https://doi.org/10.1021/acs.jpca.2c03828
J. Phys. Chem. A XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry A pubs.acs.org/JPCA Article

(2) Brando, P.; Macedo, M.; Silvério, D.; Rattis, L.; Paolucci, L.; (17) Krause, T.; Tubbesing, C.; Benzing, K.; Schöler, H. F. Model
Alencar, A.; Coe, M.; Amorim, C. Amazon wildfires: Scenes from a reactions and natural occurrence of furans from hypersaline
foreseeable disaster. Flora 2020, 268, No. 151609. environments. Biogeosciences 2014, 11, 2871−2882.
(3) Paquet, M.; Arlt, D.; Knape, J.; Low, M.; Forslund, P.; Pärt, T. (18) Naeher, L. P.; Brauer, M.; Lipsett, M.; Zelikoff, J. T.; Simpson,
Quantifying the links between land use and population growth rate in C. D.; Koenig, J. Q.; Smith, K. R. Woodsmoke health effects: A
a declining farmland bird. Ecol. Evol. 2019, 9, 868−879. review. Inhalation Toxicol. 2007, 19, 67−106.
(4) Stehfest, E.; van Zeist, W. J.; Valin, H.; Havlik, P.; Popp, A.; (19) Qian, Y.; Zhu, L.; Wang, Y.; Lu, X. Recent progress in the
Kyle, P.; Tabeau, A.; Mason-D’Croz, D.; Hasegawa, T.; Bodirsky, B. development of biofuel 2,5-dimethylfuran. Renewable Sustainable
L.; et al. Key determinants of global land-use projections. Nat. Energy Rev. 2015, 41, 633−646.
Commun. 2019, 10, No. 2166. (20) Thewes, M.; Muether, M.; Pischinger, S.; Budde, M.; Brunn, A.;
(5) Crounse, J. D.; DeCarlo, P. F.; Blake, D. R.; Emmons, L. K.; Sehr, A.; Adomeit, P.; Klankermayer, J. Analysis of the impact of 2-
Campos, T. L.; Apel, E. C.; Clarke, A. D.; Weinheimer, A. J.; McCabe, methylfuran on mixture formation and combustion in a direct-
D. C.; Yokelson, R. J.; et al. Biomass burning and urban air pollution injection spark-ignition engine. Energy Fuels 2011, 25, 5549−5561.
over the Central Mexican Plateau. Atmos. Chem. Phys. 2009, 9, 4929− (21) Tapia, A.; Villanueva, F.; Salgado, M. S.; Cabañas, B.; Martínez,
4944. E.; Martín, P. Atmospheric degradation of 3-methylfuran: Kinetic and
(6) Kumar, V.; Chandra, B. P.; Sinha, V. Large unexplained suite of products study. Atmos. Chem. Phys. 2011, 11, 3227−3241.
(22) Aschmann, S. M.; Nishino, N.; Arey, J.; Atkinson, R. Kinetics of
chemically reactive compounds present in ambient air due to biomass
the reactions of OH radicals with 2- and 3-methylfuran, 2,3- and 2,5-
fires. Sci. Rep. 2018, 8, No. 626.
dimethylfuran, and E- and Z-3-hexene-2,5-dione, and products of OH
(7) Decker, Z. C. J.; Robinson, M.; Barsanti, K.; Bourgeois, I.;
+ 2,5-dimethylfuran. Environ. Sci. Technol. 2011, 45, 1859−1865.
Coggon, M.; DiGangi, J.; Diskin, G.; Flocke, F.; Franchin, A.;
(23) Whelan, C. A.; Eble, J.; Mir, Z. S.; Blitz, M. A.; Seakins, P. W.;
Fredrickson, C.; Hall, S.; et al. Nighttime chemical transformation in Olzmann, M.; Stone, D. Kinetics of the reactions of hydroxyl radicals
biomass burning plumes: a box model analysis initialized with aircraft with furan and its alkylated derivatives 2-methylfuran and 2,5-
observations. Environ. Sci. Technol. 2019, 53, 2529−2538. dimethylfuran. J. Phys. Chem. A 2020, 124, 7416−7426.
(8) Lalchandani, V.; Srivastava, D.; Dave, J.; Mishra, S.; Tripathi, N.; (24) Cabañas, B.; Tapia, A.; Villaneuva, F.; Salgado, S. Kinetic study
Shukla, A. K.; Sahu, R.; Thamban, N. M.; Gaddamidi, S.; Dixit, K.; of 2-furanaldehyde, reactions initiated 5-methyl-2-furanaldehyde 3-
et al. Effect of biomass burning on PM 2.5 composition and secondary furanaldehyde, and by Cl atoms. Int. J. Chem. Kinet. 2008, 43, 671−
aerosol formation during post-monsoon and winter haze episodes in 678.
Delhi. J. Geophys. Res.: Atmos. 2021, 127, No. e2021JD035232. (25) Berndt, T.; Böge, O.; Rolle, W. Products of the gas-phase
(9) Koss, A. R.; Sekimoto, K.; Gilman, J. B.; Selimovic, V.; Coggon, reactions of NO3 radicals with furan and tetramethylfuran. Environ.
M. M.; Zarzana, K. J.; Yuan, B.; Lerner, B. M.; Brown, S. S.; Jimenez, Sci. Technol. 1997, 31, 1157−1162.
J. L.; et al. Non-methane organic gas emissions from biomass burning: (26) Joo, T.; Rivera-Rios, J. C.; Takeuchi, M.; Alvarado, M. J.; Ng,
Identification, quantification, and emission factors from PTR-ToF N. L. Secondary organic aerosol formation from reaction of 3-
during the FIREX 2016 laboratory experiment. Atmos. Chem. Phys. methylfuran with nitrate radicals. ACS Earth Space Chem. 2019, 3,
2018, 18, 3299−3319. 922−934.
(10) Gilman, J. B.; Lerner, B. M.; Kuster, W. C.; Goldan, P. D.; (27) Liang, Y.; Weber, R. J.; Misztal, P. K.; Jen, C. N.; Goldstein, A.
Warneke, C.; Veres, P. R.; Roberts, J. M.; De Gouw, J. A.; Burling, I. H. Aging of volatile organic compounds in October 2017 northern
R.; Yokelson, R. J. Biomass burning emissions and potential air quality california wildfire plumes. Environ. Sci. Technol. 2022, 56, 1557−1567.
impacts of volatile organic compounds and other trace gases from (28) Atkinson, R.; Aschmann, S. M.; Winer, A. M.; Carter, W. P. L.
fuels common in the US. Atmos. Chem. Phys. 2015, 15, 13915−13938. Rate Constants for the gas-phase reactions of NO3 radicals with furan,
(11) Hatch, L. E.; Luo, W.; Pankow, J. F.; Yokelson, R. J.; Stockwell, thiophene, and pyrrole at 295 ± 1 K and atmospheric pressure.
C. E.; Barsanti, K. C. Identification and quantification of gaseous Environ. Sci. Technol. 1985, 19, 87−90.
organic compounds emitted from biomass burning using two- (29) Kind, I.; Berndt, T.; Bisge, O.; Rolle, W. Gas-phase rate
dimensional gas chromatography-time-of-flight mass spectrometry. constants for the reaction of NO3 radicals with furan and methyl-
Atmos. Chem. Phys. 2015, 15, 1865−1899. substituted furans. Chem. Phys. Lett. 1996, 256, 679−683.
(12) Andreae, M. Emission of trace gases and aerosols from biomass (30) Newland, M. J.; Ren, Y.; Mcgillen, M. R.; Michelat, L.; Daële,
burning. Atmos. Chem. Phys. 2019, 15, 955−966. V.; Mellouki, A. NO3 chemistry of wildfire emissions: A kinetic study
(13) Stockwell, C. E.; Veres, P. R.; Williams, J.; Yokelson, R. J. of the gas-phase reactions of furans with the NO3 radical. Atmos.
Characterization of biomass burning emissions from cooking fires, Chem. Phys. 2022, 22, 1761−1772.
peat, crop residue, and other fuels with high-resolution proton- (31) Fayad, L.; Coeur, C.; Fagniez, T.; Secordel, X.; Houzel, N.;
Mouret, G. Kinetic and mechanistic study of the gas-phase reaction of
transfer-reaction time-of-flight mass spectrometry. Atmos. Chem. Phys.
ozone with γ-terpinene. Atmos. Environ. 2021, 246, No. 118073.
2015, 15, 845−865.
(32) Atkinson, R.; Aschmann, S. M.; Winer, A. M.; Pitts, J. N.
(14) Decker, Z. C. J.; Robinson, M.; Barsanti, K.; Bourgeois, I.;
Kinetics and atmospheric implications of the gas-phase reactions of
Coggon, M.; DiGangi, J.; Diskin, G.; Flocke, F.; Franchin, A.;
NO3 radicals with a series of monoterpenes and related organics at
Fredrickson, C.; et al. Nighttime and daytime dark oxidation 294 ± 2 K. Environ. Sci. 1985, 19, 159−163.
chemistry in wildfire plumes: an observation and model analysis of (33) IUPAC. IUPAC task group on atmospheric chemical kinetic
FIREX-AQ aircraft data. Atmos. Chem. Phys. 2021, 21, 16293−16317. data evaluation − data sheet NO3_VOC, (http://iupac.pole-ether.fr).
(15) Coggon, M. M.; Lim, C. Y.; Koss, A. R.; Sekimoto, K.; Yuan, B.; (34) Atkinson, R.; Aschmann, S. M.; Pitts, J. N. Rate constants for
Gilman, J. B.; Hagan, D. H.; Selimovic, V.; Zarzana, K. J.; Brown, S. the gas-phase reactions of the NO3 radical with a series of organic
S.; et al. OH chemistry of non-methane organic gases (NMOGs) compounds at 296 ± 2 K. J. Phys. Chem. A 1988, 92, 3454−3457.
emitted from laboratory and ambient biomass burning smoke: (35) Lançar, I.; Daele, V.; Le Bras, G.; Poulet, G. É tude de la
Evaluating the influence of furans and oxygenated aromatics on réactivité des radicaux NO3 avec le diméthyl-2,3 butène-2, le
ozone and secondary NMOG formation. Atmos. Chem. Phys. 2019, 19, butadiène-1,3 et le diméthyl-2,3 butadiène-1,3. J. Chim. Phys. 1991,
14875−14899. 88, 1777−1792.
(16) Yokelson, R. J.; Karl, T.; Artaxo, P.; Blake, D. R.; Christian, T. (36) Grira, A.; Amarandei, C.; Romanias, M. N.; El Dib, G.; Canosa,
J.; Griffith, D. W. T.; Guenther, A.; Hao, W. M. The tropical forest A.; Arsene, C.; Bejan, I. G.; Olariu, R. I.; Coddeville, P.; Tomas, A.
and fire emissions experiment: Overview and airborne fire emission Kinetic measurements of Cl atom reactions with C5-C8 unsaturated
factor measurements. Atmos. Chem. Phys. 2007, 7, 5175−5196. alcohols. Atmosphere 2020, 11, No. 256.

G https://doi.org/10.1021/acs.jpca.2c03828
J. Phys. Chem. A XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry A pubs.acs.org/JPCA Article

(37) Cabañas, B.; Baeza, M. T.; Salgado, S.; Martín, P.; Taccone, R.;
Martínez, E. Oxidation of heterocycles in the atmosphere: Kinetic
study of their reactions with NO3 radical. J. Phys. Chem. A 2004, 108,
10818−10823.
(38) Alvarado, A.; Atkinson, R.; Arey, J. Kinetics of the gas-phase
reactions of NO3 radicals and O3 with 3-methylfuran and the OH
radical yield from the O3 reaction. Int. J. Chem. Kinet. 1996, 28, 905−
909.
(39) Brown, S. S.; Stutz, J. Nighttime radical observations and
chemistry. Chem. Soc. Rev. 2012, 41, 6405−6447.
(40) Hein, R.; Crutzen, P. J.; Heimann, M. An inverse modeling
approach to investigate the global atmospheric methane cycle. Global
Biogeochem. Cycles 1997, 11, 43−76.
(41) Matsumoto, J. Kinetics of the reactions of ozone with 2,5-
dimethylfuran and its atmospheric implications. Chem. Lett. 2011, 40,
582−583.
(42) Olariu, R. I.; Bejan, I.; Barnes, I.; Klotz, I.; Becker, K. H.; Wirtz,
K. Rate coefficients for the gas-phase reaction of NO3 radicals with
selected dihydroxybenzenes. Int. J. Chem. Kinet. 2004, 36, 577−583.
(43) Cabañas, B.; Villanueva, F.; Martín, P.; Baeza, M. T.; Salgado,
S.; Jiménez, E. Study of reaction processes of furan and some furan
derivatives initiated by Cl atoms. Atmos. Environ. 2004, 39, 1935−
1944.
(44) Bierbach, A.; Barnes, I.; Becker, K. H. Rate coefficients for the
gas-phase reactions of hydroxyl radicals with furan, 2-methylfuran, 2-
ethylfuran and 2,5-dimethylfuran at 300 ± 2 K. Atmos. Environ., Part A
1992, 26, 813−817.
(45) Andersen, C.; Nielsen, O. J.; Østerstrøm, F. F.; Ausmeel, S.;
Nilsson, E. J. K.; Sulbaek Andersen, M. P. Atmospheric chemistry of
tetrahydrofuran, 2-methyltetrahydrofuran, and 2,5-dimethyltetrahy-
drofuran: kinetics of reactions with chlorine atoms, OD radicals, and
ozone. J. Phys. Chem. A 2016, 120, 7320−7326.
(46) Wingenter, O. W.; Kubo, M. K.; Blake, N. J.; Smith, T. W.;
Blake, D. R.; Rowland, F. S. Hydrocarbon and halocarbon
measurements as photochemical and dynamical indicators of
atmospheric hydroxyl, atomic chlorine, and vertical mixing obtained
during Lagrangian flights. J. Geophys. Res.: Atmos. 1996, 101, 4331− Recommended by ACS
4340.
Rate Coefficients for OH + NO (+N2) in the Fall-off Regime
and the Impact of Water Vapor
Wenyu Sun, John N. Crowley, et al.
JUNE 08, 2022
THE JOURNAL OF PHYSICAL CHEMISTRY A READ

Atmospheric Oxidation of Propanesulfinic Acid Initiated by


OH Radicals: Reaction Mechanism, Energetics, Rate
Coefficients, and Atmospheric Implications
Parandaman Arathala and Rabi A. Musah
MAY 17, 2021
ACS EARTH AND SPACE CHEMISTRY READ

Gas Phase Reaction of Isocyanic Acid: Kinetics, Mechanisms,


and Formation of Isopropyl Aminocarbonyl
Tien Van Pham and Anh Van Tran
DECEMBER 12, 2021
ACS OMEGA READ

Theoretical Investigations of the OH-Initialized Oxidation of


4-Methyl-3-Penten-2-One in the Atmosphere
Benni Du and Weichao Zhang
AUGUST 26, 2022
ACS EARTH AND SPACE CHEMISTRY READ

Get More Suggestions >

H https://doi.org/10.1021/acs.jpca.2c03828
J. Phys. Chem. A XXXX, XXX, XXX−XXX

You might also like