Products of the gas-phase reaction of the OH radical with 3-methyl-1-butene in the presence of NO

11
Products of the Gas-Phase Reaction of the OH Radical with 3-Methyl-1-Butene in the Presence of NO ROGER ATKINSON,* ERNESTO C. TUAZON, SARA M. ASCHMANN Air Pollution Research Center, University of California, Riverside, California 92521 Received 29 September 1997; accepted 26 February 1998 ABSTRACT: The products of the gas-phase reaction of the OH radical with 3-methyl-1-butene in the presence of NO have been investigated at room temperature and total pressure 740 torr of air by gas chromatography with flame ionization detection, in situ Fourier transform infrared absorption spectroscopy, and direct air sampling atmospheric pressure ionization tandem mass spectrometry. The products identified and quantified by GC-FID and in situ FT-IR ab- sorption spectroscopy were HCHO, 2-methylpropanal, acetone, glycolaldehyde, and methac- rolein, with formation yields of and 0.70 6 0.06, 0.58 6 0.08, 0.17 6 0.02, 0.18 6 0.03, respectively. In addition, IR absorption bands due to organic nitrates were 0.033 6 0.007, observed, consistent with API-MS observations of product ion peaks attributed to the b-hy- droxynitrates (CH 3 ) 2 CHCH(ONO 2 )CH 2 OH and/or (CH 3 ) 2 CHCH(OH)CH 2 ONO 2 formed from the reactions of the corresponding b-hydroxyalkyl peroxy radicals with NO. A formation yield of ca. 0.15 for these nitrates was estimated using IR absorption band intensities for known or- ganic nitrates. These products account for essentially all of the reacted 3-methyl-1-butene. Analysis of the potential reaction pathways involved shows that H-atom abstraction from the allylic C 9H bond in 3-methyl-1-butene is a minor pathway which accounts for 5–10% of the overall OH radical reaction. q 1998 John Wiley & Sons, Inc. Int J Chem Kinet: 30: 577– 587, 1998 INTRODUCTION Alkenes are a significant component of ambient air in urban areas [1 – 3] and react with OH radicals, NO 3 radicals, and O 3 in the troposphere [4,5], with the day- * Also, Environmental Toxicology Graduate Program and the Departments of Environmental Science and Chemistry, University of California, Riverside, California 92521 Correspondence to: R. Atkinson Contract grant Sponsor: U.S. Environmental Protection Agency, Office of Research and Development Contract grant number: Assistance Agreement R-825252-01-0 Contract grant Sponsor: National Science Foundation Contract grant number: Grant No. ATM-9015361 Contract grant Sponsor: University of California, Riverside q 1998 John Wiley & Sons, Inc. CCC 0538-8066/98/080577-11 time OH radical reaction often dominating as the al- kene removal process [6,7]. While the kinetics of the gas-phase reactions of the OH radical with alkenes have been studied previously [4,5,8] and are reason- ably reliably known [5], few product studies have been conducted at atmospheric pressure of air for simple acyclic alkenes [5,9 – 18]. Kinetic and product data show that at room temperature and atmospheric pres- sure the OH radical reactions proceed mainly by ad- dition to the bond [4,5,8]. Extrapolations . C " C , of rate constants measured at elevated temperatures [4,8] indicate that the fractions of the OH (. 600 K) radical reactions with ethene, propene, and 1-butene which proceed by H-atom abstraction at are ca. 298 K

Transcript of Products of the gas-phase reaction of the OH radical with 3-methyl-1-butene in the presence of NO

JCK(Wiley) RIGHT INTERACTIVE

shortstandardlong

Products of the Gas-PhaseReaction of the OH Radicalwith 3-Methyl-1-Butene inthe Presence of NOROGER ATKINSON,* ERNESTO C. TUAZON, SARA M. ASCHMANN

Air Pollution Research Center, University of California, Riverside, California 92521

Received 29 September 1997; accepted 26 February 1998

ABSTRACT: The products of the gas-phase reaction of the OH radical with 3-methyl-1-butenein the presence of NO have been investigated at room temperature and total pressure740 torrof air by gas chromatography with flame ionization detection, in situ Fourier transform infraredabsorption spectroscopy, and direct air sampling atmospheric pressure ionization tandemmass spectrometry. The products identified and quantified by GC-FID and in situ FT-IR ab-sorption spectroscopy were HCHO, 2-methylpropanal, acetone, glycolaldehyde, and methac-rolein, with formation yields of and0.70 6 0.06, 0.58 6 0.08, 0.17 6 0.02, 0.18 6 0.03,

respectively. In addition, IR absorption bands due to organic nitrates were0.033 6 0.007,observed, consistent with API-MS observations of product ion peaks attributed to the b-hy-droxynitrates (CH3)2CHCH(ONO2)CH2OH and/or (CH3)2CHCH(OH)CH2ONO2 formed from thereactions of the corresponding b-hydroxyalkyl peroxy radicals with NO. A formation yield ofca. 0.15 for these nitrates was estimated using IR absorption band intensities for known or-ganic nitrates. These products account for essentially all of the reacted 3-methyl-1-butene.Analysis of the potential reaction pathways involved shows that H-atom abstraction from theallylic C9H bond in 3-methyl-1-butene is a minor pathway which accounts for 5–10% of theoverall OH radical reaction. q 1998 John Wiley & Sons, Inc. Int J Chem Kinet: 30: 577–587, 1998

INTRODUCTION

Alkenes are a significant component of ambient air inurban areas [1–3] and react with OH radicals, NO3

radicals, and O3 in the troposphere [4,5], with the day-

* Also, Environmental Toxicology Graduate Program and theDepartments of Environmental Science and Chemistry, Universityof California, Riverside, California 92521

Correspondence to: R. AtkinsonContract grant Sponsor: U.S. Environmental Protection Agency,

Office of Research and DevelopmentContract grant number: Assistance Agreement R-825252-01-0Contract grant Sponsor: National Science FoundationContract grant number: Grant No. ATM-9015361Contract grant Sponsor: University of California, Riverside

q 1998 John Wiley & Sons, Inc. CCC 0538-8066/98/080577-11

time OH radical reaction often dominating as the al-kene removal process [6,7]. While the kinetics of thegas-phase reactions of the OH radical with alkeneshave been studied previously [4,5,8] and are reason-ably reliably known [5], few product studies have beenconducted at atmospheric pressure of air for simpleacyclic alkenes [5,9–18]. Kinetic and product datashow that at room temperature and atmospheric pres-sure the OH radical reactions proceed mainly by ad-dition to the bond [4,5,8]. Extrapolations. C"C ,of rate constants measured at elevated temperatures

[4,8] indicate that the fractions of the OH(.600 K)radical reactions with ethene, propene, and 1-butenewhich proceed by H-atom abstraction at are ca.298 K

578 ATKINSON, TUAZON, AND ASCHMANN

JCK(Wiley) LEFT INTERACTIVE

shortstandardlong

0.1%, ca. 3%, and ca. 7%, respectively [4,8]. Rate con-stants measured at for 2-methylpro-1237–1275 Kpene, trans-2-butene, and 2,3-dimethyl-2-butene [19]are also consistent with H-atom abstraction accountingfor ca. 2–3% of the overall OH radical reaction at

[8]. Product studies of the reactions of the OH298 Kradical with propene, 1-butene, 2-methylpropene, cis-2-butene, and trans-2-butene [11,20,21] have con-cluded that H-atom abstraction is minor at room tem-perature and atmospheric pressure (,5%, ,10%,

respectively).,5%, ,10%, and ,10%,The bond dissociation energies of the allylic

C9H bonds in propene, 1-butene, and 3-methyl-1-butene are 86.3 6 1.5, 82.5 6 1.3, and 77.2 6 1.5

respectively [22,23]. Extrapolation of the21kcal mol ,estimated H-atom abstraction rate constants for pro-pene and 1-butene [4,8] to 3-methyl-1-butene, assum-ing a linear correlation of ln (H-atom abstraction rateconstant at ) against the allylic C9H bond dis-298 Ksociation energy, suggests that H-atom abstractionfrom the allylic C9H bond in 3-methyl-1-butenecould account for ca. 30% of the overall reaction at

and atmospheric pressure. We have used gas298 Kchromatography with flame ionization detection (GC-FID), in situ atmospheric pressure ionization tandemmass spectrometry (API-MS), and in situ Fouriertransform infrared absorption spectroscopy (FT-IR) toinvestigate the products formed from the gas-phasereaction of the OH radical with 3-methyl-1-buteneat room temperature and atmospheric pressure ofair.

EXPERIMENTAL

Experiments were carried out in a evacua-5800 literble, Teflon-coated chamber containing an in situ mul-tiple reflection optical system interfaced to a Nicolet7199 FT-IR absorption spectrometer and with irradi-ation provided by a 24-kW xenon arc filtered througha thick Pyrex pane (to remove wavelengths0.25 in.

), in a Teflon chamber with anal-,300 nm 7900 literysis by GC-FID and with irradiation provided by twoparallel banks of blacklamps, and in a second 7900liter Teflon chamber interfaced to a PE SCIEX API IIIMS/MS direct air sampling, atmospheric pressure ion-ization tandem mass spectrometer (API-MS), againwith irradiation provided by two parallel banks ofblacklamps. All three chambers are fitted with Teflon-coated fans to ensure rapid mixing of reactants duringtheir introduction into the chamber.

Hydroxyl radicals were generated by the photolysisof methyl nitrite (CH3ONO) or ethyl nitrite

(C2H5ONO) in air at wavelengths .300 nm[11,12,24]:

R BCH ONO 1 hn !: RCH BO 1 NO (1)2 2

RCH BO 1 O !: RCHO 1 HO (2)2 2 2

HO 1 NO !: OH 1 NO (3)2 2

where or CH3. NO was added to the reactantR 5 Hmixtures to suppress the formation of O3 and, hence,of NO3 radicals [24]. Because HCHO is the primaryphotolytic product of methyl nitrite (see above) andHCHO is expected to be a major product of the OHradical-initiated reaction of 3-methyl-1-butene in thepresence of NO [5], the photolysis of ethyl nitrite inair was used as the OH radical source for the deter-mination of HCHO formation yields in the 5800 literevacuable chamber. In the 7900 liter Teflon chamberswith analyses by GC-FID and API-MS, OH radicalswere generated by the photolysis of CH3ONO in air.

Teflon Chamber with Analysis by GC-FID

The initial reactant concentrations (in molecule 23cmunits) were: CH3ONO, NO, ca. 2.4 314ca. 2.4 3 10 ;10 and 3-methyl-1-butene, (2.03–2.29) 3 10 Ir-14 13; .radiations were carried out at 20% of the maximumlight intensity for resulting in reaction of4–12 mins,up to 62% of the initially present 3-methyl-1-butene.The concentrations of 3-methyl-1-butene were mea-sured during the experiments by GC-FID as describedpreviously [25]. The concentrations of certain car-bonyl products were measured by GC-FID by collect-ing volume gas samples from the chamber3100 cmonto Tenax-TA solid adsorbent, with subsequent ther-mal desorption at ca. 2257C onto a DB-170130 mmegabore column in a Hewlett Packard (HP) 5710GC, initially held at 2407C and then temperature pro-grammed at to 2007C. GC-FID response2187C minfactors were measured as described previously[25,26]. In addition, combined gas chromatography-mass spectrometry (GC-MS) analyses of samples col-lected on Tenax solid adsorbent were carried out usinga DB-1701 fused silica capillary column in a HP30 m5890 GC interfaced to a HP 5971 Mass Selective De-tector, operating in the scanning mode.

Evacuable Chamber with Analysis by FT-IRAbsorption Spectroscopy

The initial concentrations of 3-methyl-1-butene, ethylnitrite, and NO in the two experiments carried outwere 4.92 3 10 molecule cm 2.46 3 10 mole-14 23 14,

GAS-PHASE REACTION OF OH RADICAL 579

JCK(Wiley) RIGHT INTERACTIVE

shortstandardlong

(CH3)2CHCH(O)CH2OH (CH3)2CHC(O)CH2OH 1 HO2

Reaction Scheme I

O2 O2

O2

4.1 3 104 s21

2.6 3 105 s21decomposition

5.0 3 105 s21decomposition

?

(CH3)2CH 1

NO NO2

HC(O)CH2OH? ?

O2

CH3CH(O)CH3

CH3C(O)CH3 1 HO2

?

(CH3)2CHCHO 1 CH2OH

HCHO 1 HO2

cule cm and molecule23 14 23, (1.85–3.08) 3 10 cm ,respectively. In the first experiment which employedan initial NO concentration of molecule141.85 3 10

an additional molecule of NO23 14 23cm , 1.85 3 10 cmwas added halfway through the experiment. Partialpressures of 3-methyl-1-butene, ethyl nitrite, and NOwere measured into calibrated 5 liter and 2 liter Pyrexbulbs with an MKS Baratron sensor and0–100 torrflushed into the 5800 liter chamber with N2 gas. Fourintermittent irradiations of duration were1.0–3.0 mincarried out during each experiment, resulting in reac-tion of up to 25% of the initially present 3-methyl-1-butene. Analyses by FT-IR spectroscopy were carriedout during the dark periods, with 64 coadded interfer-ograms (scans) per spectrum ( measurement1.8 mintime) being recorded with a full width at half-maxi-mum (fwhm) resolution of and a pathlength210.7 cmof The product spectra were analyzed by de-62.9 m.synthesis, which consisted of successively subtractingthe absorptions of known compounds with the use ofcalibrated reference spectra. The methods employedin the calibration and subtractive analysis have beendescribed in detail previously [27]. IR reference spec-tra of the products measured were available from pre-vious calibrations in this laboratory [27].

Teflon Chamber with Analysis by API-MS

The PE SCIEX API III MS/MS instrument was inter-faced to the Teflon chamber via a 25 mm diameter 375 cm length Pyrex tube, and the chamber contentswere sampled at a flow rate of ca. 20 liter di-21minrectly into the API mass spectrometer source. The op-eration of the API-MS in the MS (scanning) and MS/MS [with collision activated dissociation (CAD)]modes has been described elsewhere [17,18,28,29].

The MS/MS mode allows the “daughter ion” or “par-ent ion” spectrum of a given ion peak observed in theMS scanning mode to be obtained [17,18,28,29]. Thepositive ion mode was used in this work, with proton-ated water hydrates generated by the1(H O (H O) )3 2 n

corona discharge in the chamber diluent gas being re-sponsible for the protonation of analytes. Ions aredrawn by an electric potential from the ion sourcethrough the sampling orifice into the mass-analyzingfirst quadrupole or third quadrupole. For these exper-iments, the API-MS instrument was operated underconditions that favored the formation of dimer ions inthe ion source region [28–30]. Neutral molecules andparticles are prevented from entering the orifice by aflow of high-purity nitrogen (“curtain” gas), and as aresult of the declustering action of the curtain gas onthe hydrated ions, the ions that are mass-analyzed aremainly protonated molecular ions and1([M 1 H] )their protonated homo- and hetero-dimers [17,18,28–30]. The initial concentrations of CH3ONO, NO,and 3-methyl-1-butene were ca. 4.8 3 10 molecule13

each, and irradiations were carried out for23cmat 20% of the maximum light inten-3 min

sity.

Chemicals

The chemicals used, and their stated purities, were:acetone, .99.6% (Fisher Scientific); methacrolein(95%), 3-methyl-2-butenal (97%), 2-methylpropanal(99 1 %), and methyl vinyl ketone (99%), AldrichChemical Company; NO ($ 99.0%), Matheson GasProducts; and 3-methyl-1-butene (99 1 %), LiquidCarbonic. Methyl nitrite was prepared as described byTaylor et al. [31] and was stored at under vac-77 Kuum.

580 ATKINSON, TUAZON, AND ASCHMANN

JCK(Wiley) LEFT INTERACTIVE

shortstandardlong

RESULTS

Reaction Schemes I and II show the reaction mecha-nisms and the products which can be formed fromthe alkoxy radicals and(CH ) CHCH( BO)CH OH3 2 2

OH 1 (CH ) CHCH"CH !: (CH ) CH BCHCH OH (4)3 2 2 3 2 2

(CH ) CH BCHCH OH 1 O !:(CH ) CHCH(O BO)CH OH (5)3 2 2 2 3 2 2

(6a)

(6b)(CH3)2CHCH(OO)CH2OH 1 NO

(CH3)2CHCH(ONO2)CH2OH

(CH3)2CHCH(O)CH2OH 1 NO2

M

??

produced after initial OH(CH ) CHCH(OH)CH BO,3 2 2

radical addition to the 1- and 2-positions through thereactions illustrated for OH radical addition at the1-position [5],

and similarly for OH radical addition at the 2-position.Reaction Scheme III shows the reaction mechanismsand the products which can be formed from the

radical formed after H-atom ab-(CH ) BCCH"CH3 2 2

straction from the allylic C9H bond of the carbonatom at the 3-position. For clarity, organic nitrate for-mation from the reactions of radicals with NO [5]R BO2

via Reaction (6a) and analogous reactions is omittedfrom Reaction Schemes I– III. Also, shown in Reac-tion Schemes I– III are the estimated reaction rates ofthe important alkoxy radicals involved, calculated asdescribed by Atkinson [32]. H-atom abstraction fromthe two 9CH3 groups in 3-methyl-1-butene is esti-mated to account for only ca. 1% of the overall OHradical reaction [33], and the subsequent reactions ofthe radical are expectedBCH CH(CH )CH"CH2 3 2

[5,32] to lead to the formation of methyl vinyl ketoneplus formaldehyde. The potential products shown inReaction Schemes I through III form the basis for

the discussion (observed products are shown by“boxes”).

Teflon Chamber with GC-FID Analyses andEvacuable Chamber with FT-IR Analyses

GC-FID and GC-MS analyses of irradiated CH3-ONO-NO-3-methyl-1-butene-air mixtures showed theformation of acetone, 2-methylpropanal, and methac-rolein [CH2 "C(CH3)CHO]. However, no evidencewas obtained from the GC-FID and GC-MS analysesfor the formation of methyl vinyl ketone[CH3C(O)CH"CH2] or 3-methyl-2-butenal[(CH3)2C"CHCHO], and upper limits to the amounts ofthese two products were obtained from conservativelimits of detection of the GC-FID analyses. FT-IRanalyses of irradiated C2H5ONO-NO-3-methyl-1-butene-air mixtures showed the formation of 2-methylpropanal, HCHO, acetone, glycolaldehyde

(CH3)2CHCH(OH)CH2O

CH2CH(CH3)CH(OH)CH2OH

(CH3)2CHCH(OH)CHO 1 HO2

Reaction Scheme II

O2 O2

O2

3.6 3 104 s21

4.0 3 105 s21isomerization

2.5 3 106 s21decomposition

?

NO NO2

1 HO2

?

OOCH2CH(CH3)CH(OH)CH2OH?

O2

HOCH2CH(CH3)CH(OH)CHO

OCH2CH(CH3)CH(OH)CH2OH?

HOCH2CH(CH3)CH(OH)CHOH?

?

isomerization

(CH3)2CHCHO

1 (CH3)2CHCHOHHCHO

1 HO2

GA

S-P

HA

SE

RE

AC

TION

OF

OH

RA

DIC

AL

581

JCK

(Wiley)

RIG

HT

INT

ER

AC

TIV

E

shortstandardlong

CH"CH2 1 CH3C(O)CH3

O2

?

?

?

CH3 1 CH3C(O)CH"CH2? CH2C(CH3)"CHCH2OH? CH2"C(CH3)CHCH2OH?

?

HCO 1 HCHO

HCHO

O2

O2

isomerization6.9 3 106 s21

NO NO2

OCH2C(CH3)"CHCH2OH

O2

NO NO2

CH2"C(CH3)CH(O)CH2OH

1 CH2OH

?

?

?

3.1 3 106 s21decomposition

2 3 105 s21O28.5 3 104 s21

O2

CH2"C(CH3)CHO CH2"C(CH3)C(O)CH2OH

1 HO2

O2

HCHOHO2 1

HOCH2C(CH3)"CHCHOHHO2 1 HC(O)C(CH3)"CHCH2OH

(CH3)2CCH"CH2

O2

O2

decomposition8.6 3 102 s21

NO NO2

(CH3)2C(O)CH"CH2

?

?

(CH3)2C"CHCH2

O2

isomerization3.2 3 105 s21

1.0 3 105 s21

NO NO2

(CH3)2C"CHCH2O

?

5.7 3 101 s21decomposition

HOCH2C(CH3)"CHCHO 1 HO2

(CH3)2C"CHCHO 1 HO2

Reaction Scheme III

582 ATKINSON, TUAZON, AND ASCHMANN

JCK(Wiley) LEFT INTERACTIVE

shortstandardlong

Figure 1 Infrared spectra from a C2H5ONO-NO-3-methyl-1-butene-air irradiation. (A) Products attributed to reactionof 3-methyl-1-butene (see text) after of photolysis,9 minwith molecule of 3-methyl-1-butene con-14 231.24 3 10 cmsumed. (B) From (A) after subtraction of HCHO absorp-tions. (C) From (B) after subtraction of absorptions due toacetone and 2-methylpropanal. Asterisks denote RONO2-type absorption bands, while arrows indicate RC(O)OONO2-type absorption bands. The numbers in parentheses are con-centrations in units of molecule13 2310 cm .

[HOCH2CHO], methacrolein, and unidentified organicnitrates.

The results of FT-IR analyses for the two irradia-tions conducted in the 5800 liter chamber were similar,except for the formation of small amounts ofHOONO2 in the first experiment after NO was con-sumed, which was suppressed as a second aliquot ofNO was added. NO remained in excess throughout thesecond experiment, which employed a higher initialNO concentration. Product spectra from the latter ex-periment, after a cumulative irradiation time of

are shown in Figure 1. Figure 1(A) is the spec-9 min,trum of the products which are attributable to3-methyl-1-butene, obtained after subtraction of ab-sorptions by the remaining reactants and products as-sociated with irradiated C2H5ONO-NO-air mixtures[CH3CHO, C2H5ONO2, CH3C(O)OONO2 (PAN),NO2, HNO3, HONO, and HC(O)OH]. The subtractivesteps depicted in Figure 1(A)–(C) reveal the compo-nents of the overlapping absorptions in the

region. In Figure 1(C), prominent absorp-211740 cmtion bands at 852, 1284, and indicate the211658 cmpresence of an organic nitrate(s), RONO2. A distinctabsorption band at , most likely due to a211035 cmC9O stretch, is consistent with the API-MS results

(see below) that the RONO2 compounds formed con-tain hydroxyl groups (the weak O9H stretch absorp-tion band near was beyond the optimal213700 cmrange of the IR detector). The formation of a peroxy-acyl nitrate, most likely (CH3)2CHC(O)OONO2,which became more important during the latter part ofthe irradiation, is indicated by the characteristic “PAN-type” absorption bands at 794, 1300, 1738, and

211830 cm .The observed products also react with the OH rad-

ical [4,8] and, hence, their secondary reactions mustbe considered in deriving their formation yields (orupper limits thereof). Secondary reactions of the ob-served or (for methyl vinyl ketone and 3-methyl-2-butenal) potential products with the OH radical weretaken into account as described previously [34], usingthe recommended OH radical reaction rate constantsfor 3-methyl-1-butene, acetone, 2-methylpropanal,HCHO, glycolaldehyde, methacrolein, and methyl vi-nyl ketone [4,5,8] and the estimated rate constant for3-methyl-2-butenal [33]. The multiplicative correctionfactors to the measured product concentrations to takeinto account these secondary reactions depend onthe rate constant ratio k(OH) 1 product)/k(OH 1

and increase with increasing ex-3-methyl-1-butene)tent of reaction and were therefore larger for the ex-periments in the Teflon chamber with GC-FID anal-yses than in the evacuable chamber with FT-IRanalyses. The maximum calculated values of the mul-tiplicative factors were: for 2-methylpropanal, 1.55(experiments with GC-FID analyses) and 1.13 (exper-iments with FT-IR analyses); acetone, ,1.004 in allcases; methacrolein, 1.73 (experiments with GC-FIDanalyses) and 1.17 (experiments with FT-IR analyses);methyl vinyl ketone, 1.35; 3-methyl-2-butenal, 2.02;HCHO, 1.05; and glycolaldehyde, 1.05. The reactionof the OH radical with 2-methylpropanal may be ex-pected to lead to the formation of acetone and/or theperoxyacyl nitrate (CH3)2CHC(O)OONO2, dependingon the NO/NO2 concentration ratio [4]. Additionally,the formation of small amounts of HCHO from thereaction of the OH radical with acetaldehyde (the ma-jor product from the photolysis of ethyl nitrite) is alsoexpected [16] in the experiments carried out in theevacuable chamber with FT-IR analyses. These pos-sibilities of secondary formation of acetone andHCHO are discussed below.

Figure 2 shows plots of the amounts of HCHO,2-methylpropanal, acetone, glycolaldehyde, and meth-acrolein formed, corrected for reaction with the OHradical, against the amounts of 3-methyl-1-butene re-acted in the evacuable chamber experiments with FT-IR analyses, while Figure 3 shows analogous plots forthe formation of 2-methylpropanal, acetone, and meth-

GAS-PHASE REACTION OF OH RADICAL 583

JCK(Wiley) RIGHT INTERACTIVE

shortstandardlong

Figure 2 Plots of the amounts of HCHO, 2-methylpro-panal, acetone, glycolaldehyde, and methacrolein formed,corrected for secondary reactions (see text), against theamounts of 3-methyl-1-butene reacted with the OH radicalin the presence of NO for experiments carried out in an eva-cuable chamber with analyses by FT-IR absorption spec-troscopy. The data for acetone, 2-methylpropanal, andHCHO have been displaced vertically by 1.0 3

and molecule respec-13 13 13 2310 , 1.0 3 10 , 2.0 3 10 cm ,tively, for clarity.

Figure 3 Plots of the amounts of 2-methylpropanal, ace-tone, and methacrolein formed, corrected for secondary re-actions (see text), against the amounts of 3-methyl-1-butenereacted with the OH radical in the presence of NO for ex-periments carried out in a Teflon chamber with analyses byGC-FID. The curve shown through the acetone data pointsis for illustrative purposes only.

Table I Products and their Formation YieldsObserved from the Reaction of the OH Radical with 3-Methyl-1-Butene in the Presence of NO at 298 6 2 Kand 740 Torr Total Pressure of Air

Product

Formation Yield

GC-FID analysesa FT-IR analysesb

2-methylpropanal 0.54 6 0.05 0.63 6 0.05formaldehyde 0.70 6 0.06acetone 0.17 6 0.02glycolaldehyde 0.18 6 0.03methacrolein 0.034 6 0.004 0.031 6 0.011methyl vinyl ke-

tone,0.01

3-methyl-2-bu-tenal

,0.01

organic nitratesc ca. 0.15

Indicated errors are two least-squares standard deviations com-a

bined with estimated overall uncertainties in the GC-FID responsefactors for 3-methyl-1-butene and the products of 65% each.

Indicated errors are two least-squares standard deviations com-b

bined with estimated overall uncertainties of 65% each in the cal-ibration and subtractive analysis of 3-methyl-1-butene, 2-methyl-propanal, formaldehyde, and acetone, with correspondinguncertainties of 610% and 625% for glycolaldehyde and meth-acrolein, respectively.

Based on the API-MS analyses (see text), the organic nitratesc

are probably the b-hydroxynitrates (CH3)2CHCH(ONO2)CH2OHand/or (CH3)2CHCH(OH)CH2ONO2 formed from the reactions ofthe corresponding b-hydroxyalkyl peroxy radicals with NO.

acrolein in the Teflon chamber experiments with GC-FID analyses. All of the plots shown in Figure 2 andthose for 2-methylpropanal and methacrolein in Figure3 are good straight lines, and the formation yields ob-tained from least-squares analyses of the data aregiven in Table I. Also included in Table I are the upperlimits to the formation yields of methyl vinyl ketoneand 3-methyl-2-butenal derived from the GC-FIDanalyses.

The formation yield of HCHO given in Table I isrigorously an upper limit because of formation ofHCHO from acetaldehyde produced by the photolysisof ethyl nitrite. However, as discussed by Atkinson etal. [16], formation of HCHO from secondary reactionsof acetaldehyde in irradiated C2H5ONO9NO-3-methyl-1-butene-air mixtures is expected to lead to anHCHO formation yield of # 0.04, and, hence, theHCHO formation yield in Table I is almost totally thatformed from the OH radical reaction with 3-methyl-1-butene. Although acetone is expected to be formedfrom the secondary reaction of 2-methylpropanal,there is no evidence for any departure from linearityin the plot shown in Figure 2. Furthermore, the acetoneformation yield in the experiments with FT-IR anal-yses is identical, within the experimental uncertainties,to the yield of glycolaldehyde, the expected coproduct

584 ATKINSON, TUAZON, AND ASCHMANN

JCK(Wiley) LEFT INTERACTIVE

shortstandardlong

O2

H2O 1 (CH3)2CHC(O)OO

(CH3)2CHC(O)OONO2(CH3)2CH 1 CO2 1 NO2

NO2NO

OH 1 (CH3)2CHCHO

NO NO2

?

?

O2

O2

CH3C(O)CH3

Reaction Scheme IV

from decomposition of the (CH ) CHCH( BO)CH OH3 2 2

radical (Reaction Scheme I).The absence of an increase in the acetone yield with

reaction time in the evacuable chamber experimentswith FT-IR analyses can be explained if the peroxy-acyl nitrate absorption bands depicted in Figure 1 areindeed those of (CH3)2CHC(O)OONO2, with the lossof 2-methylpropanal by reaction with the OH radicalproceeding mainly by abstraction of the aldehydic hy-drogen atom followed by steps leading to the forma-tion of (CH3)2CHC(O)OONO2 in high yield (ReactionScheme IV). From the amount of 2-methylpropanalpresent in the spectrum of Figure 1(B) and the corre-sponding multiplicative correction factor of 1.13, aloss of molecule of 2-methylpropanal12 238.9 3 10 cmby secondary reaction with the OH radical can be es-timated. Assuming that the integrated absorption co-efficient determined for the band of211741 cmCH3C(O)OONO2 [35] applies to the analogous

band of (CH3)2CHC(O)OONO2, then the211738 cmlatter’s concentration is estimated from Figure 1(C)(after subtraction of contributions from glycolalde-hyde and methacrolein) as molecule127.1 3 10

accounting for ca. 80% (and possibly all) of the23cm ,2-methylpropanal consumed by reaction with the OHradical.

As shown in Figure 3, the acetone yield in the ex-periments with GC-FID analyses clearly increaseswith increasing amount of 3-methyl-1-butene reacted,suggesting secondary formation from 2-methylpro-panal. These experiments resulted in significantlyhigher conversions of 3-methyl-1-butene than was thecase in the experiments with FT-IR analyses (up to62% vs. up to 25% conversion, respectively), but em-ployed higher initial NO/3-methyl-1-butene concen-tration ratios such that the final NO/NO2 concentrationratios were $ 1. Reaction of 2-methylpropanal with

the OH radical would therefore be expected to lead tothe formation of acetone, at least in part [4] (ReactionScheme IV). At the lowest extents of reaction in theexperiments with GC-FID analyses, the acetone yieldapproached that observed in the experiments with FT-IR analyses. We therefore conclude that the acetoneformation yield obtained from the experiments withFT-IR analyses conducted at low conversions of 3-methyl-1-butene, and under conditions such that(CH3)2CHC(O)OONO2 was the major product formedfrom the secondary reaction of 2-methylpropanal, re-flects the formation of “first-generation” acetone, andthis yield is given in Table I.

The yield of RONO2 was estimated from theca.1280 absorption band based on an average21cmof the integrated absorption coefficients previouslymeasured for other organic nitrates [12]. Becausea peroxyacyl nitrate (assumed to be(CH3)2CHC(O)OONO2; see above) also increasinglycontributed to the intensity of this band with increas-ing irradiation time, RONO2 concentrations were de-rived from the early portions of the experiments whenthe spectra did not show measurable absorptions fromthe peroxyacyl nitrate species. Both experiments gaveRONO2 formation yields of ca. 0.15 during the earlystages of irradiation, and this value is reported in TableI.

Teflon Chamber with Analyses by API-MS

API-MS spectra of irradiated CH3ONO-NO-3-methyl-1-butene-air mixtures showed the presence of ionpeaks at 59, 73, 117, 131, 145, 159, 208, and 222 u.API-MS/MS “daughter ion” and “parent ion” spectrawere obtained for these and other ion peaks observedin the API-MS analyses. Product ion peaks were iden-tified based on the observation of homo- or hetero-dimer ions (for example, and1[(M ) 1 H] [M 1P1 2 P1

respectively, where P1 and P2 are prod-1M 1 H] ,P2

ucts) in the API-MS/MS “parent ion” spectra, and con-sistency of the API-MS/MS “daughter ion” spectrumof a homo- or hetero-dimer ion with the “parent ion”spectra of the various ion peaks. The API-1[M 1 H]P

MS/MS spectra obtained indicated the presence ofproducts of molecular weight 58 (attributed to ace-tone), 72 (attributed to 2-methylpropanal), 100, and149. As an example of the spectra obtained, Figure 4shows the API-MS/MS “parent ion” spectrum of the

ion peak, with the dimer ions being labeled.59 uBased on the product identifications, the ion peaks ob-served in the API-MS spectra at 59, 73, 117, 131, 145,159, 208, and are then 1222 u [acetone 1 H] ,

1[2-methylpropanal 1 H] , [acetone 1 acetone 1[acetone 1 2-methylpropanal 1 H] [2-methyl-1 1H] , ,

GAS-PHASE REACTION OF OH RADICAL 585

JCK(Wiley) RIGHT INTERACTIVE

shortstandardlong

Figure 4 API-MS/MS CAD “parent ion” spectrum of theion peak observed in the API-MS spectrum of an ir-59 u

radiated CH3ONO9NO—3-methyl-1-butene—air mix-ture. The ion peaks at 145 and are attributed to frag-178 ument ions of the ion peak (the former being a loss of208 uHNO3).

Figure 5 API-MS/MS CAD “duaghter ion” spectrum of the weak ion peak observed in the150 uAPI-MS spectrum of an irradiated CH3ONO9NO—3-methyl-1-butene—air mixture. The fragmention peaks at 132 and are attributed to losses of H2O and HNO3, respectively, and the ion87 u 46 upeak is attributed to 1[NO ] .2

propanal 1 2-methylpropanal 1 H] [acetone 1 1001,1 H] [acetone 1 149 1 and [2-1 1, H] ,methylpropanal respectively.11 149 1 H] ,

The API-MS/MS “daughter ion” spectrum of theweak ion peak showed fragment ions at150 u 132 u(2H2O), (2HNO3), and (Fig. 5),187 u 46 u ([NO ] )2

suggesting that the molecular weight 149 product was

the hydroxynitrate (CH3)2CHCH(OH)CH2ONO2 and/or (CH3)2CHCH(ONO2)CH2OH formed in Reaction(6a) and its analog. The API-MS/MS “daughter ion”spectrum of the weak ion peak present in the101 uAPI-MS spectrum showed fragment ions at 83, 73, 71,59 (weak), 55, 43, 31, and and, apart from the29 urelative intensities of the peaks, was83, 73, and 71 usimilar to the API-MS/MS spectrum of the ion101 upeak observed in the OH radical-initiated reaction ofisoprene [17]. This molecular weight 100 product maybe HOCH2C(CH3)"CHCHO (see Reaction SchemeIII).

DISCUSSION

The formation yields of 2-methylpropanal measuredfrom the GC-FID and FT-IR analyses (0.54 6 0.05and respectively) differ by ca. 15%, but0.63 6 0.05,are in agreement within the combined overall uncer-tainties. The reason for this difference is not knownbut could be due in part to uncertainties in the rateconstant ratio k(OH 1 2-methylpropanal)/k(OH 1

which is needed to calculate the3-methyl-1-butene)correction factors to account for secondary reactionsof 2-methylpropanal [a 10% increase in the rate con-stant ratio k(OH 1 2-methylpropanal)/k(OH 1

increases the maximum multiplic-3-methyl-1-butene)ative correction factor for 2-methylpropanal by 4.2%for the experiments with GC-FID analyses and by1.1% for the experiments with FT-IR analyses]. The

586 ATKINSON, TUAZON, AND ASCHMANN

JCK(Wiley) LEFT INTERACTIVE

shortstandardlong

two yield determinations are averaged, resulting in aformation yield of 2-methylpropanal of 0.58 6 0.08.

Consistent with the reaction pathways shown in Re-action Schemes I– III, formation of 2-methylpropanalplus HCHO, acetone plus glycolaldehyde, and meth-acrolein plus HCHO account for 58 6 8%, 18 6

and of the overall reaction products,3%, 3.3 6 0.7%with the molar HCHO yield of being in0.70 6 0.06reasonable agreement with the sum of the 2-methyl-propanal and methacrolein yields To-(0.61 6 0.09).gether with the estimated yield of “organic nitrates” ofca. 0.15 during the initial portions of the experiments,94% of the total reaction products and reaction path-ways of the OH radical-initiated reaction of 3-methyl-1-butene in the presence of NO are accounted for. Fur-thermore, as shown by the estimated reaction rates[32] of the major alkoxy radicals involved in the OHradical-initiated reaction of 3-methyl-1-butene in thepresence of NO, the observed products are consistentwith the calculated dominant reaction pathways ofthese alkoxy radicals. No evidence for the formationof methyl vinyl ketone or 3-methyl-2-butenal was ob-tained, and conservative upper limits to the formationyields of these two potential H-atom abstraction path-way products (Reaction Scheme III) are given in TableI. The formation of organic nitrates during the earlystages of the reaction is anticipated to arise from thereaction of the organic peroxy radicals [including(CH3)2CHCH CH2OH, (CH3)2CHCH(OH)CH2 2,(O BO) BO(CH3)2C CH"CH2, and (CH3)2C"CHCH2 2](O BO) BOwith NO (Reaction (6a) and analogous reactions). Theapproximate organic nitrate formation yield of 0.15 isreasonably consistent with organic nitrate formationyield measurements for the OH radical-initiated reac-tions of C4 9C6 alkenes in the presence of NO [36–38], with reported organic nitrate yields of 0.037 6

for the cis-2-butene reaction [37] and ca. 0.150.009for the 2,3-dimethyl-2-butene reaction [36,38].

The API-MS and API-MS/MS analyses areconsistent with the GC-FID and FT-IR analyses,with the major products being attributed to acetone,2-methylpropanal and the b-hydroxynitrates(CH3)2CHCH(ONO2)CH2OH and/or(CH3)2CHCH(OH)CH2ONO2. The API-MS analysesalso suggest the presence of a product (or products) ofmolecular weight 100, possibly HOCH2C(CH3)"CHCHO and/or CH2 "C(CH3)C(O)CH2OH formedsubsequent to the H-atom abstraction pathway (Re-action Scheme III).

Therefore, our data show that at room temperatureand atmospheric pressure, OH radical addition domi-nates for 3-methyl-1-butene, with H-atom abstractionaccounting for a minimum of 3.3% (based on the for-mation yield of methacrolein). The OH radical addi-

tion pathway accounts for of the overall91 6 12%OH radical reaction with 3-methyl-1-butene. Takinginto account the possible formation ofHOCH2C(CH3)"CHCHO [or possibly CH2 "C(CH3)C(O)CH2OH; see Reaction Scheme III] as sug-gested by our API-MS data and noting our upper limitsto the formation of methyl vinyl ketone and 3-methyl-2-butenal of ,1% each, a reasonable estimate of theH-atom abstraction pathway in the reaction of the OHradical with 3-methyl-1-butene is 5–10%. Combinedwith previous literature data [4,8,9,11], this suggeststhat H-atom abstraction from the allylic C9H bondsof alkenes during the OH radical reactions is of minorimportance, accounting for only a few percent of theoverall reactions under tropospheric conditions.

The authors gratefully thank the U.S. Environmental Pro-tection Agency, Office of Research and Development (As-sistance Agreement R-825252-01-0) for supporting this re-search, and thank the National Science Foundation (GrantNo. ATM-9015361) and the University of California, Riv-erside, for funds for the purchase of the SCIEX API III MS/MS instrument. While this research has been supported bythe U.S. Environmental Protection Agency, it has not beensubjected to Agency review and, therefore, does not neces-sarily reflect the views of the Agency, and no official en-dorsement should be inferred.

BIBLIOGRAPHY

1. R. L. Seila, W. A. Lonneman, and S. A. Meeks, Deter-mination of C2 to Ambient Air Hydrocarbons in 39C12

U.S. Cities, from 1984 through 1986, U.S. Environmen-tal Protection Agency Report EPA/600/3-89/058, At-mospheric Research and Exposure Assessment Labo-ratory, Research Triangle Park, North Carolina,September 1989.

2. F. W. Lurmann and H. H. Main, Analysis of the AmbientVOC Data Collected in the Southern California AirQuality Study, Final Report to California Air ResourcesBoard Contract No. A832-130, Sacramento, CA, Feb-ruary 1992.

3. W. L. Chameides, F. Fehsenfeld, M. O. Rodgers, C.Cardelino, J. Martinez, D. Parrish, W. Lonneman, D. R.Lawson, R. A. Rasmussen, P. Zimmerman, J. Green-berg, P. Middleton, and T. Wang, J. Geophys. Res., 97,6037 (1992).

4. R. Atkinson, J. Phys. Chem. Ref. Data, Monograph 2,1 (1994).

5. R. Atkinson, J. Phys. Chem. Ref. Data, 26, 215 (1997).6. N. J. Blake, S. A. Penkett, K. C. Clemitshaw, P. Anwyl,

P. Lightman, A. R. W. Marsh, and G. Butcher, J. Geo-phys. Res., 98, 2851 (1993).

7. R. Atkinson, Gas Phase Tropospheric Chemistry of Or-ganic Compounds, In Volatile Organic Compounds in

GAS-PHASE REACTION OF OH RADICAL 587

JCK(Wiley) RIGHT INTERACTIVE

shortstandardlong

the Atmosphere, Issues in Environmental Science andTechnology, 4, Eds. R. E. Hester and R. M. Harrison,The Royal Society of Chemistry, Cambridge, UK, 1995,pp. 65–89.

8. R. Atkinson, J. Phys. Chem. Ref. Data, Monograph 1,1 (1989).

9. H. Niki, P. D. Maker, C. M. Savage, and L. P. Breiten-bach, J. Phys. Chem., 82, 135 (1978).

10. H. Niki, P. D. Maker, C. M. Savage, and L. P. Breiten-bach, Chem. Phys. Lett., 80, 499 (1981).

11. R. Atkinson, E. C. Tuazon, and W. P. L. Carter, Int. J.Chem. Kinet., 17, 725 (1985).

12. E. C. Tuazon and R. Atkinson, Int. J. Chem. Kinet., 22,1221 (1990).

13. S. E. Paulson, R. C. Flagan, and J. H. Seinfeld, Int. J.Chem. Kinet., 24, 79 (1992).

14. S. E. Paulson and J. H. Seinfeld, Environ. Sci. Technol.,26, 1165 (1992).

15. A. Miyoshi, S. Hatakeyama, and N. Washida, J. Geo-phys. Res., 99, 18779 (1994).

16. R. Atkinson, E. C. Tuazon, and S. M. Aschmann, En-viron. Sci. Technol. 29, 1674 (1995).

17. E. S. C. Kwok, R. Atkinson, and J. Arey, Environ. Sci.Technol. 29, 2467 (1995).

18. E. S. C. Kwok, R. Atkinson, and J. Arey, Environ. Sci.Technol. 30, 1048 (1996).

19. G. P. Smith, Int. J. Chem. Kinet., 19, 269 (1987).20. K. Hoyermann and R. Sievert, Ber. Bunsenges. Phys.

Chem., 83, 933 (1979).21. K. Hoyermann and R. Sievert, Ber. Bunsenges. Phys.

Chem., 87, 1027 (1983).22. D. F. McMillen and D. M. Golden, Ann. Rev. Phys.

Chem., 33, 493 (1982).23. J. A. Kerr, Strengths of Chemical Bonds, in Handbook

of Chemistry and Physics, 74th Ed., Ed. D. R. Lide,

CRC Press, Boca Raton, FL, 1993–94, pp. 9-123 to 9-145.

24. R. Atkinson, W. P. L. Carter, A. M. Winer, and J. N.Pitts, Jr., J. Air Pollut. Control Assoc., 31, 1090 (1981).

25. R. Atkinson and S. M. Aschmann, Environ. Sci. Tech-nol., 27, 1357 (1993).

26. S. M. Aschmann and R. Atkinson, Environ. Sci. Tech-nol., 28, 1539 (1994).

27. E. C. Tuazon and R. Atkinson, Int. J. Chem. Kinet., 21,1141 (1989).

28. R. Atkinson, E. S. C. Kwok, J. Arey, and S. M. Asch-mann, Faraday Discuss. Chem. Soc., 100, 23 (1995).

29. E. S. C. Kwok, J. Arey, and R. Atkinson, J. Phys. Chem.100, 214 (1996).

30. E. S. C. Kwok and J. Arey, to be submitted for publi-cation (1998).

31. W. D. Taylor, T. D. Allston, M. J. Moscato, G. B. Fa-zekas, R. Kozlowski, and G. A. Takacs, Int. J. Chem.Kinet., 12, 231 (1980).

32. R. Atkinson, Int. J. Chem. Kinet., 29, 99 (1997).33. E. S. C. Kwok and R. Atkinson, Atmos. Environ., 29,

1685 (1995).34. R. Atkinson, S. M. Aschmann, W. P. L. Carter, A. M.

Winer, and J. N. Pitts, Jr., J. Phys. Chem., 86, 4563(1982).

35. N. Tsalkani and G. Toupance, Atmos. Environ., 23,1849 (1989).

36. H. Niki, P. D. Maker, C. M. Savage, L. P. Breitenbach,and M. D. Hurley, J. Phys. Chem., 91, 941 (1987).

37. K. Muthuramu, P. B. Shepson, and J. M. O’Brien, En-viron. Sci. Technol., 27, 1117 (1993).

38. E. C. Tuazon, S. M. Aschmann, J. Arey, and R. Atkin-son, Environ. Sci. Technol., in press (1998).