Kinetic Study of Gas-Phase Reactions of Pinonaldehyde and Structurally Related Compounds

advertisement
36257_5306_p527-533
5/21/97 8:06 AM
36257
Page 527
(Black plate)
Journal of Chemical Kinetics
(Wiley-Interscience)
art
5306 5/7/97 bw
Kinetic Study of Gas-Phase
Reactions of Pinonaldehyde
and Structurally Related
Compounds
M. GLASIUS, A. CALOGIROU*, N. R. JENSEN, J. HJORTH, C. J. NIELSEN**
European Commission, Joint Research Centre, Environmental Institute, TP 272, I-21020 Ispra (VA), Italy
Received 9 August 1996; accepted 25 November 1996
ABSTRACT: Rate constants for the reactions of OH, NO3, and O3 with pinonaldehyde and
the structurally related compounds 3-methylbutanal, 3-methylbutan-2-one, cyclobutylmethylketone, and 2,2,3-trimethyl-cyclobutyl-1-ethanone have been measured at 300 6 5 K
using on-line Fourier transform infrared spectroscopy. The rate constants obtained for
the reactions with pinonaldehyde were: kOH 5 (9.1 6 1.8) 3 10211 cm3 molecule21 s21,
kNO3 5 (5.4 6 1.8) 3 10214 cm3 molecule21 s21, and kO3 5 (8.9 6 1.4) 3 10220 cm3 molecule21 s21. The results obtained indicate a chemical lifetime of pinonaldehyde in the
troposphere of about two hours under typical daytime conditions, [OH] 5 1.6 3 106
molecule cm23. © 1997 John Wiley and Sons, Inc. Int J Chem Kinet 29: 527 – 533, 1997.
INTRODUCTION
Monoterpenes (C10H16) from vegetation contribute an
important fraction of the total nonmethane hydrocarbons emitted to the troposphere [1 – 4]. In the troposphere they react with OH and NO3 radicals as
well as with O3 , leading to the formation of a variety
of products.
a-Pinene is one of the most abundant monoterpenes in the troposphere [2 – 4], and its gas-phase
oxidation reactions have been the subject of several
investigations [5 – 8]. Pinonaldehyde (cis-4-acetyl*Also affiliated with Institute of Ecological Chemistry and Environmental Analysis, Technical University of Munich, Germany
**On leave from Department of Chemistry, University of Oslo,
Norway
Correspondence to: N. R. Jensen
Contract grant Sponsor: European Commission
Contract grant Sponsor: Department of Chemistry, University
of Oslo (N)
© 1997 John Wiley & Sons, Inc.
CCC 0538-8066/97/070527-07
2,2-dimethyl-cyclobutyl-ethanal) has been found to
be one of the major products in the tropospheric oxidation of a-pinene [5 – 8]. The tropospheric fate of
pinonaldehyde is unknown, but, taking its structure
into account, only photolysis and reaction with OH
are expected to be of importance.
To elucidate the atmospheric degradation of pinonaldehyde we have measured rate constants for its
reactions of OH, NO3 , and O3 and the structurally related compounds 3-methylbutanal, 3-methylbutan-2one, cyclobutyl-methylketone, and 2,2,3-trimethylcyclobutyl-1-ethanone.
EXPERIMENTAL
All the experiments were performed in purified air at
740 6 5 torr and 300 6 5 K in a 480 L Teflon-coated
reaction chamber surrounded by photolysis lamps
36257_5306_p527-533
36257
5/21/97 8:06 AM
Page 528
Journal of Chemical Kinetics
528
(Black plate)
(Wiley-Interscience) art 5306
2/13/96 jr
5/7/97 bw
GLASIUS ET AL.
(black lamps, l $ 300 nm) and equipped with a multiple reflection mirror system for on-line Fourier
Transform Infrared Spectroscopy (FT-IR) with a total
optical pathlength of 81 m. In addition, a low pressure mercury lamp (l 5 254 nm) has been introduced inside of the reaction chamber to photolyze
H2O2 . A detailed description of the experimental system has been previously published [9]. The spectra
were obtained with a Bruker IFS 113v at 1 cm21
nominal resolution by co-adding 20 to 50 scans.
The OH radicals were formed by photolysis of either CH3ONO or H2O2 . CH3ONO was synthesized by
adding H2SO4 (50 wt % aqueous solution) to sodium
nitrite dissolved in a methanol/water mixture [10].
Pure gaseous H2O2 was obtained by heating a commercial Urea-H2O2 complex to ca. 70°C. Ozone was
prepared by silent discharge of pure oxygen. NO3
radicals were formed by thermal dissociation of
N2O5 , which was prepared in solid form according to
a previously described procedure [11]. In a typical
experiment the temperature was 295 6 3 K in the beginning of an experiment; but when CH3ONO was
used as the precursor for OH radicals the temperature
increased by 3 – 7 K due to the heating from the UVlamps.
The rate constants for the ozone reactions were determined under pseudo-first-order conditions:
k
A 1 O3 9: Products
(1)
At 5 Aoexp{2k[O3]t}
(2)
ln At 5 const. 2 (k[O3]t)
(3)
5 const. 2 (k9t)
where k is the reaction rate constant for the reaction
between compound A and O3 , and k9 5 k[O3].
With O3 in large excess its concentration will be
approximately constant during an experiment and
pseudo-first-order conditions are obtained. A plot of
lnAt vs. time t will then give pseudo-first-order decay
k9. A plot of k9 vs. [O3] will subsequently give the reaction rate constant k as the slope. Typical experimental conditions during the O3 experiments were:
50 ppmV , [O3] , 400 ppmV, 1 ppmV , [A] , 10
ppmV, but [O3] was always , 10 times [A].
The OH and NO3 reaction rate constants were determined by the “relative rate” method:
k
A
A 1 OH 9:
Products
kB
B 1 OH 9: Products
(4)
(5)
For these two bimolecular reactions the following expression for the ratio kA/kB can be obtained:
Figure 1 Plot of ln{[pinonaldehyde]o /[pinonaldehyde]t},
corrected for losses, vs. ln{[1,3-butadiene]o / [1,3butadiene]t} for the decay of pinonaldehyde and 1,3-butadiene during photolysis of methylnitrite producing OH radicals.
kA/kB 5 ln(Ao /At)/ln(Bo /Bt)
(6)
Thus when kB is known, kA can be calculated.
In the case that reactant A has an additional firstorder loss process described by a rate constant kw , eq.
(6) is modified as shown in eq. (7):
kA/kB 5 {ln(Ao /At) 2 (kwt)}/ln(Bo /Bt)
(7)
A plot of ln(Ao /At) 2 (kw t)} vs. ln(Bo /Bt) will give
the ratio kA/kB as the slope (see Fig. 1). Typical
experimental conditions: 10 ppmV , [CH3ONO]
, 30 ppmV, 5 ppmV , [N2O5] , 100 ppmV, 1
ppmV , [A] , 10 ppmV, and 5 ppmV , [B] , 25
ppmV.
Pinonaldehyde was introduced into the reaction
chamber using a “double filter system” as explained
below: Pinonaldehyde is difficult to evaporate, because it has a very low vapor pressure and decomposes easily when heated; thus we added 10 to 20 ml
of pinonaldehyde to the first of two filters in series in
a Teflon tube and a minor fraction of it was blown
into the reaction chamber by a stream of air. The first
filter was used to enhance the surface area of the
compound and the second filter was used to prevent
aerosols from being blown into the reaction chamber.
Figure 2 shows an infrared spectrum of pinonaldehyde. The spectrum resembles the infrared spectrum
previously published [8], which was tentatively identified as due to pinonaldehyde.
3-Methylbutanal, 3-methylbutan-2-one, cyclobutyl-methylketone, ethene, isoprene, 1,3-butadiene,
and 1-butene were added in known quantities to the
reaction chamber by standard manometric methods.
36257_5306_p527-533
5/21/97 8:06 AM
36257
Page 529
(Black plate)
Journal of Chemical Kinetics
(Wiley-Interscience) art 5306
2/13/96
jr 5/7/97 bw
GAS-PHASE REACTIONS OF PINONALDEHYDE
529
mercial samples of p.a. quality and were used without further purification: 3-Methylbutanal (97% pure,
Aldrich), 3-methylbutan-2-one (99% pure, Aldrich),
cyclobutyl-methylketone (98% pure, Aldrich), ethene
( $ 99.5% pure, Air Liquide), isoprene (99% pure,
Aldrich), 1,3-butadiene (99.5% pure, Ucar), and 1butene (99.9% pure, Ucar).
(a)
RESULTS AND DISCUSSION
(b)
Figure 2 (a), (b) Infrared spectrum of pinonaldehyde
from 700 to 3500 cm21, with spectral features at 1178,
1379, 1726, 1740, 2713, 2812, 2889, 2933, and 2962 cm21.
The spectral features from CO2 have been removed.
Chemicals
Pinonaldehyde was synthesized in small batches by
oxidative bond cleavage of pinandiol (99% pure,
Aldrich). A solution of 0.46 g of HIO4 ?2H2O in 40
mL H2O was brought to pH 5 4 – 5 by adding
Na2HPO4 and subsequently added in small portions
to a solution of 0.34 g of pinandiol in 6 mL tetrahydrofuran (THF) and 6 mL of H2O. The mixture was
left standing for two days, then the THF was removed
and the solution extracted with methyl-tertbutylether. Pinonaldehyde was finally obtained by
evaporation of the methyl-tert-butylether. For the purpose of this investigation no further purification was
necessary. The MS data obtained here for pinonaldehyde were in accordance with those of ref. [12].:
EI-MS m/z 168(M1, 0), 109(11), 98(29), 97(24),
83(100), 71(35), 69(80), and 55(56); CI-MS
(methane) m/z 169(M 1 H1, 7), 151(46), 107(40),
99(40), and 71(100).
2,2,3-trimethylcyclobutyl-1-ethanone was prepared as previously described [12] starting from
pinonaldehyde synthesized as described above.
EI-MS m/z 140(M1, 2), 125(10), 83(82), 71(18),
70(92), and 55(100); CI-MS (methane) m/z
141(M 1 H1, 22), 123(8), 111(12), 99(22), and
71(100).
Other chemicals used in this study were all com-
In this investigation we have determined the rate
constants for the reactions of OH, NO3 , and O3
with pinonaldehyde and the structurally related
compounds 3-methylbutanal, 3-methylbutan-2-one,
cyclobutyl-methylketone, and 2,2,3-trimethylcyclobutyl-1-ethanone.
In purified air pinonaldehyde showed a first-order
decay in our reaction chamber, which was attributed
to wall loss. This wall loss varied somewhat with the
condition or “history” of the cell and with temperature. On the average, the rate constant for the wall
loss first-order decay of pinonaldehyde was
kwall-loss 5 (2.9 6 0.3) 3 1025 s21 (uncertainty with
2 s). With the external photolysis lamps turned on,
l $ 300 nm, the decay of pinonaldehyde was enhanced. The average photolysis and wall loss constant
of pinonaldehyde in our system, JUV/VIS 1 kwall , was
found to be between 1 3 1024 s21 and 2 3 1024 s21
(when UV-light $ 300 nm was used), which correspond to up to 25% of the overall decay of pinonaldehyde during an experiment with OH radicals. The
reference compounds and the other compounds studied were all very stable in the cell with
kwall-loss , 1 3 1025 s21, even with the photolysis
lamps turned on.
The analysis of a relative rate measurement of the
OH reaction with a pinonaldehyde/1,3-butadiene
mixture is presented in Figure 1. In the example
shown the relative concentration measurements were
obtained by spectral subtraction using the C — Hald
stretch band of pinonaldehyde at 2650 – 2750 cm21
and the " C — H bend band of 1,3-butadiene at
850 – 950 cm21. The data were analyzed assuming additional first-order loss of pinonaldehyde to the cell
walls and first-order photolytical degradation as discussed above. In the spectral subtraction we also included the spectral features from the products
formaldehyde and acrolein. Table I lists the bands
that have been used in spectral subtraction for the individual compounds.
The rate constants for the reactions of OH, NO3 ,
and O3 with the different organics are presented
in Table II. The rate constant for the reaction between
36257_5306_p527-533
36257
5/21/97 8:06 AM
Page 530
Journal of Chemical Kinetics
530
(Black plate)
(Wiley-Interscience) art 5306
2/13/96 jr
5/7/97 bw
GLASIUS ET AL.
Table I Bands used for Spectral Subtraction
Band
(cm21)
Pinonaldehyde
2,2,3-Trimethylcyclobutyl-1-ethanone
3-Methylbutanal
3-Methylbutan-2-one
Cyclobutyl-methylketone
Isoprene
1,3-Butadiene
1-Butene
Ethene
Vibration
2650 – 2750
1150 – 1210
2600 – 2850
1070 – 1170
1125 – 1225
3050 – 3150
850 – 950
3050 – 3150
850 – 1050
C — Hald.
C—C
C — Hald.
C — Hald.
C—C
C — Halk.
"C—H
C — Halk.
"C—H
stretch
stretch
stretch
stretch
stretch
stretch
out-of plane bend
stretch
out-of plane bend
Interfering bands from products produced in the experiments, such as formaldehyde, metacrolein, and
acrolein, have been taken into account by spectral subtraction.
pinonaldehyde and OH was obtained from two series
of experiments using isoprene with kOH 5
1.01 3 10210 cm3 molecule21 s21 at 298 K [13]
and 1,3-butadiene with kOH 5 6.66 3 10211 cm3
molecule21 s21 at 298 K [13] as reference compounds, respectively. There is no significant difference between the rate constants obtained using the
two different reference compounds and only the average value is given in Table I. The influence of a possible variation in the wall and photolysis loss of pinonaldehyde during the experiments was estimated by
letting JUV/VIS 1 kwall loss for pinonaldehyde vary
within its statistical 95% confidence interval; the resulting variation was less than 10% of the average
rate constant. Typical plots for pinonaldehyde 1 NO3
and for pinonaldehyde 1 O3 are shown by Figures 3
and 4, respectively. These plots are corrected for wall
loss, but in Figure 4 it can be seen that the slope does
not intercept with {0 ; 0}. It is tentatively suggested
that an additional surface reaction with ozone may be
the reason.
The rate constant for the OH reaction with 3methyl-butanal has previously been measured as
(2.7 6 0.1) 3 10211 cm3 molecule21 s21 [14] (see
Table III) with the use of GC-FID as their analytical
technique. As can be seen the value reported here of
(4.0 6 0.7) 3 10211 cm3 molecule21 s21 obtained
using isoprene with kOH 5 1.01 3 10210 cm3
molecule21 s21 [13] as the reference compound is significantly larger. No explanation for this discrepancy
between the two values can be given at this time.
It was not possible to find any IR bands unique
to the organic reactants in the spectra obtained from
the OH reaction with 2,2,3-trimethylcyclobutyl-1ethanone, cyclobutyl-methyl-ketone and 3-methylbutan-2-one. Changes in the IR band shapes showed
Table II Rate Constants Determined at 740 6 5 torr and 300 6 5 K (Units in cm3 molecule21 s 21) in this
Investigation
R
e
a
c
t
a
n
t
OH
NO3
O3
CO
CO
CO
CHO
CO
CHO
Pinonaldehyde
3-Methylbutanal
9.1 6 1.8 3 10211a
5.4 6 1.8 3 10214
8.9 6 1.4 3 10220
4.0 6 0.7 3 10211
1.2 6 0.3 3 10214
3.9 6 0.6 3 10220
2,2,3Trimethylcyclobutyl- Cyclobutylmethyl1-ethanone
ketone
n.d.b
< 5 3 10216
< 5 3 10221
n.d.
< 5 3 10216
< 5 3 10221
3-Methylbutan2-one
n.d.
< 5 3 10216
6.3 6 1.3 3 10221
The uncertainties are given as 2 times the standard deviation.
a
For this reaction two reference compounds were used and we obtained the following results: When isoprene was used as reference compound the ratio (0.896 6 0.260 ) was obtained, corresponding to k pinonaldehyde1OH 5 (9.05 6 2.62) 3 10211 cm3 molecule21 s21
(kisoprene1OH 5 1.01 3 10210 cm3 molecule21 s 21 [13]) and with 1, 3-butadiene as reference compound the ratio (1.37 6 0.19) was obtained,
corresponding to (kpinonaldehyde1OH 5 (9.18 6 1.25) 3 1021 cm3 molecule21 s21 (k1,3-butadiene1OH 5 6.66 3 10211 cm3 molecule21 s21 [13]).
b
n.d.: Not determined, see text.
36257_5306_p527-533
5/21/97 8:06 AM
36257
Page 531
(Black plate)
Journal of Chemical Kinetics
(Wiley-Interscience) art 5306
2/13/96
jr 5/7/97 bw
GAS-PHASE REACTIONS OF PINONALDEHYDE
Figure 3 Plot of ln{[pinonaldehyde]o /[pinonaldehyde]t},
corrected for losses (kwt), vs. ln{[1-butene]o /[1-butene]t for
the decay of pinonaldehyde and 1-butene during reactions
with the NO3 radical.
that the compounds reacted, but it was not possible to
determine the corresponding rate constants.
The reactions of O3 and NO3 with 2,2,3-trimethylcyclobutyl-1-ethanone and cyclobutyl-methylketone
are all very slow and only conservative upper estimates for these rate constants are presented. Reference values for the NO3 reactions used in these investigations were: k(NO3 1 1-Butene) 5 1.25 3
10214 cm3 molecule21 s21 [15] and k(NO3 1
Ethene) 5 2.0 3 10216 cm3 molecule21 s21 [15]. The
values presented here for the reaction of NO3 with
pinonaldehyde and 3-methylbutanal, 5.4 and
1.2 3 10214 cm3 molecule21 s21, respectively, are one
to two orders of magnitude higher than the values for
formaldehyde and acetaldehyde, 6.8 3 10216 and
2.8 3 10215 cm3 molecule21 s21, respectively. Taking into account the rate constants for higher homo-
531
logues of acetaldehyde [16] our values lie well within
the trend of increasing rate constants by increasing
carbon atom number. The present knowledge about
the mechanism of the NO3 radical reaction with aldehydes gives no satisfactory explanation for this trend.
The reaction between ozone and 3-methyl-butan2-one is somewhat faster than those of ozone with
2,2,3-trimethylcyclobutyl-1-ethanone and cyclobutylmethylketone, and the observed loss of organic is apparently reasonably well correlated with the ozone
concentration in the reaction chamber.
It has been suggested that the reactivity of the OH
radical towards organics, including the oxygenated
hydrocarbons presently studied, may be estimated
from a simple structure-reactivity relationship [17].
This incremental reactivity method predicts a rate
constant for the OH reaction of 2.4 3 10211 cm3
molecule21 s21 for both pinonaldehyde and 3methylbutanal, which is significantly lower than our
values of 9.1 and 4.0 3 10211 cm3 molecule21 s21,
respectively (see Table III). The rate constants for the
OH reaction with the three ketones investigated are
estimated by the same method to range between 1.5
and 3 3 10212 cm3 molecule21 s21, i.e., an order of
magnitude lower than for pinonaldehyde and 3methylbutanal. All of the structure elements in pinonaldehyde and 3-methylbutanal are included in the
database used in the derivation of the structure-reactivity relationship [17]. The incremental reactivity
method is claimed to be reliable within a factor of 2
for ca. 90% of organic compounds belonging to the
classes used in the database development, but pinonaldehyde may very well belong to the 10% “ill-behaving” molecules. Table III also includes the predicted OH rate constants for a structure-reactivity
method based upon molecular orbital calculations
[18 – 20]: The values for the aldehydes, 2.6 3 10211
and 2.2 3 10211 cm3 molecule21 s21 for pinonaldehyde and 3-methylbutanal, respectively, are similar to
those calculated by the incremental reactivity
method, showing however, a trend towards a faster reaction for pinonaldehyde. For the ketone, 2,2,3trimethylcyclobutyl-1-ethanone, a rate constant of
1.4 3 10211 cm3 molecule21 s21 is predicted, which
is about one order of magnitude higher than the rate
constant obtained by the incremental reactivity
method.
CONCLUSIONS AND ATMOSPHERIC
IMPLICATIONS
Figure 4 Plot of k9 vs. [O3] for the reaction between
pinonaldehyde and O3 .
This study shows that the chemical lifetime of pinonaldehyde in the atmosphere is essentially determined
36257_5306_p527-533
36257
5/21/97 8:06 AM
Page 532
Journal of Chemical Kinetics
532
(Black plate)
(Wiley-Interscience) art 5306
2/13/96 jr
5/7/97 bw
GLASIUS ET AL.
Table III Measured and Estimated OH Reaction Rate Constants (Units in cm3
molecule21 s 21)
CO
CO
CHO
This work
(Experimental)
Kwok and Atkinson [17]
(Estimated)
Klamt [18 – 20]
(Estimated)
Kerr and Sheppard [14]
(Experimental)
a
CHO
(9.1 6 1.8) 3 10211
(4.0 6 0.7) 3 10211
n.d.a
2.4 3 10211
2.4 3 10211
2.4 3 10212
2.6 3 10211
2.2 3 10211
1.4 3 10211
n.d.a
(2.7 6 0.1) 3 10211
n.d.a
n.d.: Not determined, see text.
by the reaction with the OH radical. The reaction
with the NO3 radical is only of minor importance
even in polluted areas, whereas the reaction with O3
is of no importance. The present results indicate an
atmospheric chemical lifetime of pinonaldehyde of
about two hours, assuming [OH] 5 1.6 3 106
molecules cm23 [21].
It is concluded from the results shown in Table II
and from the estimates of the OH radical reaction
rates in Table III, that the reactivity of pinonaldehyde
with the NO3 radical and with O3 is mainly linked to
the aldehyde group, because the reactions rates of
structurally similar compounds without the aldehyde
group are significantly slower as shown by Table II.
For the reactions of pinonaldehyde with OH radicals
other sites like the two tertiary H-atoms could be of
importance, see Table III under Klamt [18 – 20], who
does not calculate a significant difference between
the rates of the reaction of pinonaldehyde with OH
and that of the structurally similar compound with
OH.
To our knowledge, pinonaldehyde has only been
tentatively identified at low pptV levels in ambient air
[22]. This is somewhat puzzling considering that laboratory studies have shown pinonaldehyde to be a
major product from oxidation of a-pinene [5 – 8].
Hence, we have estimated the daytime steady-state
concentration of pinonaldehyde in the atmosphere using k(OH+pinonaldehyde) obtained in this study, and
assuming that photolysis and heterogeneous losses
are unimportant sinks. Pinonaldehyde is formed by
the oxidation of a-pinene by OH radicals with a yield
of approximately 60% in an NOx-containing atmosphere [7]. The yield of pinonaldehyde from
the ozonolysis of a-pinene is not known so we consider the two extreme cases where the yield is
0% and 100%, respectively. By equating sources
and sink terms of pinonaldehyde (i.e., applying
the
steady-state
assumption)
and
using
k(OH 1 a-pinene) 5 5.37 3 10211 cm3 molecule21
s21 [13], k(O3 1 a-pinene) 5 8.5 3 10217 cm3
molecule21 s21 [23], [OH] 5 1.6 3 106 molecules
cm23, and O3 5 1.2 3 1012 molecules cm23 (50
ppbV), ratios of [pinonaldehyde] : [a-pinene] of 0.35
and 1.05, respectively, are obtained. This means that
under typical ambient conditions, pinonaldehyde
should be detected at low ppbV levels and not at low
pptV levels [22]. From the simple estimate given, we
conclude that other effective atmospheric sinks for
pinonaldehyde, like heterogenic losses or photolytic
degradation, probably exist.
It should be mentioned that other terpenes in the
atmosphere are also oxidized to keto-aldehydes like
pinonaldehyde [8,24] and therefore the fate of pinonaldehyde may give an indication of the fate of other
keto-aldehydes in the troposphere.
Further work on the atmospheric lifetime and fate
for other terpene oxidation products is in progress.
A. C. would like to thank the European Commission for a
Ph.D. grant. C. J. N. acknowledges financial support from
the Department of Chemistry, University of Oslo (N), during his stay at the JRC, EI. The authors would like to thank
Prof. C. Lohse and Dr. D. Kotzias for helpful scientific discussions. The authors gratefully acknowledge the help of
G. Ottobrini with the experimental work.
BIBLIOGRAPHY
1. I. S. A. Isaksen and Ø. Hov, Tellus, 39b, 271(1987).
2. A. Guenther, C. N. Hewitt, D. Erickson, R. Fall, C.
36257_5306_p527-533
5/21/97 8:06 AM
36257
Page 533
(Black plate)
Journal of Chemical Kinetics
(Wiley-Interscience) art 5306
2/13/96
jr 5/7/97 bw
GAS-PHASE REACTIONS OF PINONALDEHYDE
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
Geron, T. Graedel, P. Harley, L. Klinger, M. Lerdau,
W. A. McKay, T. Pierce, B. Scholes, R. Steinbrecher,
R. Tallamraju, J. Taylor, and P. Zimmerman, J. Geophys. Res., 100, 8873(1995).
P. R. Zimmerman, R. B. Chatfield, J. Fishmann, P. J.
Crutzen, and P. L. Hanst, Geophys. Res. Lett., 5,
679(1978).
R. A. Rasmussen, J. Air. Pollut. Control Assoc., 22,
537(1972).
S. Hatakeyama, K. Izumi, T. Fukuyama, and H. Akimoto, J. Geophys. Res., 94, 13013(1989).
J. Arey, R. Atkinson, and S. A. Aschmann, J. Geophys.
Res., 95, 18539(1989).
S. Hatakeyama, K. Izumi, T. Fukuyama, H. Akimoto,
and N. Washida, J. Geophys. Res., 96, 947(1991).
H. Hakola, J. Arey, S. M. Aschmann, and R. Atkinson,
J. Atm. Chem., 18, 75(1994).
N. R. Jensen, J. Hjorth, C. Lohse, H. Skov, and G.
Restelli, Atmos. Environ., 25A, 1897(1991).
W. D. Taylor, T. D. Allston, M. J. Moscato, G. B.,
Fazekas, R. Kozlowski, and G. A. Takacs, Int. J. Chem.
Kinet., 12, 231(1980).
G. Schott and N. Davidson, J. Am. Chem. Soc., 80,
1841(1958).
V. Rautenstrauch, B. Willhalm, W. Thommen, and G.
Ohloff, Helv. Chim. Acta, 67, 325(1984).
533
13. R. Atkinson, J. Phys. Chem. Ref. Data, Monograph
no. 1 (1989).
14. J. A. Kerr and D. W. Sheppard, Environ. Sci. Technol.,
15, 960(1981).
15. R. P. Wayne, I. Barnes, P. Biggs, J. P. Burrows, C. E.
Canosa-Mas, J. Hjorth, G. LeBras, G. K. Moortgat, D.
Perner, G. Poulet, G. Restelli, and H. Sidebottom,
Commission of the European Communities: Air Pollution Research Report 31, (1991).
16. C. J. Nielsen, private communication, 1996.
17. E. S. Kwok and R. Atkinson, Atmos. Environ., 29,
1685(1995).
18. A. Klamt, Chemosphere, 26, 1273(1993).
19. A. Klamt, Chemosphere, 32, 717(1996).
20. A. Klamt and J. Lohrenz, BAYER AG. private communication, 1996.
21. R. Prinn, D. Cunnold, P. Simmonds, F. Alyea, R. Boldi,
A. Crawford, P. Fraser, D. Guntzler, D. Hartley, R.
Rosen, and R. Rasmussen, J. Geophys. Res., 97,
2445(1992).
22. A. Calogirou, M. Duane, A. Kettrup, D. Kotzias, M.
Lahaniati, and B. R. Larsen, Atmos. Environ., submitted.
23. R. Atkinson, Atmos. Environ., 24A, 1(1990).
24. A. Calogirou, D. Kotzias, and A. Kettrup, Atmos. Environ., 31, 283(1997).
36257_5306_p527-533
36257
5/21/97 8:06 AM
Page 534
Journal of Chemical Kinetics
(Black plate)
(Wiley-Interscience) art 5306
2/13/96 jr
5/7/97 bw
534 Blank
Download