You are on page 1of 6

Arabian Journal of Chemistry (2017) 10, S665–S670

King Saud University

Arabian Journal of Chemistry


www.ksu.edu.sa
www.sciencedirect.com

ORIGINAL ARTICLE

A new look at the b-pinene–ozone reaction using


the atmospheric pressure reactor
Fadel Alwedian *

Department of Chemistry, Faculty of Science, Tafila Technical University, P.O. Box 179, Tafila 66110, Jordan

Received 8 April 2012; accepted 13 November 2012


Available online 23 November 2012

KEYWORDS Abstract b-Pinene–ozone reaction without a scavenger was performed for the first time, in the sta-
Static bag reactor; tic bag atmospheric pressure reactor coupled into ion trap mass spectrometer. The diffusion coef-
Gas-phase reactions; ficients and evolution times (t10–90%) of multiple fragments were used to characterize their neutral
Ozone; parent ions. Mainly identified products (yields in parentheses) of the reaction such as nopinone
b-Pinene; (0.58 ± 0.05), formaldehyde (0.19 ± 0.05), HCOOH, CH3COOH, and acetone were successfully
Evolution time associated with their fragment ions. While, 3-hydroxy-nopinone and cis-pinic acid were unsuccess-
fully associated with their fragments because of the lack of available standard reference mass spec-
tra of these compounds for comparison. Meanwhile, two short-lived intermediates were detected
with tentative contributions of products of masses 70 and 97.
ª 2012 Production and hosting by Elsevier B.V. on behalf of King Saud University. This is an open access
article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/3.0/).

1. Introduction leading to the formation of photochemical air pollution in


urban and rural areas (Atkinson and Arey, 1998; Seinfeld
Alkenes are an important fraction of nonmethane hydrocar- and Pandis, 1998).
bons which are introduced in the atmosphere from anthropo- The kinetics and products of the gas phase reaction of ozone
genic and biogenic sources (Atkinson, 1997; Seinfeld and with alkene have been extensively studied. The ozone–alkene
Pandis, 1998). They are the major components in gasoline reaction proceeds via the addition of ozone to the alkene dou-
fuels, automobile exhaust emission, and ambient air in the ble bond, followed by the decomposition of resulting molozo-
highly urbanized areas. In troposphere, they react with O3, nide to yield carbonyl compounds (primary products) and
OH, and NO3 which result in a complex series of chemical initially energy-rich Criegee biradical, the Criegee biradical
transformations and physical sink reactions in the atmosphere can undergo unimolecular decomposition or isomerization to
yield the secondary products (Atkinson and Arey, 1998;
Seinfeld and Pandis, 1998). The kinetics for large ozone/alkene
* Mobile: +962 03 22 50 326; fax: +962 03 22 50 002. reactions are well understood, while there are less available
E-mail address: alwedfad@yahoo.com. data concerning mechanism and product of small alkenes under
Peer review under responsibility of King Saud University. atmospheric conditions (Atkinson, 1997; Atkinson and Arey,
1998; Guenther et al., 1995; Hasson et al., 2001; Seinfeld and
Pandis, 1998; Tuazon et al., 1998).
An important fraction of nonmethane hydrocarbons are
Production and hosting by Elsevier the monoterpenes (C10H16), a-pinene and b-pinene which
http://dx.doi.org/10.1016/j.arabjc.2012.11.006
1878-5352 ª 2012 Production and hosting by Elsevier B.V. on behalf of King Saud University.
This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/3.0/).
S666 F. Alwedian

are among the most abundant terpenes emitted into the 2. Experimental
atmosphere by vegetation (Atkinson and Arey, 1998; Atkin-
son and Aschmann, 1993; CalogIrou et al., 1999; Grosjean 2.1. The ion trap mass spectrometer
et al., 1993; Hakola et al., 1994; Guenther et al., 1995;
Hatakeyama et al., 1991; Librando and Tringali, 2005).
The reactor used in this work contained a flowing reacting
The gas-phase oxidation reactions of terpenes have been
atmospheric pressure gas mixture that was admitted to an ion
the subject of several investigations. Major products that re-
trap mass spectrometer through a flow restrictor for analysis.
sulted from a-pinene oxidation (Hatakeyama et al.,1989;
The custom reactor was coupled to a Varian Saturn 2000 ion
Hatakeyama et al., 1991; Librando and Tringali, 2005;
trap mass spectrometer via the 35.6 mm outer diameter direct
Venkatachari and Hopke, 2008) are pinonaldehyde, norpi-
insertion probe interface. The reactor interface was inserted into
nonaldehyde, formaldehyde, CO, and CO2; and nopinone
the vacuum adaptor installed in the GC transfer line inlet port.
and formaldehyde from the oxidation of b-pinene (CalogI-
The experiments reported here were conducted using the
rou et al., 1999; Hatakeyama et al., 1989; Hatakeyama
methane chemical ionization. The mass spectra (m/z 30–250)
et al., 1989; Jay and Stieglitz, 1987; Hakola et al., 1994;
were collected using the ion trap control software in the GC
Winterhalter et al., 2000). However, very little is known
mode. The WSearch reduction program was used to convert
about the atmospheric fate of these terpene oxidation prod-
the measured mass spectra to the ASCII form for further
ucts (Atkinson, 1997; CalogIrou et al., 1999; Librando and
analysis.
Tringali, 2005; Seinfeld and Pandis, 1998; Winterhalter
et al., 2000). Scheme 1 shows the chemical structures of a-
2.2. Static reactor setup
pinene, b-pinene, nopinone, pinonaldehyde, and
norpinonaldehyde.
The static bag atmospheric pressure reactor coupled into The reactor setup was described in a previously published work
ion trap mass spectrometer (ITMS) (Wedian and Atkinson, (Wedian and Atkinson 2010). Briefly, all reactions were carried
2010) is convenient to use because it permits continuous, sen- out in a 12 L Teflon bag. One side of the bag was connected to
sitive, and specific monitoring of the gas-phase reactions by mass spectrometer via a custom made interface while the other
the ITMS. The ITMS is potentially valuable as a detector side was connected to a 3-way valve. In one valve position, the
since it provides various clean chemical ionization modes valve admitted a He flow sufficient to supply the mass spec-
and Tandem MS/MS experiments which can be used for trometer interface and to maintain the bag shape. He served
structure determinations. In addition, the ITMS is an inex- as the gas medium for reactions as well as the carrier gas that
pensive bench top instrument. The static bag atmospheric delivered the reactants and products into the mass spectrome-
pressure reactor can provide kinetic measurements in the ter. In the second valve position, the bag was connected to a
range of 10–200 s, as well as calibration of product yield mea- diaphragm pump to evacuate the bag’s contents between suc-
surements (Wedian and Atkinson, 2010). cessive experiments. The vapor phase of b-pinene and ozone,
The objective of this work is to demonstrate the use of the was injected with a syringe through a hole in the side of the bag.
static pressure reactor to study the gas phase b-pinene–ozone
reaction in order to validate the detection of the primary and 2.3. Static reactor operation
secondary products and to detect new stable intermediates
and then identify their chemical formulae. The time evolutions The reactions were performed in this experiment by filling with
(t10–90%) of fragment ions of products in the static bag atmo- He then taking the background of the bag, after few minutes,
spheric pressure reactor coupled into ion trap mass spectrom- the b-pinene (initial calculated concentration of 6 ppmv) was
eter were determined and used to correlate the ions to their injected into the bag using a 25 ml syringe until the signal
distinct neutral species. reached the steady state, and then finally, excess amount of

(a) (b) (c)

(d) (e) (f)


Scheme 1 The chemical structures of a-Pinene and b-pinene and their major oxidation products: (a) a-pinene (b) b-pinene (c) nopinone
(d) pinonaldehyde (e) pinic acid (f) norpinonaldehyde.
A new look at the b-pinene–ozone reaction usingthe atmospheric pressure reactor S667

purified ozone in nitrogen was introduced into the bag using a 3. Results and discussion
50 mL syringe to complete the reaction. It is expected that the
primary and secondary products of ozonation are formed in 3.1. b-pinene–ozone reaction
few seconds. Moreover, it is expected all reactants and prod-
ucts were exposed to similar conditions of wall reactions, dilu-
Fig. 1A shows the mass spectrum of the reaction mixture at
tion, and transport of all products (Wedian and Atkinson
long reaction time, after subtracting the pre-reaction back-
2010). Between experiments, the reactor bag was flushed and
ground. The mass spectrum is dominated by ions of unreacted
filled several times with He to remove the contents from the
b-pinene and nopinone which produce strong peaks at m/z 57,
previous experiment. The ozone concentration in the reactor
59, 69, 83, 97, and 107 and a few weaker fragments. Fig. 1B
was determined in separate experiments by exhausting the
shows the mass spectrum of b-pinene which is dominated by
reactor through a Dasibi ozone instrument.
major fragments (m/z = 67, 81, 93, 107, and 121), Fig. 1C is
chemical ionization mass spectrum of pure Nopinone which
2.4. Chemicals
is dominated by the fragments (m/z = 55, 69, 81, 93, 97,
107, 135 and 137). The CI mass spectrum of pure pinene and
Chemicals and gases used in these experiments were: b-pinene nopinone was measured in separate experiments. Multiplying
(Fluka, 99.5%), acetone (Aldrich, 99%), nopinone (Aldrich, the reference b-pinene and nopinone CI spectra by an appro-
98%), formaldehyde (Aldrich, 36.5–38.0%), He (commercial priate weighting factor, and then subtracting them from the
grade 99.998, Airgas Inc.), Methane (UHP 99.9995, Airgas b-pinene–ozone reaction mixture (Fig. 1A) allowed the contri-
Inc.), N2 (UHP 99.9995, Airgas Inc.), and O2 (UHP 99.9995, bution of nopinone and un-reacted b-pinene to be removed
Airgas Inc.). and Fig. 1D shows the resulting mass spectrum. The other

69
100
57
59
A
83
Relative Intensity

80
60 45
40 31 97
107
20 177
0

81
100
93
B
Relative Intensity

80
60
40
107
121 136
20
0

81
100 C
Relative Intensity

80
93
60
40 109
57 138
20

350 69
300
57 D
250 59 83
Intensity

200 31
45
150
100 97 100
50
0
40 60 80 100 120 140 160 180
m/z

Figure 1 (A) The mass spectrum of the b-pinene–ozone reaction mixture after subtraction of the ozone background. The spectrum is
dominated by b-pinene and nopinone. (B) The mass spectrum of b-pinene. (C) The mass spectrum of nopinone. (D) The net mass
spectrum of the b-pinene–ozone mixture after subtracting contributions of ozone, pure b-pinene, and pure nopinone.
S668 F. Alwedian

200 80
A B
150 60

Intensity
m/z 57
100 40

m/z 61
m/z 93
50 20

0 0
0 50 100 150 200 250 300 0 50 100 150 200 250 300
160 150
C D
120 m/z 57
100
Intensity

80 m/z 55
m/z 83
50
40 m/z 60
m/z 96

0 0
0 50 100 150 200 250 300 0 50 100 150 200 250 300
m/z m/z

Figure 2 Function (1 exp ( a1t)) a2, with adjustable coefficients, used to fit the initial growth of product ions of b-pinene /O3 in the
static bag reactor: (A) the growth ion profile of m/z 57 and 93 of nopinone. (B) The growth ion profile of m/z 61. (C) The growth ion
profile of m/z 55 and 60. (D) The growth ion profiles of m/z 57, 83, and 96.

Table 1 Major ions of the b-pinene–ozone reaction observed Table 2 The molar masses of the reported primary and
by the ion trap using CH4 chemical ionization, from Fig. 1A secondary products of the b-pinene–ozone reaction Winterhal-
and D, listed in order of decreasing rise coefficients. ter et al. (2000).
m/z Rise Tentative assignment Compound name Molar mass (g/mol) Chemical formula
coefficient a, s 1
of neutral parent
Carbon monoxide 28 CO
31 0.01639 HCHO Formaldehyde 30 HCHO
93, 81, 107, 138 0.01512 Nopinoneb Carbon dioxide 44 CO2
70 0.01213 C4H5O Formic acid 46 HCOOH
45, 55, 56,60, 69, 99 0.01113 – Nopinone 138 C9H14O
47 0.01012 HCOOH 3-Hydroxy-nopinone 154 C9H14O2
97 0.009211 C5H4O2 cis-Pinic acid 186 C9H14O4
57, 71, 72, 82, 83, 96, 98 0.009013 –
59 0.009009 Acetone
61 0.008013 CH3COOH
characterize the ions as a primary product by showing a short
a
Obtained by fitting (1 exp ( a1t)) a2 to measured signals vs. revolution time or as a secondary product by showing a
time after injection of excess O3 (with O2) in He into an unstirred relatively long revolution time (Table 1).
bag reactor containing b-pinene. It was concluded from the differences in the fitted rise
b
Data from Fig. 1C.
coefficients in Table 1 that an ion with a unique rise coefficient
corresponds to a distinct neutral species in the atmospheric-
pressure reactor, moreover, the multiple ions with the same
products of the b-pinene–ozone reaction are now more easily rise coefficient in Fig. 1 also belong to a single species and
recognized (m/z 31, 45, 47, 57, 59, 69, 83, and 97). those fragments are produced by the low-pressure chemical
The logistic growth function (1 exp ( a1t))a2, where a1 is ionization of the single species inside the ion source of the
the rise coefficient adjustable in (s 1), and a2 (the upper limit of ITMS.
the growth), was used to fit the initial growth of product ions The ions in Fig. 1A with (m/z 57, 81, 93, 109, 135 and 137)
vs. time after injection of excess ozone into the unstirred reac- are associated with nopinone (scheme 1C) which is one of the
tor (Fig. 2). This experiment was carried out by injecting excess two primary products of the b-pinene–ozone reaction. Of the
ozone into the reactor containing b-pinene in order to monitor ions in Fig. 1D, the growth of m/z = 31, was satisfactorily fit-
the behavior of the unreacted b-pinene and the products of ted with the logistic growth function, with an evolution time of
b-pinene–ozone reaction (Fig. 1A). The measured growth coef- 21 ± 7 s, and is associated with protonated formaldehyde
ficients and evolution times of all ions (Fig. 1D) were used to which is the other primary product identified by Winterhalter
A new look at the b-pinene–ozone reaction usingthe atmospheric pressure reactor S669

Table 3 Yields of the primary products from the reaction of ozone with b-pinene.
Static bag reactor Winterhalter et al. (2000) Grosjean et al. (1993) Hakola et al. (1994)
Formaldehyde 0.58 ± 0.05 0.65 ± 0.04 0.42 –
Nopinone 0.19 ± 0.05 0.16 ± 0.04 0.22 0.23 ± 0.05

et al. (2000), Atkinson (1997), and Grosjean et al. (1996). This The ion at m/z = 70 showed a slow evolution time,
conclusion was supported by MS/MS data. diagnostic for a secondary product which may associate with
The growth of m/z = 47, is tentatively associated with a neutral parent ion or a short-lived neutral intermediate
HCOOH, which showed a slow evolution time, diagnostic of yielding m/z = 70 on protonation. The MS/MS spectrum of
a secondary product. The MS/MS experiments confirmed this m/z = 70 dominated with (m/z 70, 52 (base peak), and 30)
conclusion. Winterhalter et al. (2000) showed that HCOOH is which did not provide enough information to reach a definite
a secondary product resulting from the decomposition of the conclusion regarding the nature of this ion. The tentative
C9-Criegee intermediate and other radicals. The growth of chemical formula of m/z = 70 is C4H5O.
m/z = 59 is associated with presumably protonated acetone, The ion at m/z = 97 is associated with a secondary
which is probably present as an impurity in the reactor or as component with an induction delay of 42 ± 19 s; The
a product of the reaction of ozone with an olefinic impurity MS/MS spectrum of m/z = 97 dominated with (m/z 97, 79,
(Grosjean et al., 1993). The MS/MS experiments supported 69 (base peak), and 44), these data did not provide more infor-
this conclusion. mation to support whether the m/z = 97 is an unidentified new
The ion at m/z = 61 showed a slow evolution time, charac- secondary product or a contribution of short-lived intermedi-
teristic of a secondary product which is a contribution of a ate. It is believed that m/z = 97 may have a tentative chemical
protonation product of mass 60 such as CH3COOH. This con- formula of C5H4O2.
clusion was confirmed by the MS/MS data. Winterhalter et al.
(2000) who used ion chromatograms were able to identify 3.3. Calibrations and yields of the primary products
traces of CH3COOH as a result of the decomposition of unsta-
ble intermediates such as acetaldehyde monopeacetate. The calibrations of the primary products and their yields were
Fig. 1D showed the presence of overlapped multiple frag- performed without a scavenger in a similar procedure that was
ments which may result from at least two unidentified second- described in detail in previous work (Wedian and Atkinson
ary products. The mathematical fitting was used to correlate 2010). The yields of formaldehyde and nopinone were 0.58
the fragments in clusters, Table 1; the first cluster of ions at (±0.05), and 0.19 (±0.05), respectively. Table 3 shows a
m/z 45, 55, 56, 60, 69 and 99 showed a slow evolution time comparison of the yields of this work and other previous
of 55 ± 19 s which is a characteristic of a secondary product; b-pinene–ozone studies.
and another cluster of ions at m/z 57, 71, 72, 82, 83, 96 and 98 The limitations of this reactor in its present state of devel-
are fragments of the other secondary product of the opment lie mainly with the lack of accurate identification of
b-pinene–ozone reaction with an induction delay of the parent ion mass of fragments; Moreover, the system does
48 ± 14 s. Based on Winterhalter et al. (2000) work, these not exclude any possible contribution of thermalized neutral
fragments are associated with two detected secondary prod- fragments. Those limitations could be overcome when a stan-
ucts: 3-hydroxy-nopinone and the cis-pinic acid. The MS/MS dard mass spectra library of the ion trap mass spectrometer is
experiments for selected m/z from each cluster showed severe available for most detected compounds.
fragmentation and produced no unique ion fragment in the
daughter mass spectra making it difficult to correlate them
to either possible compound. Up to the time of doing these 4. Conclusion
experiments, standard mass spectra of pure 3-hydroxy-nopi-
none and the cis-pinic acid were not available in our research The b-pinene–ozone reaction was studied without scavenger in
group which limited the ability to get a positive correlation of a atmospheric bag reactor coupled into ITMS. The detected
ions to their neutral compound. ions of the products showed different rise coefficients in the
The main products of the b-pinene–ozone reaction static reactor. The observed differences in rise coefficients were
identified by Winterhalter et al. (2000) are HCHO, nopinone, used to distinguish the ions and then associated with their
3-hydroxy-nopinone, CO2, CO, HCOOH, the secondary possible parent ions. In this system, at least two secondary
ozonide of b-pinene, and cis-pinic acid. There were two product fragments were failed to be correlated to their possible
primary products of the b-pinene–ozone reaction identified parent ions. The reactor showed a potential advantage in the
by Winterhalter et al. (2000), Atkinson (1997), and Grosjean gas phase studies by the detection of short-lived intermediates
et al. (1996): formaldehyde and nopinone. Table 2 lists all at m/z = 70 and 97.
primary and secondary products of the b-pinene–ozone
reaction and their molar masses identified in the literature.
Acknowledgments
3.2. New short-lived intermediates
The author would like to acknowledge the financial support of
This reactor was able to detect new ions with unique rise the Tafila Technical University (Grand number 176/2008), and
coefficients which could have possibly resulted from contribu- the author would like to thank the Portland State University’s
tions of thermalized neutral intermediates: atmospheric research group for their inspiration.
S670 F. Alwedian

References Hakola, H., Arey, J., Aschmann, S.M., Atkinson, R., 1994. J. Atmos
Chem. 18, 75. http://dx.doi.org/10.1007/BF00694375.
Atkinson, R., 1997. J. Phys. Chem. Ref. Data 26, 215–290. http:// Hatakeyama, S., Izumi, K., Fukuyama, T., Akimoto, H., 1989. J.
dx.doi.org/10.1063/1.556012. Geophys. Res. 94, 13013–13024. http://dx.doi.org/10.1029/
Atkinson, R., Arey, J., 1998. Acc. Chem. Res. 31, 574. http:// JD094iD10p13013.
dx.doi.org/10.1021/ar970143z. Hatakeyama, S.I.K., Fukuyama, T., Akimoto, H., Washida, N., 1991. J.
Atkinson, R., Aschmann, S.M., 1993. J. Atmos. Chem. 16, 337–348. Geophys. Res. 96, 947–958. http://dx.doi.org/10.1029/90JD02341.
http://dx.doi.org/10.1007/BF01032629. Jay, K., Stieglitz, L., 1987. Air Pollut. Ecosyst., 542–547.
CalogIrou, A., Jensen, N.R., Nielsen, C.J., Kotzias, D., Hjorth, J.l., Librando, V., Tringali, G., 2005. J. Environ. Manage. 75, 275–282.
1999. Environ. Sci.Tech. 33, 453–460. http://dx.doi.org/10.1021/ http://dx.doi.org/10.1016/j.jenvman.2005.01.001.
es980530j. Seinfeld, J.H., Pandis, S.N., 1998. Atmospheric Chemistry and Physics:
Grosjean, D., Williams II, E.L., Grosjean, E., Andino, J.M., Seinfeld, From Air Pollution to Climate Change, 2nd ed. Wiley,
J.H., 1993. Envirn. Sci. Technol. 27, 2754–2758. http://dx.doi.org/ New York.
10.1021/es00049a014. Tuazon, E.C., Aschmann, S.M., Arey, J., Atkinson, R., 1998. Environ.
Grosjean, E., Andrade, J.B., Grosjean, D., 1996. Environ. Sci. Sci. Technol. 32, 2106–2112. http://dx.doi.org/10.1021/es980153a.
Technol. 30, 975–983. http://dx.doi.org/10.1021/es950442o. Venkatachari, P., Hopke, P.K., 2008. J. Envorn. Monitor. 10, 966–
Guenther, A., Nicholas, H.C., Erickson, D., Fall, R., Geron, C., 974. http://dx.doi.org/10.1039/b804357d.
Graedel, T., Harley, P., Klinger, L., Lerdau, M., Mckay, W.A., Wedian, F., Atkinson, D.B., 2010. Canad. J. Chem. 88, 1017–1025.
Pierce, T., Scholes, B., Steinbrecher, R., Tallamraju, R., Taylor, T., http://dx.doi.org/10.1139/V10-119.
Zimmerman, P., 1995. J. Geophys. Res. 100, 8873–8892. http:// Winterhalter, R., Neeb, P., Grossmann, D., Kolloff, A., Horie, O.,
dx.doi.org/10.1029/94JD02950. Moortgat, G., 2000. J. Atmos. Chem. 35, 165–197. http://
Hasson, A.S., Orzechowska, G., Paulson, S.E., 2001. J. Geophys. Res. dx.doi.org/10.1023/A:1006257800929.
106, 3431–3442. http://dx.doi.org/10.1029/2001JD000597.

You might also like