• Nebyly nalezeny žádné výsledky

Ideas and plans for further work

In document Text práce (2.331Mb) (Stránka 80-124)

4. Conclusions

4.2. Ideas and plans for further work

The experimental work on this thesis was finished in the first quarter of year 2006.

Since then the author continues to work in the same laboratory carrying out research related to SIFT-MS but not any longer as a part of the topic of the doctoral thesis. The legacy SIFT instrument has worked hard well over 25 years and is now being dismantled. In September 2006 a new SIFT-MS instrument was commissioned in our laboratory, this is a new generation of instrumentation commercially manufactured (Profile 3, Instrument Science Limited, Crewe, England) as a transportable unit weighing 120 kg with external dimensions of 80 x 60 x 60 cm based on a flow tube 5 cm long. We have started several research projects using this new instrument including research of processes within the microwave plasma ion source and improvement of detection limits for quantitative real time breath analysis. From January 2007 we are starting new projects concerning the research of ion chemistry of Titan atmosphere and a project in collaboration with an explosives manufacturer Synthesia for detection of explosives and products of their combustion. We have initiated pilot projects in breath analysis of healthy volunteers, that already resulted in two papers already published, see the author’s C.V., and we are staring collaboration with microbiology and experimental medicine institutes. Other than this we intend to follow up the research of diffusion processes in the smaller flow tube.

81

References

[1] Yarris L. Clusters: a new state of matter, Lawrence Berkeley National Laboratory Currents Weekly Newsletter, 1991.

Available on-line: http://www.lbl.gov/LBL-Science-Articles/Archive/clusters.html [2] Columbia Encyclopedia, 6th ed, Columbia University Press, 2005.

Available on-line: http://www.encyclopedia.com/html/c1/complexi.asp

[3] Santos L.S., Catharino R., Eberlin M.N. The proton-bound dimer of acetone, J. Mass Spectrom.

40 (1) (2005) 127 – 128.

[4] Engel A. Ionised gases, Phys.-Math. Literature Press, Moscow, 1959.

[5] Bowers M.T. Gas Phase Ion Chemistry, Vol. 1, Academic Press, New York, 1979.

[6] Smith D., Španěl P. Ions in the terrestrial atmosphere and in interstellar clouds, Mass Spectrom.

Rew. 14 (1995) 255-278.

[7] Amann A., Smith D. Breath analysis for clinical diagnosis and therapeutic monitoring, World Scientific, Singapore, 2005, pp. 3-35.

[8] Smith D., Španěl P. Selected ion flow tube mass spectrometry (SIFT-MS) for on-line trace gas analysis, Mass Spectrom. Rev. 24 (2005) 661–700.

[9] Smith D., Adams N.G. Elementary interactions between charged and neutral species in plasmas, Pure Appl. Chem. 56 (2) (1984) 175–188.

[10] McDaniel E.W. Collision phenomena in ionised gases, John Wiley & Sons, Inc, New York, 1964.

[11] Knewstubb P.F. Mass spectrometry and ion-molecule reactions. University Press, Cambridge, 1969.

[12] Gioumousis G., Stevenson D.P. Reactions of gaseous molecule ions with gaseous molecules. V.

Theory, J. Chem. Phys. 29 (2) (1958) 294-299.

[13] Su T., Chesnavich W.J. Parametrization of the ion-polar molecule collision rate constant by trajectory calculations, J. Chem. Phys. 76 (10) (1982) 5183-5185.

[14] Atkins P.W. Physical Chemistry. 4th ed., Oxford University Press, Oxford, 1992.

[15] Elrod M.J. A comprehensive computational investigation of the enthalpies of formation and proton affinities of C4H7N and C3H3ON compounds, Int. J. Mass Spectrom. 228 (2003) 91–105.

[16] Jost C., Sprung D., Kenntner T., Reiner T. Atmospheric pressure chemical ionisation mass spectrometry for the detection of tropospheric trace gases: the influence of clustering on sensitivity and precision, Int. J. Mass Spectrom. 223–224 (2003) 771–782.

[17] Biondi M.A., Brown S.C. Measurement of Electron-Ion Recombination, Phys. Rev. 76 (1949) 1697-1700.

[18] Plasil R., Glosik J., Poterya V., Kudrna P., Rusz J., Tichy M., Pysanenko A. Advanced integrated stationary afterglow method for experimental study of recombination of processes of H3

+ and D3 +

ions with electrons, Int. J. Mass Spectrom. 218 (2) (2002) 105–130.

[19] Rebrion-Rowe C., Le Garrec J.L., Hassouna M., Travers D., Rowe B.R. Experimental evaluation of the recombination rate of cations formed from fluoranthene, Int. J. Mass Spectrom. 223 (2003) 237–251.

82

[20] Meek J.M., Craggs J.D. Electrical breakdown of gases, Clarendon Press, Oxford, 1953.

[21] Bardsley J.N., Biondi M.A. Dissociative recombination, Adv. At. Mol. Phys. 6 (1970) 1–57.

[22] Smith D., Goodall C.V., Copsey M.J. Investigation of the helium afterglow II. Langmuir probe observations, J. Phys. B, 1 (1968) 660-668.

[23] Ferguson E.E., Fehsenfeld F.C., Schmeltekopf A.L. Ion-molecule reaction rates measured in a discharge afterglow. Advances in Chemistry Series 80 (1969) 83-91.

[24] Rowe B.R., Canosa A., Lepage V. FALP and CRESU studies of ionic reactions, Int. J. Mass Spectrom. Ion Proces., 149 (1995) 573–596.

[25] Smith D., Španěl P. Studies of electron attachment at thermal energies using the FALP technique, Adv. At. Mol. Phys. 32 (1994) 307–343.

[26] Španěl P., Smith D. Radiation from the Reactions of NO+ with Cl- and I-, Chem. Phys. Let. 258 (1996) 477–484.

[27] Pysanenko A., Novotny O., Zakouril P., Plasil R., Poterya V., Glosik J. Recombination of H3 + and H5

+ ions with electrons in He-Ar-H2 flowing afterglow plasma. Czech. J. Phys. 52 (2002) D681–

694.

[28] Miller T.M., Van Doren J.M., Viggiano A.A. Electron attachment and detachment: C6F6, Int. J.

Mass Spectrom. 233 (2004) 67–73.

[29] Adams N.G., Smith D. The selected ion flow tube (SIFT): A technique for studying ion-neutral reactions, Int. J. Mass Spectrom. Ion Phys. 21 (1976) 349–359.

[30] Mackay G.I., Vlachos G.D., Bohme D.K., Schiff H.I. Studies of reactions involving C2Hx + ions with HCN using a modified selected ion flow tube, Int. J. Mass Spectrom. Ion Phys. 36 (1980) 259–270.

[31] Van Doren J.M., Barlow S.E., DePuy C.H., Bierbaum V.M. The tandem flowing afterglow-SIFT-Drift. Int. J. Mass Spectrom. Ion Processes 81 (1987) 85-100.

[32] Glosik J., Twiddy N.D., Javahery G., Ferguson E.E. Measurement of the equilibrium constant of the reaction HeH+ + Ne NeH+ + He in a selected ion flow tube, Int. J. Mass Spectrom. Ion Proces. 109 (1991) 75–81.

[33] Iraqi M., Petrank A., Peres M., Lifshitz C. Proton transfer reactions of C2H2

+: the bond energy D0

(C2H-H), Int. J. Mass Spectrom. Ion Proces. 100 (1990) 679-691.

[34] Knight J.S., Freeman C.G., McEwan M.J. Selected-ion flow tube studies of HC3N, Int. J. Mass Spectrom. Ion Proces. 67 (1985) 317–330.

[35] Viggiano A.A., Miller T.M., Stevens Miller A.E., Morris R.A., Van Doren J.M., Paulson J.F. SF4: electron affinity determination by charge-transfer reactions, Int. J. Mass Spectrom. Ion Proces.

109 (1991) 327–338.

[36] Jackson D.M., Stibrich N.J., Adams N.G., Babcock L.M., A selected ion flow tube study of the reactions of a sequence of ions with amines, Int. J. Mass Spectrom. 243 (2005) 115–120.

[37] Koyanagi G.K., Zhao X., Blagojevic V., Jarvis M.J.Y., Bohme D.K. Gas-phase reactions of atomic lanthanide cations with methyl fluoride: periodicities reactivity, Int. J. Mass Spectrom.

241 (2005) 189–196.

83

[38] Schoon N., Amelynck C., Vereecken L., Arijs E. A selected ion flow tube study of the reactions of H3O+, NO+ and O2

+ with a series of monoterpenes, Int. J. Mass Spectrom. 229 (2003) 231–240.

[39] Španěl P., Smith D. SIFT studies of the reactions of H3O+, NO+ and O2

+ with a series of alcohols.

Int. J. Mass Spectrom. Ion Proc. 167/168 (1997) 375-388.

[40] Roithova J., Herman Z., Schreder D., Schwarz H. Competition of proton and electron transfers in gas-phase reactions of hydrogen-containing dications CHX2+ (X=F, Cl, Br, I) with atoms, nonpolar and polar molecules, Chem. Eur. J. 12 (2006) 2465–2471.

[41] Španěl P., Smith D. Selected ion flow tube - mass spectrometry: detection and real-time

monitoring of flavours released by food products Rapid Commun. Mass Spectrom. 13 (1999) 585-597.

[42] Španěl P., Van Doren J.M., Smith D. A selected ion flow tube study of the reactions of H3O+, NO+ and O2

+ with saturated and unsaturated aldehydes and subsequent hydration of the product ions. Int. J. Mass Spectrom 213 (2002) 163–176.

[43] Španěl P, Smith D. Selected ion flow tube: a technique for quantitative trace gas analysis of air and breath. Med. Biol. Eng. Comput. 34 (1996) 409–419.

[44] Manolis A. The diagnostic potential of breath analysis. Clin. Chem. 29 (1983) 5–15.

[45] Turner C., Španěl P., Smith D. A longitudinal study methanol in the exhaled breath of 30 subjects using selected ion flow tube mass spectrometry, SIFT-MS. Physiol. Meas. 27 (2006) 637–648.

[46] Jones J.D.C., Lister D.G., Twiddy N.D. Equilibrium-constant for the reaction Xe+ +

2Ar-reversible-XeAr+ +Ar in the temperature-range 150K-300K and the dissociation-energy of XeAr+, Chem. Phys. Lett. 70(3) (1980) 575–578.

[47] Španěl P., Smith D., Henchman M. The reactions of some interstellar ions with benzene, cyclopropane and cyclohexane, J. Chem. Phys. 141 (1995) 117-126.

[48] Španěl P., Smith D. Application of ion chemistry and the SIFT technique to the quantitative analysis of trace gases in air and on breath, Int. Rev. Phys. Chem. 15 (1996) 231–271.

[49] Miller P.E., Denton M.B. The quadrupole mass filter: Basic operating concepts. J. Chem Ed. 63 (7) (1986) 617–622.

Douglas D.J., Konenkov N.V. Influence of the 6th and 10th spatial harmonics on the peak shape of a quadrupole mass filter with round rods. Rapid commun. mass spectrom 16 (2002) 1425-1431.

[50] Mathieu E. Memoire sur le mouvement vibratoire dune membrane de forme elliptique. J. Math.

Pure Appl. 13 (1868) 137-203.

[51] March R.E. An Introduction to Quadrupole Ion Trap Mass Spectrometry. J. Mass Spectrom. 32 (1997) 351–369.

[52] Dawson P.H., Douglas D.J. Quadrupoles, Used of in Mass Spectrometry, in Encyclopedia of Spectroscopy and Spectrometry, Ed. by J. Lindon, G. Trantner and J. Holmes, Academic Press, London, 1999.

[53] Pedder R.E. Practical Quadrupole Theory: Graphical Theory. Extrel Application Note, Pennsylvania, 2001.

Available on-line: http://www.ardaratech.com/app_notes.html.

84

[54] Španěl P., Smith D. On-line measurement of the absolute humidity of air, breath and liquid headspace samples by selected ion flow tube mass spectrometry, Rapid Comm. Mass Spectrom.

15 (2001) 563-569.

[55] Španěl P., Smith D. Influence of water vapour on SIFT-MS analyses of trace gases in humid air and breath. Rapid Comm. Mass Spectrom. 14 (2000) 1898-1906.

[56] Lide D.R. (ed.), CRC Handbook of Chemistry and Physics, CRC: Boca Raton, 1991.

[57] KDB - Hydrocarbons Properties. POWERED BY Thermodynamics and Properties LAB. Dept. of Chemical Engineering, Korea Univerity.

[58] Dewar M.J.S., Shanshal M., Worley S.D. Calculated and observed ionization potentials of nitroalkanes and of nitrous and nitric acids and esters. Extension of Mindo method to nitrogen-oxygen compounds. J. Am. Chem. Soc. 91 (13) (1969) 3590.

[59] Smith D., Wang T.S., Španěl P. Analysis of ketones by selected ion flow tube mass spectrometry Rapid Comm. Mass Spectrom. 17 (2003) 2655-2660.

[60] Španěl P., Yufeng J., Smith D. SIFT studies of the reactions of H3O+, NO+ and O2

+ with a series of aldehydes and ketones. Int. J. Mass Spectrom. Ion Proc., 165/166 (1997) 25-37.

[61] Smith D., Španěl P., Jones J.B. Analysis of volatile emissions from porcine faeces and urine using selected ion flow tube mass spectrometry. Bioresource Technology 75 (2000) 27-33.

[62] Deakyne C.A., Meot-Ner (Mautner) M., Campbell C.L., Hughes M.G., Murphy S.P.

Multicomponent Cluster Ions. 1. The Acetonitrile - Water System. J. Chem. Phys. 84 (1986) 4958-4969.

[63] Abbott S.M., Elder J.B., Španěl P., Smith D. Quantification of acetonitrile in exhaled breath and urinary headspace using selected ion flow tube mass spectrometry. Int. J Mass Spectrom. 228 (2003) 655-665.

[64] Committee on Confronting Terrorism in Russia, Office for Central Europe and Eurasia Development, Security, and Cooperation, National Research Council in Cooperation with the Russian Academy of Sciences. High-Impact Terrorism: Proceedings of a Russian-American Workshop (2002), The National Academies Press,

Available on-line: http://www.nap.edu/books/0309082706/html

[65] United Nations Office on Drugs and Crimes (UNODC), Convention on the Marking of Plastic Explosives for the Purpose of Detection. Montreal, March 1, 1991. (date of entry into force, June 21, 1998).

[66] Schulte-Ladbeck R., in: Proceedings of the Pittsburgh Conference on Novel Analytical Methods for the Determination of Peroxide-Based Explosives (lecture 1336), Pittsburgh, 2002.

[67] Chladek J. The identification of organic peroxides, Adv. Anal. Detect. Explos., Proc. Int. Symp. 4 (1992) 73–76.

[68] Buttigieg G.A., Knight A.K., Denson S., Pommier C., Denton M.B. Characterization of the explosive triacetone triperoxide and detection by ion mobility spectrometry. Forensic Sci. Int. 135 (2003) 53–59

[69] Wilson P.F., Prince B.J., McEwan M.J. Application of selected-ion flow tube mass spectrometry to the real-time detection of triacetone triperoxide. Analytical Chemistry 78 (2006) 575-579.

85

[70] Currie L.A. Nomenclature in evaluation of analytical methods including detection and quantification capabilities. Anal. Chim. Acta 391 (1999) 105–126.

[71] Španěl P, Smith D. The Selected Ion Flow Tube a Novel Technique for the Quantitative Trace Gases Analysis of Air and Human Breath. Med. Biol. eng. Comput. 34 (1996) 409-419.

[72] Španěl P., Smith D. Quantitative selected ion flow tube mass spectrometry: the influence of ionic diffusion and mass discrimination, J. Am. Soc. Mass Spectrom., 12 (2001) 863-872.

[73] Miller T.M., Viggiano A.A. Electron attachment and detachment: C6F5Cl, C6F5Br, and C6F5I and the electron affinity of C6F5Cl, Phys. Rev. A 71 (1) (2004) 1-11.

[74] de Gouw J.A., Krishnamurthy M., Bierbaum V.M., Leone S.R. Measured and calculated mobilities of cluster ions drifting in helium and in nitrogen, Int. J. Mass Spectrom. Ion Proc.

167/168 (1997) 281-289.

[75] A. Blanc, J. Phys., 7 (1908) 825.

[76] M.P. Langevin, Ann. Chim. Phys., 5 (1905) 245.

[77] Mason E.A.,. McDaniel E.W Transport properties of ions in gases, Wiley, New York, 1988.

[78] Takebe M., Satoh Y., Iinuma K., Seto K. Mobilities and longitudinal diffusion coefficient for K+ ions in nitrogen and argon, J. Chem. Phys., 73 (8) (1980) 4071-4076

[79] Hegerberg R., Elford M.T., Skullerud H.R. The cross section for symmetric charge exchange of Ne+ in Ne and Ar+ in Ar at low energies, J. Phys. B: At. Mol. Phys. 15 (1982) 797-811.

[80] Mechlinska-Drewkoa J., Wroblewski T., Lj Z., Novakovic V., Karwasz G.P. Electron scattering on N2O-from cross sections to diffusion coefficients, Radiat. Phys. Chem. 68 (2003) 205–209.

[81] Bastian M.J., Lauenstein C.P., Bierbaum V.M., Leone S.R. Single-frequency laser probing of velocity component correlations and transport-properties of Ba+ drifting in Ar, J. Chem. Phys. 98 (12) (1993) 9496-9512.

[82] Smith D., Dean A.G., Adams N.G. Space-charge fields in afterglow plasmas. J. Phys D 7 (14) (1974) 1944-1962.

[83] Viehland L.A., Mason E.A. Transport properties of Gaseous ions over a wide energy range. Part IV, ADND Tables, 60 (1995) 37-95.

[84] de Gouw J.A., Ding L.N., Krishnamurthy M., Lee H.S., Anthony E.B., Bierbaum V.M., Leone S.R. The mobilities of NO+(CH3CN)n cluster ions (n = 0–3) drifting in helium and in helium-acetonitrile mixtures. J .Chem. Phys. 105 (23) (1996) 10398-10409.

[85] Viehland L.A., Mason E.A., Morrison W.F., Flannery M.R. Tables of transport collision integrals for (n, 6, 4) ion-neutral potentials, ADND Tables, 16 (1975) 495-514.

[86] Dryahina K., Španěl P. Theoretical Study of Diffusion of Ions Important for Breath Analysis by the SIFT-MS Method. in: J. Safrankova (Ed.), WDS'04 Proceedings of Contributed Papers, Part II, MATFYZPRESS, Prague, 2004, pp. 407-412.

[87] Spartan programm, Wavefunction, Inc, Irvine, CA

[88] Frisch M.J., G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman,

V.G. Zakrzewski, J.A. Montgomery Jr., R.E. Stratmann, J.C. Burant, S. Dapprich, J.M. Millam, A.D. Daniels, K.N. Kudin, M.C. Strain, O. Farkas, J. Tomasi, V. Barone, M. Cossi, R. Cammi, B. Mennucci, C. Pomelli, C. Adamo, S. Clifford, J. Ochterski, G.A. Petersson, P.Y. Ayala,

86

Q. Cui, K. Morokuma, D.K. Malick, A.D. Rabuck, K. Raghavachari, J.B. Foresman, J. Cioslowski, J.V. Ortiz, B.B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. Gomperts, R.L. Martin, D.J. Fox, T. Keith, M A. Al-Laham, C.Y. Peng, A. Nanayakkara, C. Gonzalez, M. Challacombe, P.M.W. Gill, B. Johnson, W. Chen, M.W. Wong, J.L. Andres, C. Gonzalez, M. Head-Gordon, E.S. Replogle, J.A. Pople, Gaussian 98, Revision A.6, Gaussian Inc., Pittsburgh PA, 1998.

[89] Bondi A. van der Waals volumes and radii, J. Phys. Chem., 68 (1964) 441-451.

[90] von Helden G., Hsu M.T., Gotts N., Bowers M.T. Carbon cluster cations with up to 84 atoms:

structures, formation mechanism, and reactivity, J. Phys. Chem. 97 (1993) 8182-8192.

[91] James A.M., Lord M.P. Macmillan’s Chemical and Physical Data, Macmillan, London, 1992.

[92] Dryahina K. DiffusionCalc software, 2005.

Available on-line at http://www.jh-inst.cas.cz/~dryahina/.

[93] Koutselos A.D., Mason E.A. Correlation and prediction of dispersion coefficients of isoelectronik systems. J. Chem. Phys. 85 (1986) 2154-2160.

[94] Selle S., RiedelU. Transport coefficients of reacting air at high temperatures, AIAA Paper 2000-0211, Reston, 2000.

[95] Miller T.M., Bederson B., Atomic and molecular polarizabilities – a review of recent advances Adv. At. Mol. Phys. 13 (1977) 1-55.

[96] Ellis H.W., McDaniel E.W., Albritton D.L., Viehland L.A., Lin S.L., Mason E.A. Transport properties of Gaseous ions over a wide energy range. Part II. At. Data Nucl. Data Tables 22 (1978) 179-217.

[97] Ellis H.W., Pai R.Y., McDaniel E.W., Mason E.A., Viehland L.A. Transport properties of Gaseous ions over a wide energy range. At. Data Nucl. Data Tables 17 (1976) 177-210.

A1. Dryahina K., Polášek M., Španěl P. A selected ion flow tube, SIFT, study of the ion chemistry of H3O+, NO+ and O2+• ions with several nitroalkanes in the presence of water vapour.

International Journal of Mass Spectrometry 239 (2004) 57-65.

International Journal of Mass Spectrometry 239 (2004) 57–65

A selected ion flow tube, SIFT, study of the ion chemistry of H

3

O

+

, NO

+

and O

2+

ions with several nitroalkanes in the presence of water vapour

Kseniya Dryahina, Miroslav Pol´aˇsek, Patrik ˇSpanˇel

J. Heyrovsk´y Institute of Physical Chemistry, Academy of Sciences of the Czech Republic, Dolejˇskova 3, 18223 Prague 8, Czech Republic

Received 17 August 2004; accepted 22 September 2004 Available online 28 October 2004

Abstract

We have carried out a selected ion flow tube (SIFT) study of the reactions of H3O+, NO+and O2+with the following six nitroalkanes:

nitromethane, nitroethane, 1-nitropropane, 2-nitropropane, 1-nitrobutane and 2-methyl-2-nitropropane. The primary purpose of this work was to extend the kinetics database to allow these compounds, M, to be analysed in air by selected ion flow tube mass spectrometry, SIFT–MS.

Some nitroalkanes are used as industrial solvents and some are component of agricultural agents that are known health hazards. The initial step in all the H3O+reactions is exothermic proton transfer to produce MH+ions, which are seen to be the only products for the two smallest nitroalkanes, but for the isomers of nitropropane and nitrobutane, fragmentation of the nascent MH+ions occurs. NO+reacts with the four smallest compounds via association resulting in NO+M product ions, whilst for the isomers of nitrobutane the C4H9+hydrocarbon ion is produced. The reaction of O2+with nitromethane proceeds via charge transfer giving M+as the major product, whilst the O2+ reactions with all the remaining nitroalkanes in this study lead to a single hydrocarbon ion product CnH2n+1+. The secondary chemistry of the ion products with H2O and with M, which is relevant to SIFT–MS applications, is fully described, with the interesting finding that water cluster ions of the kind MH+(H2O)3containing three water molecules are formed at 300 K. The mechanisms of the reactions are described with the aid of ab initio calculations of the ion energetics that were not previously available for some of the ions involved in the chemistry.

© 2004 Elsevier B.V. All rights reserved.

Keywords: SIFT; Nitropropane; Nitromethane; SIFT–MS

1. Introduction

Nitroalkanes are commonly used as solvents in indus-trial construction and maintenance, printing, highway main-tenance and food packaging. They have been widely used as chemical intermediates, industrial solvents, and as compo-nents of inks, paints, varnishes and other coatings[1]. Poten-tial occupational exposure to nitro compounds occurs during their manufacture and use. Nitromethane, along with other nitroalkanes, has been detected in cigarette smoke[2]. Ni-tromethane and 2-nitropropane are considered to be carcino-genic to humans and animals[3,4]. High concentrations of nitroalkanes (above 600 ppm) are acutely toxic and have

pro-∗Corresponding author. Tel.: +420 2 6605 2112; fax: +420 2 858 2307.

E-mail address: spanel@seznam.cz (P. ˇSpanˇel).

duced industrial fatalities[5]. Occupational exposure limits set by the US Department of Labour are 100 ppm for ni-tromethane and nitroethane and 25 ppm for both isomers of nitropropane[6]. Selected ion flow tube (SIFT)–MS[7]has the potential for accurate on-line real time measurement and monitoring of concentrations of these nitroalkanes in air.

SIFT–MS may also be feasible analytical method for bi-ological monitoring via breath analysis of the exposure of workers to the most widely used nitroalkane, 2-nitropropane.

Human uptake of nitroalkanes occurs mainly through the lungs. In experimental studies on animals (rats)[5]it has been shown that 2-nitropropane is not only rapidly absorbed via the lungs but also via the peritoneal cavity and the gastrointestinal tract; however, there is no satisfactory information on absorp-tion via the skin. 2-Nitropropane is not retained in the body for more than a few hours, since it is rapidly metabolised,

1387-3806/$ – see front matter © 2004 Elsevier B.V. All rights reserved.

doi:10.1016/j.ijms.2004.09.007

58 K. Dryahina et al. / International Journal of Mass Spectrometry 239 (2004) 57–65

mainly to acetone, nitrite salts and possibly some isopropyl alcohol[8]. These metabolites and 2-nitropropane are then rapidly lost from the body by further metabolic transforma-tion, exhalation and excretion[5].

The data also may be of some relevance to atmospheric ion chemistry. In situ measurements of the ion composition of the atmosphere indicate that CH3NO2is also present as a component of complex hydrated H3O+based clusters[9].

In order to establish the library of kinetic data and to un-derstand the trends of reactivity of the common SIFT–MS precursor ions H3O+, NO+, O2+, a series of studies of the reaction of these ions with different classes of molecules has been carried out; see the recent studies of the reactions of diols and the references there in[10]. The objective of the present study is to extend this library of data by providing information on the trends of reactivity of several nitroalka-nes with the SIFT–MS precursor ions. As the information on proton affinities (PAs) is not available for all the nitroalka-nes included, the study was complemented by their ab initio calculations. Details of the reaction mechanisms involving fragmentation were elucidated by information from parallel study using collisionally activated dissociation (CAD) in a sector mass spectrometer.

Protonated parent molecules that are produced in the pri-mary reactions of H3O+with small molecules can bind one or more water molecules under the typical conditions of the SIFT experiments, viz. helium carrier gas, temperature 300 K, pressure 100 Pa. Notable exception are protonated hy-drocarbons that do not associate efficiently with H2O under these SIFT conditions[11]. The maximum number of bound water molecules is characteristic of the different classes of compounds. Thus ketones, ethers and esters typically bind only a single water molecule [12], but aldehydes [13], al-cohols [14] and carboxylic acids [15] can bind two H2O molecules. The biggest number of added water molecules observed to date in SIFT experiments at 300 K is three for H3O+ [16] and CH3CNH+ [17]. As it will be shown later, the protonated nitroalkanes also are seen to add three water molecules.

2. Methods

2.1. Selected ion flow tube (SIFT)

The well-known SIFT technique for the determination of the rate coefficients and ion product distributions for the re-actions of H3O+, NO+ and O2+ ions with organic com-pounds has been described in numerous previous publications [18,19], so it is sufficient here to summarise it as follows.

Precursor ions (H3O+, NO+and O2+•) are generated in a mi-crowave discharge ion source, mass selected by a quadrupole mass filter and then injected as into helium carrier gas, flow-ing at a velocity of 170 m/s in the Prague SIFT. The reactant vapours of interest (in these experiments, nitromethane, ni-troethane, 1-nitropropane, 2-nitropropane, 1-nitrobutane and

2-methyl-2-nitropropane) are then introduced into the ion swarm/carrier gas where they react with the chosen precur-sor ion species. Measurement of the absolute flow rates of their neat vapours into the carrier gas of the SIFT instrument, required for determination of absolute rate coefficients[18], cannot be achieved to an accuracy better than 20% due to the surface activity of these compounds. In these circumstances, the established procedure is to measure relative rate coeffi-cients of the H3O+, NO+and O2+ions with the compound.

Thus, the reactant nitroalkane is introduced as its saturated vapour in helium by introducing a small droplet of liquid ni-troalkane in a sealed plastic bag, which is then inflated with helium. This helium/nitro compound vapour mixture is then introduced into the helium carrier gas/precursor ion swarm of the SIFT via a flow meter in the usual way. The H3O+, NO+ and O2+ions are injected simultaneously into the carrier gas by switching the injection mass filter to the total ion mode [20].

The relative concentrations of the primary and product ions of the reactions are determined from the count rates recorded by a downstream quadrupole mass spectrometer with a pulse counting single channel electron multiplier de-tector. This can be operated in the full scan mode (FSM) over a predetermined m/z range to obtain a spectrum of the reactant and product ions or in the multi-ion mode (MIM) in which the spectrometer is switched and dwells on selected reactant/product ions as their count rates are determined[20].

The FSM is primarily used to identify the product ions and the MIM is used to accurately determine the rate coefficients and the product ion distributions[20,21]. Primary product branching ratios are determined by extrapolation of the mea-sured ion count rate ratios towards zero reactant flow rate [18].

The relative rate coefficients for the reactions of the three precursor ions with the given molecule are then determined from the slopes of the plots of logarithm of ion count rates against the vapour mixture flow rate. It can be justifiably assumed that the proton transfer reactions of H3O+ with the nitroalkane compounds are exothermic and proceed at their collisional rates, because the proton affinities of the ni-troalkanes exceed the proton affinity of water molecules[22]

by >20 kJ/mol[23]. Collisional rate coefficients, kc, for the H3O+reactions have been calculated from the polarizabilities and the dipole moments of the reactant nitroalkanes using the parameterised trajectory formulation of Su and Chesnavich [24]. Values of the dipole moments and of the polorizabilities are from[25]with the single exception of 2-nitropropane, the polarizability of which is unknown, but we estimate it to be the same as the polarizability of its isomer 1-nitropropane.

The rate coefficients, k, for the NO+and O2+reactions are then experimentally derived from the relative decay rates of the three ion species (see the previous successful applications of this procedure, for example, in[12,13]).

In SIFT–MS analyses of moist air samples, it is impor-tant to know if any of the product ions, R+, of the analytical reactions undergo association with water molecules[16]. To

K. Dryahina et al. / International Journal of Mass Spectrometry 239 (2004) 57–65 59

Fig. 1. Experimental SIFT data tracing the ion chemistry that occurs when H3O+ions are injected into the helium carrier gas into which a small steady flow of nitromethane and variable flows of water vapour are introduced. The count rates, c/s, for the precursor and product ions indicated are plotted on a semilogarithmic scale as a function of the number density of water molecules [H2O] in the carrier gas.

investigate this, controlled amounts of an air/water vapour mixture were introduced into the carrier gas whilst monitor-ing the R+ ions using the MIM mode (seeFig. 1). The rate coefficients for association reactions that occur can be esti-mated from the dependencies of the [R+·H2O]/[R+] count rate ratio on the H2O molecule number density in the carrier gas, which is calculated from the distribution of the H3O+ ions and its hydrates H3O+·(H2O)1,2,3. This procedure has been discussed previously in a paper on the influence of humidity on SIFT–MS analyses[16].

2.2. Collisionally activated dissociation (CAD), mass spectrometry

In order to elucidate mechanisms of the fragmentation observed in the O2+reactions, additional information was used from collisionally activated dissociation mass spectra recorded on ZAB2-SEQ hybrid tandem mass spectrome-ter equipped with chemical ionisation (CI) ion source. The ions of interest were generated by self-chemical ionisation (self-CI) using the standard ion source conditions (0.5 mA, 100 eV, 150C) at the pressure of 6 × 103Pa measured at the ion source diffusion pump intake. The ions were ex-tracted from the ionisation chamber, accelerated to 8 keV, and products of their collisional activation in the first and second field-free regions were analysed using the B/E = const. linked scan and mass-analysed ion kinetic energy (MIKE) scan, respectively. Helium was used as a collision gas.

2.3. Quantum chemical calculations

Standard ab initio calculations of proton affinities of nitroalkanes were preformed using the Gaussian 98 suite of programs [26]. Geometries were optimised with den-sity functional calculations using Becke’s hybrid functional

(B3LYP) [27] and the 6-31+G(d,p) basis set. Stationary points were characterized by harmonic frequency lations as local minima (all frequencies real). The calcu-lated frequencies were scaled by the factor of 0.963 [28]

and used to obtain zero-point energy (ZPVE) corrections and enthalpies. For selected species, the geometries were also optimised at the MP2/6-31+G(d,p) level of theory. Im-proved energies were obtained by single-point calculations at two levels of theory. The first set of energies resulted from B3LYP/6-311+G(3df,2p) calculations. The second set of energies was calculated using the G2(MP2) method[29]

consisting of MP2/6-311+G(3df,2p), MP2/6-311G(d,p), and QCISD(T)/6-311G(d,p) calculations that were combined to provide effective QCISD(T)/6-311+G(3df,2p) energies cor-rected for ZPVE and the number of valence electrons. Ac-cording to definition, the proton affinity of molecule M is the negative enthalpy change of the hypothetical protonation re-action H++ MMH+at 298 K. Therefore, the geometries of the lowest energy conformers of neutral nitroalkanes and their O-protonated ionic counterparts were found and used for accurate single-point energy calculation at two levels of theory (vide supra).

3. Results

3.1. Proton affinities of nitroalkanes

To the best of our knowledge, proton affinities are only available for nitromethane and nitroethane[22]. Since there is also very few theoretical values available[30,31], as a part of the present study we have calculated ab initio the pro-ton affinities of nitroethane, 1-nitropropane, 2-nitropropane, 1-nitrobutane and 2-methyl-2-nitropropane. The results are summarized inTable 1. For the three compounds where previ-ous theoretical or experimental values are available the agree-ment is within the estimated uncertainty of the theoretical method,±7 kJ/mol[29].

3.2. Rate coefficients for the H3O+, NO+and O2+ reactions

The calculated collisional rate coefficient, kc, for the re-actions of H3O+ with the nitroalkanes and the derived rate coefficients, k, for the corresponding NO+and O2+reactions are listed inTable 2. As can be seen, the k values for the NO+ reactions increase with the number of carbon atoms in the nitroalkane molecules. The k values for the O2+reactions are close to their respective kcvalues (within the 20% exper-imental error). The only significant exception is the reaction of O2+•with nitromethane, the k value being somewhat lower at 0.75 kc.

Previous measurements are available for the reaction of H3O+with nitromethane (k reported as 4.1±1 cm3s1[32]

and 4.11±1 cm3s1[33]), agreeing well with our calculated kc.

60 K. Dryahina et al. / International Journal of Mass Spectrometry 239 (2004) 57–65 Table 1

Calculated proton affinities (298 K) of nitroalkanes in kJ/mol

Molecule B3LYP/6-31+G(d,p) B3LYP/6-311+G(3df,2p) G2(MP2)a Previousc

Nitromethane 740 745 747b 754.6

Nitroethane 761 767 766 765.7

1-Nitropropane 769 774 772

2-Nitropropane 774 780 778 777.1

1-Nitrobutane 773 779 776

2-Methyl-2-nitropropane 788 793 790

a Uncertainty estimated as mean average deviation is±7 kJ/mol[29].

b Value taken from[31].

c The values of proton affinities for nitromethane and nitroethane are from[22,33]and that for 2-nitropropane is from[30].

Table 2

Rate coefficients for the reactions of H3O+, NO+and O2+•with the nitroalkanes indicated

Molecule M (Da) αa(10−24cm3) µ(D) H3O+[kc] (10−9cm3s−1) NO+kb[kc] (10−9cm3s−1) O2+•kb[kc] (10−9cm3s−1)

Nitromethane 61 7.37 3.46 [4.6] 0.4 [3.9] 2.9 [3.8]

Nitroethane 75 9.63 3.65 [4.8] 2.4 [4.0] 4.6 [3.9]

1-Nitropropane 89 8.5 3.66 [4.6] 2.5 [3.9] 3.6 [3.8]

2-Nitropropane 89 8.5±1 3.73 [4.7] 2.6 [3.9] 3.3 [3.8]

1-Nitrobutane 103 10.4 3.59 [4.6] 3.4 [3.9] 4.4 [3.8]

2-Methyl-2-nitropropane 103 10.3 3.71 [4.7] 4.2 [3.9] 4.2 [3.8]

Also given are their molecular weights, M, in Daltons, Da, their polarisabilities,α, in units of 1024cm3and their permanent dipole moments,µin Debye, D.

The collisional rate coefficients, kc, calculated using the parameterised trajectory formulation of Su and Chesnavich[24]are given in square brackets.

a The knownαandµvalues[25]are shown in regular type. The estimatedαvalue is shown in italic.

b The rate coefficients, k, for the NO+and O2+reactions are experimentally derived by the procedure described in Section2. The absolute and relative uncertainties in these calculated rate coefficients are±30% and±15%, respectively.

The products of the reactions and their percentage (in parenthesis) are given in Table 3together with the known ionisation energies[34]of the nitroalkanes. The trends in re-activity will be discussed briefly in the following subsections.

3.3. The H3O+reactions

These reactions proceed initially via the formation of a nascent protonated parent molecule, MH+, and this is the only observed product ion for the nitromethane and nitroethane re-actions. Partial fragmentation of the nascent MH+ions occurs for both isomers of nitropropane leading to the formation of the C3H7O+ion; similarly, both isomers of nitrobutane lead to C4H9+ions (seeTable 3).

There is an obvious difference in the products of the re-action of 1-nitropropane and 2-nitropropane. The rere-action of

Table 3

Primary product ions and their percentage (in parenthesis) for the reactions of H3O+, NO+and O2+with the nitroalkanes indicated

Molecule IE (eV)a Productions

H3O+ NO+ O2+

Nitromethane CH3NO2 11.08 CH3NO2·H+(100) CH3NO2·NO+(100) CH3NO2+(90); NO+(10) Nitroethane C2H5NO2 10.90 C2H5NO2·H+(100) C2H5NO2·NO+(100) C2H5+(100)

1-Nitropropane C3H7NO2 10.78 C3H7NO2·H+(95); C3H7O+(5) C3H7NO2·NO+(100) C3H7+(100) 2-Nitropropane (CH3)2CHNO2 10.74 C3H7NO2·H+(40); C3H7O+(60) C3H7NO2·NO+(100) C3H7+(100) 1-Nitrobutane n-C4H9NO2 10.71 C4H9NO2·H+(90); C4H9+(10) C4H9NO2·NO+(85); C4H9+(15) C4H9+(100) 2-Methyl-2-nitropropane (CH3)3CNO2 C4H9NO2·H+(5); C4H9+(95) C4H9+(100) C4H9+(100)

a The values of the ionisation energy are from[47].

1-nitrpropane proceeds thus:

H3O+ +C3H7NO2C3H7NO2·H+ + H2O; 95% (1a) H3O+ +C3H7NO2

C2H5CHOH+ + (HNO +H2O); 5% (1b) Here the major product is the protonated parent molecule.

Both reactions (1a) and (1b) are exothermic. The reac-tion exothermicities of reacreac-tions (1a) and (1b) calculated at the G2(MP2) level of theory are rH298= 84 and 46 kJ/mol respectively. The latter exothermicity is in rela-tively good agreement with value of 60 kJ/mol calculated for the C2H5CHOH+product structure using data from[35](see Scheme 1). Note that the identity of the neutral products can-not be directly determined using SIFT, thus we always list the

K. Dryahina et al. / International Journal of Mass Spectrometry 239 (2004) 57–65 61

Scheme 1. The nitro-to-nitrite group isomerization by an alkyl cation mi-gration for (a) protonated 1-nitropropane C3H7NO2·H+; (b) protonated 2-nitropropane (CH3)2CHNO2·H+.

most exothermic neutral products and indicate the possibility of different products by parentheses.

In contrast to this, in the reaction of 2-nitropropane the closed shell C3H7O+fragment is the major product.

H3O+ + (CH3)2CHNO2

(CH3)2CHNO2·H+ +H2O; 40% (2a)

H3O+ + (CH3)2CHNO2

(CH3)2COH+ +(HNO + H2O); 60% (2b) The exothermicities of (reaction (2a) and (2b) were calcu-lated at the G2 level of theory to be 90 and 93 kJ/mol, respec-tively. It should be noted that structure of oxygen-protonated propanal, C2H5CHOH+, was considered as the product in reaction (2b) whereas protonated acetone, (CH3)2COH+, is considered as the product in reaction (2b).

From the mechanistic viewpoint, reactions (1b) and (2b) require some rearrangement of the reactants which allows C O bonds to form. The nitro-to-nitrite group isomeriza-tion (Scheme 1) by an alkyl caisomeriza-tion migraisomeriza-tion may serve as a plausible explanation. However, it cannot be excluded that such rearrangement takes place in a reaction complex with H2O, since methyl cation transfer has already been described in CH3NO2+[36], CH3CO+[37]and CH3NN+[38] com-plexes with various atoms and small molecules. Especially in the case of larger systems, after the proton transfer the H2O leaving the reaction complex could be trapped for a time suf-ficiently long to be involved in further reaction, a step-wise mechanism, in which the proton transfer (1a) and (2a) takes place as the first step. After the H2O leaves the reaction com-plex, the vibrationally excited C3H7NO2·H+rearranges and dissociates unimolecularly to the products (Scheme 1) as it will be discussed in the following section.

It has been found that a methyl group migrating along the NO2H moiety in protonated nitromethane must proceed over large energy barriers, and so virtually no isomerization was observed during collisional activation[31]. However,

migra-tion of larger alkyl camigra-tions should be easier due to their higher stability, as was actually observed in protonated nitroethane [39]. Stability of the migrating alkyl cations can be espe-cially important when comparing the reactivity of 1- and 2-nitropropane since the higher stability of a migrating 2-propyl cation can result in higher stabilization of the corresponding transition state. So in the case of 1-nitropropane, the frag-mentation shown inScheme 1a requires migration of a less stable 1-propyl cation over a correspondingly higher energy barrier making the dissociation of C3H7NO2·H+to C3H7O+ and HNO much slower. Therefore, protonated 1-nitropropane substantially prevails among the products, whereas in the case of 2-nitropropane, isomerization via the stable 2-propyl cation migration becomes faster, which increases the fraction of C3H7O+and HNO products.

The final step of the proposed reaction mechanism (Scheme 1) involves hydrogen migration from the propyl to the NO group in protonated propyl nitrite. Similar to methyl nitrite[31], protonation of an ether oxygen atom in propyl nitrites leads to the most stable tautomer. Our cal-culations showed, that C3H7O(H)NO+ ions are in fact a complexes of propanols with NO+, as it is indicated by the O N bond lengths which are 1.98 and 1.95 ˚A in 1-propyl nitrite and 2-propyl nitrite, respectively. The migration of the most labile -hydrogen to NO+ is considered to take place producing appropriately C3H7O+ ion and a neutral HNO. This, which from the formal point of view is hy-dride anion abstraction, was found to be the only reaction channel in the 1- and 2-propanol reactions with NO+ [14].

Since the reaction of propanol with NO+is fast (it proceeds at the collisional rate [14]), the nitro-to-nitrite group rear-rangement is most probably the rate-determining step (in the fragmentation shown inScheme 1) that controls the product distribution.

From the G2(MP2) enthalpies of reactions (1a), (1b), (2a) and (2b) it follows that the dissociation of C3H7NO2H+ to CH3CH2CHOH+ and HNO is 38 kJ/mol endothermic and dissociation of (CH3)2CHNO2·H+to (CH3)2COH+and HNO is 3 kJ/mol exothermic. Although the calculation of the isomerization barriers along the reaction path is beyond the scope of this work, from the nature of the SIFT experiment it can be concluded that none of these barriers can exceed the protonation exothermicity reduced by the internal and kinetic energy carried away by the H2O. There are similar differences between the branching ratios of 1-nitrobutane and 2-methyl-2-nitropropane reactions. For 1-nitrobutane the major product is the protonated parent molecule and C4H9+is only a minor product:

H3O+ +C4H9NO2C4H9NO2·H+ + H2O; 90% (3a) H3O+ +C4H9NO2C4H9+ + (HONO +H2O); 10%

(3b) For the 2-methyl-2-nitropropane reaction the protonated parent molecule is the minor product and C4H9+is the major

In document Text práce (2.331Mb) (Stránka 80-124)