Biogenic Sulfur in the Environment - American Chemical Society

reaction rate of IO radicals with DMS is rather fast and leads quantitatively to ... that Cl-containing species may play in the recently observed "ozo...
0 downloads 0 Views 1MB Size
Chapter 30

Fourier Transform IR Studies of the Reactions of Dimethyl Sulfoxide with OH, NO3, and Cl Radicals

Downloaded by CORNELL UNIV on May 8, 2017 | http://pubs.acs.org Publication Date: April 27, 1989 | doi: 10.1021/bk-1989-0393.ch030

I. Barnes, V. Bastian, K. H. Becker, and D. Martin Physikalische Chemie FB 9, G.H.S. Wuppertal, Bergische Universität, 400 Wuppertal, Federal Republic of Germany Dimethyl sulfoxide (DMSO) has recently been detected in marine air masses. To date nothing is known about the atmospheric fate of DMSO in the gas phase. Reported here are product and kinetic studies on the reactions of OH, NO3 and Cl radicals with DMSO. The investigations were performed in a 420 l reaction chamber at atmospheric pressure using long path in situ Fourier transform (FTIR) absorption spectroscopy for detection of reactants and products. Using the competitive kinetic technique preliminary rate constants of (6.2 ± 2.2) x 10 , (1.7 ± 0.3) x 10 and (7.4 ± 1.8) x 10 cm s have been obtained for the reaction of OH, NO and Cl with DMSO, respectively. SO and dimethyl sulfone (DMSO ) were major products of the reactions of O H and Cl with DMSO. For NO only D M S O was observed as product. Dimethyl sulfoxide (DMSO) is known to be present in seawater at higher concentrations than dimethyl sulfide (DMS) (1). Because of its low vapor pressure it has largely been neglected in the chemistry of the atmospheric sulfur cycle. However, DMSO has been recently observed in marine rain and marine air masses (2). Further, new laboratory studies (2*4) have shown that the reaction rate of IO radicals with DMS is rather fast and leads quantitatively to the formation of DMSO and I. It has been suggested that the reaction may be a sink for DMS in marine environments (2*4) and could explain the short atmospheric residence times of DMS observed in field measurements over the ocean (5) and in coastal regions (6). In a recent article (2) it has been argued that the reaction of IO with DMS could not be an important sink for DMS because the concentrations of DMSO in the atmosphere and in marine rain are very low compared to those for DMS, MSA (methane sulphonate aerosol), and NSS-SO4 " (non-sea-salt sulfate). However, in contrast to other reduced organic sulfur compounds in the troposphere nothing is known about the atmospheric fate of DMSO. If a rapid sink reaction exists for DMSO then it would invalidate the argument proposed above. The major daytime sink for most organic sulfur compounds in the atmosphere is reaction with O H radicals (8.9) and it has been shown that reaction with NO3 is an important nighttime sink for DMS (10.11). As shown -11

-11

3

-13

-1

3

2

3

2

2

0097-6156/89/0393-O476S06.00/0 • 1989 American Chemical Society

Saltzman and Cooper; Biogenic Sulfur in the Environment ACS Symposium Series; American Chemical Society: Washington, DC, 1989.

2

Downloaded by CORNELL UNIV on May 8, 2017 | http://pubs.acs.org Publication Date: April 27, 1989 | doi: 10.1021/bk-1989-0393.ch030

30. BARNES ETAL.

411

FTIR Studies ofReactions ofDMSO

by the recent study on the reaction of IO with DMS our present knowledge of atmospheric chemistry does not allow the exclusion of the possible involvement of other reactions such as those with halogen atoms or halogen oxides in the tropospheric chemistry of sulfur compounds. Interest in the reaction of Cl atoms with organic substances has presently been renewed because of the role that Cl-containing species may play in the recently observed "ozone hole" over the Antarctic (12). Although there is no significant source of chlorine in the troposphere it is known to be emitted from various chemical industries (12). Hov (12) has shown that chlorine is actively involved in the formation of photochemical oxidants in industrial coastal areas. Thus a knowledge of the reactions of C l with atmospheric sulfur compounds will be of interest for modeling the chemistry of such localized situations. Presented here are, to our knowledge, the first investigations of the reactions of O H , Cl, and NO3 with DMSO. A comparison is made with the corresponding reactions of the radicals with DMS and the possible implications for the atmospheric sulfur cycle are briefly discussed. Experimental The investigations were carried out at 298 ± 2 K in a 4201 Duran glass reaction chamber surrounded by 24 Philips T L A 40W/05 fluorescence lamps. Details of the experimental set-up can be found elsewhere (14.15) and will not be given here. The only modification to the original set-up was the insertion of a quartz tube through the center of the chamber which contained a germicidal lamp (Philips, T U V 40 W) with maximum output at 254 nm. The investigations on the O H and Cl reactions were carried out at 760 Torr total pressure and those on NO3 at 500 Torr using either synthetic air or N + 0 mixtures as diluent gas. OH radicals were generated by the photolysis of either NOx/hydrocarbon/ N / 0 (9.15) or C H O N O / N O / N / 6 (8.9) reaction mixtures using the fluorescence lamps. In a few experiments the photolysis of H 0 at 254 nm was employed (16). Cl radicals were generated by the 254 nm photolysis of COCl . The thermal decomposition of N7O5 prepared by the reaction 0103 with N 0 (1Q) was used as the source of NO3 radicals. The concentration-time behavior of the majority of reactants was followed using long-path in situ Fourier Transform infrared (FT-IR)-absorption spectroscopy. Tne organic reference compounds for the experiments with O H and C l were analyzed by gas chromatography with flame ionization detection (Hewlett Packard, Model 5/10 A) using a 2 m stainless steel column packed with Porasil C. The competitive kinetic technique was used to determine the rate constants for the reactions of OH, Cl, and N O 3 with DMSO (S). The principals of the technique are the same for all three radicals. In the reaction systems studied, provided that DMSO and the reference hydrocarbon R H are removed solely by reaction with the radical X (X = OH, Cl, or N O 3 ) , 2

2

2

3

2

2

2

2

2

2

2

DMSO + X - > products, k

(1)

RH

(2)

t

+ X - - > products, k

2

then the rate constant ratio kj/k is given by 2

(dln[DMSO]/dt)

k

(dln[RH]/dt)

k

t

2

Saltzman and Cooper; Biogenic Sulfur in the Environment ACS Symposium Series; American Chemical Society: Washington, DC, 1989.

478

BIOGENIC SULFUR IN THE ENVIRONMENT

where (dln[DMSO]/dt) and (dln[RH]/dt) are the first order decays of DMSO and reference compound, respectively, obtained from an analysis of appropiate reactions mixtures. However, under the conditions of the present experiments there is a slow loss of DMSO to the reactor wall which must be corrected for. The wall loss was first order in DMSO and wastypicallyof the order (1-2) x IO" s* . It is represented by the term (dln[DMSO]/dt)w which has to be considered in addition resulting in Equation (4),

4

1

(d In [DMSO]/dt) - (d In [DMSO]/dt)w (4)

Downloaded by CORNELL UNIV on May 8, 2017 | http://pubs.acs.org Publication Date: April 27, 1989 | doi: 10.1021/bk-1989-0393.ch030

(d In [RH]/dt)

k"

2

which was used in the rate constant determinations. Propene, cis-2-butene, trans-2-butene, or isobutene were used as the reference hydrocarbons in the experiments. The rate constants used for their reactions with OH, Cl, and NO3, are given in the relevant sections in the text. Reactant concentrations were in the range 5-25 ppm and the concentration of the radical precursors, C H 3 O N O , NO*, H2O2, and N2O5 were typically 20,3-6, 10-40, and 15-30 ppm, respectively. Product analyses were carried out on DMSO/NO^/air, and DMSO/Cl^air or N photolysis systems and the D M S O / N 0 5 / a i r dark reaction system. Reactants and products were measured using the FT-IR facility and the IR spectra were recorded within the range 40(M000 cm' with a resolution of 1 c n r . The corresponding IR absorption spectra were derived from interferograms ratioed against background and converted to absorbance. Typically, between 30 and 60 interferograms were co-added per spectrum and 10 to 15 spectra were recorded over periods up to 30 min. concentrations of known species were determined by computer-aided subtraction of calibrated reference spectra. DMSO and D M S 0 were calibrated as follows: The reactor was filled with 500 Torr of air. Weighed amounts of the substances were then heated and their vapor was swept into the reactor using a flow of air. At a pressure of 760 Torr a spectrum was immediatly recorded. The whole procedure was performed as quickly as possible, usually within 4-5 min, to minimize wall losses. The calibration for DMSO was reproducible to within ± 10%, however, the calibration for D M S 0 showed a scatter of ± 30%. 2

2

1

1

2

2

Results and Discussion O H + D M S O . Irradiation of CH ONO/NO/hydrocarbon/DMSO and N0 /NO/hydrocarbon/DMSO reaction systems were carried out at 760 Tontotal pressure using different mixtures of N + O^ as diluent gas. The measurements were carried out relative to the reaction of O H with either propene (k = 4.85 x IO" exp(504/T) cm s' , (S)), cis-2-butene (k = 1.09 x 3

2

2

12

3

1

1

s" , (8)) as reference hydrocarbon. Figure I shows typical data for a N0 /NO/isobutene/DMSO reaction mixture plotted in the form In C / C versus t. The rate constants for the reaction of O H with DMSO obtained from the analysis of such plots according to Equation 4 are listed in Table I. The values of k measured were independent of the O H source used and an average value of (6.5 ± 2.5) x 10* cm s* has been obtained as rate constant for the reaction of O H with DMSO in 760 Torr air at 298 K, which is approximately a factor of 10 higher than the corresponding reaction of O H with DMS. The reaction of O H with DMS is known to proceed primarily by H abstraction (16.18) and H abstraction rate constants are known to scale with 2

t

p

2

11

3

1

Saltzman and Cooper; Biogenic Sulfur in the Environment ACS Symposium Series; American Chemical Society: Washington, DC, 1989.

30. BARNES ETAL.

FTIR Studies ofReactions ofDMSO

0.0 vfl—-o—o—o—o—o

Isobutene

0

Downloaded by CORNELL UNIV on May 8, 2017 | http://pubs.acs.org Publication Date: April 27, 1989 | doi: 10.1021/bk-1989-0393.ch030

479

-0.5

\ o c

\ DMSO *\

-1.0

-Photolysis

Dark-1.5

10

15 t/min

20

V 25

30

Figure 1. Plots of In Q / C against t for DMSO and isobutene obtained from a DMSO/isobutene/N0 photolysis system in 760 Torr synthetic air at 298 K. 0

2

Table I. Rate constants for the reaction of O H with DMSO obtained in 760 Torr synthetic air at 298 K. For reference rate constants k see text. The errors are la and refer to precision only 2

Reactant DMSO DMSO DMSO

Ref. propene cis-2-butene isobutene

No. Expts. 3 3 6

1

klA

k^lO-Ucm^s- ) 2

2.64 ±0.9 1.10 ±0.4 1.24 ±0.4

6.9 ±2.5 6.2 ±2.2 6.4 ±2.0

average

6.5 ±2.5

Saltzman and Cooper; Biogenic Sulfur in the Environment ACS Symposium Series; American Chemical Society: Washington, DC, 1989.

480

BIOGENIC

SULFUR IN T H E

ENVIRONMENT

C-H bond energies. Therefore, such a dramatic increase in the reactivity of O H towards DMSO compared to DMS would not be expected if an abstraction mechanism was operative in each case. It is possible that the presence of the O in DMSO or the nigher oxidation state of sulfur is increasing the reactivity of the molecule towards an addition type of reaction. The reaction of the addition aduct, OH-DMSO, with O j could lead to an enchancement of the effective rate constant. Such a mechanism has been envoked to explain the dependence of the rate constant for the reaction of O H with CS2 on the partial pressure of 0 in the reaction system (26.27). The detection of both S 0 and D M S 0 as reaction products, as described below, indicates that both addition and abstraction reaction pathways are operative. On the other hand, it is now well established that relative kinetic methods which rely on the photolysis of systems containing NOx for O H production lead to erroneously high rate constants for the reaction of O H with thiols (2) and organic sulfides (16.17) due to uncontrolled processes. In the present work the 0 partial pressure was varied between 50 and 760 Torr and the N O concentration was varied by a factor of 4. The trend in the measurements was towards higher values of the rate constants with increase in the 0 partial pressure. However, the observed increase never exceeded the total error limits of the experiments. Changing the N O concentration had very little effect upon the measured rate constant. Recently, Barnes et al. (9.16) have shown that the difficulties associated with the relative method for the determination of rate constants for reactions of O H with organic thiols and sulfides when using N O containing precursor for O H generation can be overcome by using the photolysis of H Oj> at 254nm as the O H source. A few experiments were performed on DMSO/isobutene/ H 0 mixtures in 760 Torr synthetic air and 760 Torr N at 298 K. The gas phase reaction between DMSO and H 0 in the dark was found to be negligible. However, the rate constants determined for O H + DMSO from the photolysis experiments were always much higher than those obtained using the photolysis of N O or CH3ONO/NO as the O H source and also showed a dependence on the H 0 concentration. The rate constants increased with increasing concentration of H 0 > At present this chemical behavior is not understood. One possible explanation is that H 0 9 is reacting with a OH-DMSO adduct in a similar manner recently reported for the reaction of 02 with a OH-DMS adduct (16.18). This is, however, only speculation and the system needs further investigation to understand the detailed mechanism. Product analysis on the O H + DMSO reaction were performed using DMSO (20 ppm)/ N 0 (4 ppm)/ air photolysis system at 298 K. The readily identified products were S 0 , D M S 0 , CO, HCHO, CH3NO3 and an unstable intermediate compound with IR-absorption peaks at 1766,1303, and 766 cm" . This compound nas recently been assigned to methylsulfinyl peroxynitrate ( C H S ( 0 ) O O N 0 ) from an investigation on D M D S / N O / a i r photolysis systems (12). Figure 2 shows a difference spectrum for a typical investigation obtained by subtracting a spectrum recorded after 12 min irradiation from a spectrum recorded before irradiation. The absorptions due to HNO3 and CH3NO3 have also been subtracted for clarity. After subtraction of all known products the residual spectrum resembles very closely that of an aerosol consisting of CH3SO3H (2Q). The yields of S 0 and DMS0 -were, on a molar basis, 60 ± 10% and approximatly 30%, respectively. Because of a series of difficulties in calibration, wall loss, and aerosol formation it is not possible to indicate whether the observed yield of DMSOo is being over- or underestimated. As stated above the observed products show that both abstraction and addition reaction pathways are operative, 2

Downloaded by CORNELL UNIV on May 8, 2017 | http://pubs.acs.org Publication Date: April 27, 1989 | doi: 10.1021/bk-1989-0393.ch030

2

2

2

x

2

x

x

2

2

2

2

2

2

x

2

2

2

2

2

2

2

1

3

2

?

2

2

Saltzman and Cooper; Biogenic Sulfur in the Environment ACS Symposium Series; American Chemical Society: Washington, DC, 1989.

Downloaded by CORNELL UNIV on May 8, 2017 | http://pubs.acs.org Publication Date: April 27, 1989 | doi: 10.1021/bk-1989-0393.ch030

30. BARNES ETAL.

FTIR Studies ofReactions ofDMSO

481

1.0

R

0.6 1.0 U In ([isobuteneJQ/[isobutene ] )

0.2

t

Figure 2. Plot of Equation (3) for results obtained from a DMSO/isobutene/ N2C>5/air reaction system.

Saltzman and Cooper; Biogenic Sulfur in the Environment ACS Symposium Series; American Chemical Society: Washington, DC, 1989.

482

BIOGENIC SULFUR IN THE ENVIRONMENT OH + CH SOCH

-->

CH3SOCH2 + H2O

(5)

O H + CH3SOCH3 + M - >

CH3SOCH3 + M

(6)

3

3

O'H

The reactions of C H 3 S O C H 2 formed in step (5) are expected to lead to the CH3SO radical, CH3SOCH2 CH SOCH 0 3

2

+0 2

->

2

+ NO

CH SOCH 0 Downloaded by CORNELL UNIV on May 8, 2017 | http://pubs.acs.org Publication Date: April 27, 1989 | doi: 10.1021/bk-1989-0393.ch030

3

2

CH SOCH 0 2

2

(7)

2

->

CH3SOCH3O + NO3

(8)

->

CH3SO + H C H O

(9)

whose further reactions with 0 and N 0 lead to the formation of S 0 (12). DMSO? is probably being formed in the reaction system by the further reaction of the OH-DMSO adduct formed with O2, possibily via (10), 2

2

2

O CH3-SO-CH3 + 0

-->

2

CH3-L-CH3 + H 0

JH

(10)

2

y

One interesting point to note is that the S O 2 yield of 60% observed for D M S O / N 0 2 / a i r photolysis systems is considerably higher than the yield of typically 30% reported for D M S / N 0 / a i r photolysis systems (20.32-35) indicating perhaps differences in the respective oxidation mechanisms. It is not possible to say whether the S 0 is being formed entirely via the H abstraction oxidation pathway of DMSO by O H or if it also being formed partly by the oxidation of the D M S 0 product. Because of difficulties in the calibration of D M S O 2 and its possible further oxidation it is not possible to give information on the relative importance of the abstraction and addition reaction pathways. More work is needed on individual reaction steps in order to obtain a better understanding of the reaction mechanisms. 2

2

2

N Q + DMSO. The rate constant for the reaction of NO3 with DMSO was measured relative to the reaction of NO3 with isobutene using DMSO (20 ppm)/isobutene (10-20 ppm)/N 0«j (15-30 ppm) reaction mixtures in 500 Torr of synthetic air at 298 K. A value of (3.2 ± 0.$) x IO* cm s" as determined in this laboratory relative to NO3 + trans-2-butene (k = (3.06 ± 0.30) x IO" cm s" , (21)} was used as rate constant for the reaction of NO3 + isobutene. The value is in good agreement with the value obtained by Atkinson et al. (22) after correction for uncertainties in the N 2 O 5 equilibrium constant (22) and also with a more recent absolute determination using a discharge flow-mass spectrometer technique by Rahman et al. (24). The experiments were performed by flushing N^Oj from an external 201 bulb into the 420 1 reactor containing DMSO and isobutene in 500 Torr of synthetic air. The decays of DMSO and isobutene were monitored for a period of 5 min using FT-IR spectroscopy. Over the time period of the investigation the wall loss of DMSO was small (3%) compared to reaction with NO3. A typical plot of the results according to Equation (3) is shown in Figure 3. From a total of 6 experiments a rate constant ratio k (NO3 + DMSOJ / k (NO3 + 3

2

13

3

1

13

1

Saltzman and Cooper; Biogenic Sulfur in the Environment ACS Symposium Series; American Chemical Society: Washington, DC, 1989.

3

30. BARNES ETAL.

FTIR Studies ofReactions ofDMSO

483

Downloaded by CORNELL UNIV on May 8, 2017 | http://pubs.acs.org Publication Date: April 27, 1989 | doi: 10.1021/bk-1989-0393.ch030

Absorbance

3300

2900

2500

2100 1700 1300 wavenumbers (cm" )

900

500

1

1

Figure 3. IR spectrum in the range 500 to 3300 cm* for a DMSO (20 ppm)/ NO2 (4 ppm) photolysis mixture in 760 Torr synthetic air. The spectrum is a difference spectrum formed by subtracting a spectrum recorded before irradiation from one recorded after 12 min irradiation. Absorption due to HNO3 and CH3NO3 have been subtracted for clarity.

Saltzman and Cooper; Biogenic Sulfur in the Environment ACS Symposium Series; American Chemical Society: Washington, DC, 1989.

484

BIOGENIC SULFUR IN THE ENVIRONMENT

isobutene) = 0.53 ± 0.10 was obtained. Using the rate constant for NO3 with isobutene quoted above a value of (1.7 ± 0.3) x IO* cm s* is obtained as rate constant for the reaction of NO3 with DMSO. The quoted l a error represents experimental precision only. Product studies on DMSO/N2O5 mixtures were performed in 760 Torr of synthetic air. H N O 3 and DMSO? were the only observed products. H N O 3 can be accounted for entirely by the H N O 3 entering the reactor with the N2O5 and also by the decomposition of N2O5 at the reactor surface. It could, however, also be formed by the abstraction reaction (11), 13

CH3SOCH3 + NO3

->

1

CH3SOCH2 + H N O 3

(11)

The further reactions of C H S O C H are known to lead to the formation of SOj. Since SO2 is not observed in the product spectrum the abstraction reaction is considered to be a minor channel. The results suggest that the major channel is addition of N O j to D M S O probably forming an unstable adduct which decomposes to D M S O 2 and NO2, O 3

Downloaded by CORNELL UNIV on May 8, 2017 | http://pubs.acs.org Publication Date: April 27, 1989 | doi: 10.1021/bk-1989-0393.ch030

3

2

CH3SOCH3 + NO3 - > CH3SCH3 - >

CH3SO2CH3 + N 0

(12)

2

Since an addition reaction pathway appears to play a role in both the mechanisms of the reactions of O H and NO3 with DMSO it is surprising that the reactivity of NO3 towards DMSO is less than that towards DMS, whereas, the opposite is true for O H . This could reflect a difference in fate of the NO3-DMSO and OH-DMSO adducts. The former adduct can probably only decompose via reaction (12), whereas, the OH-DMSO adduct could react with O2 (reaction (10)) which would probably considerably shorten the lifetime of the adduct and lead to a faster overall rate constant for the reaction. This is only speculation and as mentioned in section 3.1 the possibily that secondary chemistry is effecting the O H rate determination needs to be elimated before meaningful comparisons can be made. C l + D M S O . The rate constant for the reaction of C l with DMSO was measured relative to the reaction of C l with propene using DMSO (10-20 ppm)/propene (10-20 ppm)/COC12 (20-30 ppm) reaction mixtures in 760 Torr ot synthetic air and 760 Torr of N at 298 K. A rate constant of (24.4 ± 0.8) x 10-" cm s* was taken for the reaction of Cl with propene (2£). As for the studies on O H + DMSO a correction had to be made for the wall loss of DMSO which contributed, in this case, approximately 40% to the DMSO decay. After correction for these wall losses rate constants ratios, k(Cl + DMSO)/k(Cl + propene), of 0.3 ± 0.08 and 0.22 ± 0.06 were obtained for the experiments performed in air and N2, respectively, from a total of five experiments in each diluent gas. Using the rate constant for C l + propene quoted above values of (7.4 ± 1.8)x 10* cm s* in air and (5.4 ± 1.4)x 10" * cm s in N were obtained for the reaction of Cl with DMSO. The errors are la and represent precision only. The rate constant obtained for the reaction of Cl with DMSO in is approximately 40% lower than the rate constant obtained for the reaction in air. The results suggest a possible O3 dependence of the reaction rate. The rate constant for the reactions of O H with DMS (16.18) and C S and of C l with C S (2&) are known to be dependent on the oxygen concentration. However, in the present case further kinetic studies are needed 2

3

1

11

3

1

-1

2

2

2

Saltzman and Cooper; Biogenic Sulfur in the Environment ACS Symposium Series; American Chemical Society: Washington, DC, 1989.

1

3

30. BARNES ETAL.

FTIR Studies ofReactions ofDMSO

485

to exclude the possibility of the interference of secondary reactions before the observed increase can be attributed more certainly to an " 0 effect". The rate constant obtained for Cl + DMSO in air is slightly more than a factor of 2 slower than the value of (2.0 ± 0.3) xlOr cm s for the reactiom Cl + D M S (this laboratory, unpublished results). No effect of the 0 concentration was observed for the reaction of Cl with DMS. However, this rate constant is already close to the collision frequency and a small 0 effect could remain undetected within the precision of the present experimental method. In air the reaction of Cl with DMSO leads to the formation of SO^ DMSOb, CO, HCHO, and HCOOH with yields of approximately 42% (S), 14% (S), 15% (C), 18% (C), and 2% (C), respectively. As discussed earlier it is not known whether the yield of DMS02 is being over- or underestimated. The total sulfur yield was 56vo indicating that probably a major sulfur containing product has not been detected. With tne inclusion of the contribution from DMSO? the total carbon yield was 63%. The formation of DMSOo and S 0 as products indicates that both, addition (13) and abstraction (14) pathways are operative, 2

10

3

_1

2

Downloaded by CORNELL UNIV on May 8, 2017 | http://pubs.acs.org Publication Date: April 27, 1989 | doi: 10.1021/bk-1989-0393.ch030

2

2

Cl + CH3SOCH3 - >

C H S O C H + HC1

(13)

Cl + CH3SOCH3 - >

CH3SOCH3

(14)

3

2

The further reactions of C H S O C H are expected to yield S 0 . The D M S O is possibly being formed by further reaction or the Cl- adduct produced in reaction (14) with O^ possibly via, 3

2

2

?

O CH3SOCH3 + 0

2

- > C H s ' c H - > C H S 0 C H + CIO 3

3

3

2

3

(15)

Cl

Before the missing products are identified it is not possible to discuss the dominating channeiany further. Conclusions Rate constants for the reaction of O, OH, Cl, N O 3 , and O 3 with DMS and DMSO are listed in Table II. With the exception of the rate constants for the reactions of O, OH, N O 3 , and O 3 with DMS which are from references (16.18. 29-31). respectively, all other values are either results of this study or of unpublished work from this laboratory. The value quoted for O H + DMS is the average of the two latest studies on this reaction (16.18). Apart from the reaction of O H with DMSO which is a factor of 10 faster tnan the reaction of O H with DMS, DMSO is less reactive towards the other atmospheric species O, Cl, N O 3 , and O 3 than DMS. Of the radicals listed in Table II Cl shows the highest reactivity towards both DMS and D M S 0 . No field measurements of the tropospheric concentration of Cl have been reported. Its tropospheric concentration is almost certainly lower than the typical atmospheric O H concentration of l x l O molecules c n r However, a Cl concentration of 10 molecules cm" , which may occur in industrial regions, would suffice to make the Cl + DMS reaction a significant atmospheric sink for DMS in addition to 2

6

3

3

Saltzman and Cooper; Biogenic Sulfur in the Environment ACS Symposium Series; American Chemical Society: Washington, DC, 1989.

4

486

BIOGENIC SULFUR IN THE ENVIRONMENT

reaction with O H in such areas. In general the reaction of Cl with DMSO will be of no tropospheric importance. As pointed out in section 3.1 relative methods which involve the use of N O containing precursors for O H production are known to give too high rate constants for the reactions of O H with mercaptans and organic sulfides probably due to secondary chemistry in the reaction systems. The possibility that secondary chemistry is affecting the determination of the rate constants for O H + D M S u in the present study cannot be completely excluded. The quoted value should, therefore, be used with some caution until the reaction rate constant is supported by additional studies applying different techniques. However, if this rate constant is used in conjunction with an average O H concentration of 1 x 10 molecules c n r a value of -3.9 h is obtained as atmospheric lifetime for DMSO due to reaction with O H . This value is approximately a factor of 10 lower than the corresponding lifetime of DMS. DMSO and DMSO2 are considerably less volatile than DMS and are also water soluble. Dissolution of DMSO and DMSO? into suspended droplets could also constitute a major sink for DMSO and DMSO2 in the atmosphere. To our knowledge nothing is known about these loss processes for DMSO or D M S 0 . A combination of loss through dissolution ana reaction with O H could lead to rather short atmospheric lifetimes for DMSO. Obviously our understanding of the involvement of DMSO and DMSO2 in the atmospheric sulfur cycle is at present very limited. The results from laboratory and field measurements as briefly discussed here imply that DMSO and DMSO2 are important products of the photooxidation of DMS which is the major source of reduced sulfur to the marine atmosphere. Further kinetic stuaies are needed, however, to elucidate the role of DMSO and DMSO2 in the atmospheric sulfur cycle.

x

Downloaded by CORNELL UNIV on May 8, 2017 | http://pubs.acs.org Publication Date: April 27, 1989 | doi: 10.1021/bk-1989-0393.ch030

6

3

2

3

1

Table II. Comparison of the rate constants (in units of cm s* ) for the reaction of O, OH, Cl, NO*, and O3 with DMS and DMSO, respectively, for conditions of 760 Torr total pressure and 298 K

0

OH Cl N03 O3

DMS

DMSO

(3.3 ± 0.3) x 10-" 7x10-12 (2.0 ± 0.3) x 10-1" (8.1 ± 1.3) x IO-" < 8 x l& 19

(1.2 ± 0.2) x 10-u (7.1 ± 2.5) x 10-u (7.4 ± 1.8) x IO-" (1.7 ± 0.3) x 10-13 < 5 x 10-19

Acknowledgments This research was supported by the "Bundesminister fur Forschung und Technologie and the "Umweltbundesamt". M

Literature Cited 1. Andreae, M . O. Limnol. Oceanogr. 1980, 25, 1054. 2. Harvey, G. R.; Lang, R.F. Geophys. Res. Letts. 1986, 13, 49. 3. Barnes, I.; Becker, K. H.; Carlier, P.; Mouvier, G. Int. J. Chem. Kinetics 1987, 19, 489.

Saltzman and Cooper; Biogenic Sulfur in the Environment ACS Symposium Series; American Chemical Society: Washington, DC, 1989.

30. BARNES ETAL.

FTIR Studies ofReactions ofDMSO

487

4. Martin, D.; Jourdain, J. L.; Laverdet, G.; Le Bras, G. Int. J. Chem. Kinetics 1987, 19, 503. 5. Nguyen, B. C.; Bergeret, C.; Lambert, G . In Gas Transfer of Water Surfaces: Brutsaert, W.; Jirke, G. H., Eds.; Reidel: Dordrecht, 1984; pp 53945.

Downloaded by CORNELL UNIV on May 8, 2017 | http://pubs.acs.org Publication Date: April 27, 1989 | doi: 10.1021/bk-1989-0393.ch030

6. Carlier, P. In Atmospheric Ozone; Zerefos, C. S.; Ghazi, A . B., Eds.; Reidel: Dordrecht, 1985; pp 815-19. 7. Charlson, R. J.; Lovelock, J. E.; Andreae, M . O.; Warren, S. G . Nature 1987, 326, 655. 8. Atkinson, R. Chem. Rev. 1986, 86, 69. 9. Barnes, I.; Bastian, V.; Becker, K. H.; Fink, E. H.; Nelsen, W. J. Atmos. Chem. 1986, 4, 445. 10. Atkinson, R.; Pitts, J. N. Jr.; Aschmann, S. M . J. Phys. Chem. 1984, 88, 1584. 11. Wallington, T. J.; Atkinson, R.; Winer, A . M.; Pitts, J. N., Jr. J. Phys. Chem. 1986,90,4640. 12. McElroy, M . B.; Salawitch, R. J.; Wofsy, S. C.; Logan, J. A . Nature. 1986, 321, 759. 13. Hov, O. Atmos. Environ. 1985, 19, 471. 14. Barnes, I.; Bastian, V.; Becker, K. H.; Fink, E . H.; Klein, Th.; Nelsen, W.; Reimer, A . ; Zabel, F. Untersuchung der Reaktionssysteme N O / C l O / H O unter troposphrischen Bedingungen, B P T Bericht 1/84 ISSN 01761/0777 GSF, München 1984. 15. Barnes, I.; Bastian, V.; Becker, K . H . ; Fink, E . H . ; Zabel, F. Atmos. Environ. 1982, 16, 545. 16. Barnes, I.; Bastian, V.; Becker, K. H . Int. J. Chem. Kinetics 1988, 20, 415. 17. Wallington, T. J.; Atkinson, R.; Tuazon, E . C.; Aschmann, S. M . Int. J. Chem. Kinetics 1986, 18, 837. 18. Hynes, A . J.; Wine, P. H.; Semmes, D. H . J. Phys. Chem. 1986,90,4148. 19. Barnes, I.; Bastian, V.; Becker, K. H.; Niki, H . Chem. Phys. Letts. 1987, 140, 451. 20. Hatakeyama, S.; Izumi, K.; Akimoto, H . Atmos. Environ. 1985, 19, 135. 21. Ravishankara, A. R.; Mauldin, R. L., III. J. Phys. Chem. 1985, 89, 3144. 22. Atkinson, R.; Plum, C. N.; Carter, W. P. L.; Winer, A . M.; Pitts, J.N., Jr.; J. Phys. Chem. 1984, 88, 1210. 23. Atkinson, R.; Aschmann, S. M . ; Winer, A . M . ; Carter, W. P. L., Environ. x

x

Sci. Technol. 1985, 19, 87.

Saltzman and Cooper; Biogenic Sulfur in the Environment ACS Symposium Series; American Chemical Society: Washington, DC, 1989.

x

488

BIOGENIC SULFUR IN T H E

ENVIRONMENT

24. Rahman, M . M.; Becker, E.; Schindler, R. N . Ber. Bunsenges. Phys. Chem. 1987,92,91. 25. Atkinson, R.; Aschmann, S. M . Int. J. Chem. Kinetics 1985, 17, 33. 26. Barnes, I.; Becker, K. H.; Fink, E. H.; Reimer, A.; Zabel, F.; Niki, H . Int. J. Chem. Kinetics 1983, 15, 631. 27. Jones, B. M . R.; Burrows, J. P.; Cox, R. A.; Penkett, S. A . Chem. Phys.

Letts. 1982, 88, 372.

Downloaded by CORNELL UNIV on May 8, 2017 | http://pubs.acs.org Publication Date: April 27, 1989 | doi: 10.1021/bk-1989-0393.ch030

28. Martin, D.; Barnes, I.; Becker, K. H . Chem. Phys. Letts. 1987, 140, 195. 29. Lee, J. H.; Tang, I. N.; Klemm, R. B. J. Chem. Phys. 1980,72,1793. 30. Wallington, T. J.; Atkinson, R.; Winer, A . M . ; Pitts, J. N . , Jr.; J. Phys. Chem. 1986,90,5393. 31. Atkinson, R.; Carter, W. P. L. Chem. Rev. 1984, 84, 437. 32. Hatakeyama, S.; Akimoto, H . J. Phys. Chem. 1983, 87, 2387. 33. Grosjean, D.; Lewis, R. Geophys. Res. Letts. 1982, 9, 1203. 34. Grosjean, D.; Environ. Sci. Technol. 1984, 18, 460. 35. Niki, H . ; Maker, P. D.; Savage, C. M . ; Breitenbach, L. P. Int. J. Chem. Knetics l983, 15, 647. RECEIVED August 11, 1988

Saltzman and Cooper; Biogenic Sulfur in the Environment ACS Symposium Series; American Chemical Society: Washington, DC, 1989.