T Vs Rate

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Temperature Dependence of the Rate

Constant for the Reaction OH H2S +


C. L. LIN
J e t Propulsion Laboratory, California Institute of Technology, Pasadena, California
91109

Abstract
Absolute rate constants for the reaction of OH with H2S have been measured over the
temperature range of 239425 K using the flash photolysisresonance fluorescence technique.
The results showed that the rate constants deviate slightly from Arrhenius behavior but can
still be represented adequately by the following Arrhenius equation:
k11 = (7.8 f 2.6) X exp[-(146 f 105)/T] cm”/molec. s
Comparisons with recent literature values are presented.

Introduction
The source of H2S in the earth’s atmosphere have been extensively in-
vestigated, and the atmospheric concentration has been measured in recent
years [l-31. The reaction of OH with H2S is generally believed to be the
major sink [1,2] for HZS. Therefore in order to estimate the chemical
lifetime of atmospheric H2S, the rate constant for its reaction with OH
under atmospheric conditions must be known accurately.
The rate constant of this reaction has been measured by several inves-
tigators [4-101 using different techniques. Room temperature rate con-
stants obtained range from 3.1 to 5.5 X 10-l2 cm3/molees. Two types of
temperature dependence of the rate constant have been reported, one
following very closely the Arrhenius behavior [5,7,8]and the other showing
non-Arrhenius behavior, with lz (7’)going through a minimum near room
temperature. The latter phenomenon has been reported by Leu and Smith
[lo] and Michael et al. [6]. Further, those studies that found Arrhenius
behavior obtained wide differences in activation energy. Westenberg and
DeHaas [8] reported 880 kcal/mol, while Perry et al. [5] obtained rate
constants with no activation energy. Wine et al. [7] reported a small ac-
tivation energy of 110 kcal/mol. Because of these discrepancies it was felt
that additional rate constant measurements were needed.
In this work absolute rate constants for the reactions of OH radicals with
hydrogen sulfide have been determined over the temperature range of
239-425 K, using a flash photolysis-resonance fluorescence technique.

International Journal of Chemical Kinetics, Vol. 14,593-598 (1982)


(c) 1982 John Wiley & Sons, Inc. CCC 0538-8066/82/050593-06$01.60
594 [,IN

Experimental
The apparatus and technique used in this work have been described
previously [ 111. Some modifications appropriate for OH production and
detection have been made. All the windows were replaced with Suprasil
quartz instead of MgF2. An interference filter with band pass centered
near 308 nm (100A FWHM) was placed in the fluorescence light path
leading to the photomultiplier tube to exclude scattered light from detec-
tion. A reflecting mirror was used to refocus the resonance lamp image
and thus enhance the resonance light intensity. In addition, two quartz
lenses, one in the resonance light path and the other in the fluorescence light
path, were used to increase the signal-to-noise ratio of the OH fluorescence
signal.
The OH radicals were produced by the pulsed ultraviolet photodisso-
ciation ( A 2 165 nm) of H20 and were monitored as a function of time after
the flash by OH fluorescence near 308 nm ( A 2 2 + - X 2 n ) . Flash energies
of 36-100 J per flash were used, and the resonance fluorescence signals were
accumulated for 200-2000 flashes, depending on the signal strengths. OH
radical half-lives ranged from 3.6 to 77 ms, and the signals were followed
over a t least two orders of magnitude. In order to avoid the accumulation
of photolysis or reaction products, all experiments were carried out under
slow flow conditions. H2S diluted with Ar gas was flowed from three in-
terconnected 5-L glass bulbs. A separate Ar gas flow was bubbled through
a water bubbler maintained a t constant temperature and used as the source
of water vapor. A third gas flow of pure Ar gas was used to further dilute
the gas mixture. All three flows converged in a mixing bulb before entering
the reaction cell. All flowmeters were calibrated with a Hastings Mini-Flo
calibrator, and MKS Baratron capacitance gauges were used for pressure
measurements.
The gases used had the following purity levels, according to the manu-
facturers: Ar 2 99.999% and H2S 2 99.5%. They were used without fur-
ther purification.

Results
Under the pseudo-first-order experimental conditions used ([HZS] >>
[OH]),the decays of the OH radical concentrations [OH] are given by the
rate expression

where [OHIOand [OH], are the concentrations of [OH] a t times t o and t ,


respectively, So and St are the corresponding resonance fluorescence in-
RATE CONSTANTS FOR REACTION OH + H2S 595

tensities, ko is the first-order rate constant for removal of OH in the absence


of added H2S, and kII is the second-order rate constant for the reaction
OH + H2S - H20 + HS
In all experiments, exponential decays of the resonance fluorescence
signal were observed, and the measured decay rates were found t o depend
linearly on the concentration of added H2S, except a t high flash energies
(>120 J). A t high flash energy the rate constant obtained was found to be
+
larger, probably due to the secondary reaction OH HS. There are three
possible sources of HS,

+ H2S
H2S
- H20 + HS
-
+ hv(>165 nm) H + HS
OH
H + H2S - H2 + HS
AH = -19.2 kcal/mol
AH = -13.3 kcal/mol
T h e flash photolysis light would be confined to a wavelength region of
about 165-175 nm, because the quartz window cuts off the light effectively
below 165 nm and the output of the medium-pressure (-100 torr) Xe flash
lamp used in the present experiments would be limited to about the
150-175-nm region [12] (see [l2, p. 1151). The integrated absorption cross
sections [12] for H2S in this flash wavelength region are smaller than for
H20 (see [l2, pp. 202,2041). Therefore the production of HS radicals from
direct photolysis was not important, especially because the H,O-to-H,S
concentration ratios used in all the experiments were greater than 30. The
HS radicals produced by the last two reactions would not complicate the
rate measurements as long as the OH concentration was not too high. In
our experiments typical initial OH concentrations produced by the flash
were estimated to be about 0 . 5 1 X 10" molec/cm3. Under these conditions
the observed OH signal decay was linear over a t least two orders of mag-
nitude. In addition, a set of experiments a t 295 K was performed where
the H20 and H2S concentrations were maintained constant, while flash
energy was varied. A variation of the flash energies from 36 to 100 J showed
no difference in the rate constant.
All the data for flash energies 5100 J are summarized in Table I, and
plots of the first-order rate constants against H2S concentrations for four
temperatures are shown in Figure 1 (for 295 K only one set of data is shown).
The slopes and their intercepts yield k11and ko, respectively. An Arrhenius
plot for the temperature dependence of rate constants is shown in Figure
2. Also included in Figure 2 are data from other investigators.

Discussion
A least-squares fit of our data in the Arrhenius form yields
k11 = (7.8 f 2.6) X 10-l2 exp[-(146 f 105)/T] cm3/molec s -
where errors are 20 standard deviation.
596 LIN

I. Rate data for the reaction of OH + H2S.


TABLE

239 50 29-38 0-0.687 81 9 4.41 f 0.60.

295 54 24-31 09.849 81-100 U 4.46 t 0.26

50 U4-131 0-1.255 36100 12 4.lb t 0.24


90 52-55 0-0.924 b4-01 0 4.63 t 0.58
4 . U t 0.48

375 50 35-41 0-1.321 71-98 11 5.U t 0.40

425 50 96-105 0-1.053 81 9 5.03 t 0.76

a Uncertainties are 2a standard deviations.

The present results agree, in general, very closely with the data of Michael
et al. [6] and are parallel in trend with the data of Leu and Smith [lo],al-
though the values of the latter are in general lower at corresponding tem-
peratures. If we examine the rate constants obtained by different studies
shown in Figure 2, we can conclude that except for the high activation en-
ergy and high rate constants at high temperatures reported by Westenberg
and DeHaas [8] and the low room temperature rate constant (3.1 X
cm3/molec.s) reported by Stuhl [4], our rate constants are in reasonable
agreement (towithin 20%) with all the rest of the literature values. Further,

-I
U
240
220
200
180
160
4
t 140
Y
- 120
1 00
80
60
40
20
0
0.4 0.8 1.2 1.6 2.0 2.4 2.8 3.2 3.6
-3
[H2S] molecule cm

Figure 1. Plots of the OH first-order rate constant against H2S concentration at


239,295,375, and 425 K. Data points are displaced vertically +20, +40, and +60
s-l for T = 295,375, and 425 K, respectively.
597

30 I I I

20 -

'-.
\..
10 -
\-.
'-.

1 .o 2.0 3.0 4.0

~ o ~ / TK ,- ~
Figure 2. Arrhenius plots for the reaction OH + H2S. a, this work, error bars
are 20 standard deviation; 0-Leu and Smith; A-Michael et al.; --Wine et
al.; - - - ,Perry et al.; - - - -Westenberg and DeHaas; A-Cox and Sheppard;
0-Stuhl.

the combined data (except the data of Stuhl and Westenberg and DeHaas)
show a positive though weak temperature dependence above room tem-
perature and a nearly flat or very weak negative temperature dependence
below room temperature.
The deviation of the data for the OH +
H2S reaction from simple
Arrhenius behavior has been discussed in detail in the two recent publi-
cations of Michael et al. [6] and Leu and Smith [lo]. A two-path scheme
(Le., direct attack and complex formation), where the direct attack path
dominates at high temperatures while the complex formation becomes
important at low temperatures, was suggested by Michael et al. However,
both our data and that of Leu and Smith [ 101 seem to be inconsistent with
this possibility, because the rate constants at low temperatures showed no
variation with pressure.
Michael et al. have performed a transition-state theory calculation and
concluded that the calculated rate constants were slightly too high to fit
their data on an absolute basis. However, their calculations did show that
the curved Arrhenius behavior shown by the combined data in Figure 2 can
be accounted for by a simple abstraction mechanism with zero energy
598 LIN

barrier. In view of the absence of any direct evidence for a complex


mechanism, we feel that the simple abstraction model is adequate to explain
all the most recent data for this reaction.

Acknowledgment
The author wishes to thank W. B. DeMore for reviewing this manuscript.
This work was supported by the National Aeronautics and Space Admin-
istration under contract NAS7-100.

Bibliography
[I] W. Jaeschke, H. Claude, and J . Herrmann, J . Ceophys. Res., 85,5639 (1980).
[2] T. E. Graedel, Reu. Geophys. Space Phys., 15,421 (1977).
[3] R. Delmas and J. Bandet, J . Geophys. Res., 85,4468 (1980).
[4] F. Stuhl, Rer. Bunsenges. Phys. Chem., 78,231 (1974).
[5] R. A. Perry, R. Atkinson, and J . N. Pitts, Jr., J. Chem. Phys., 64,3237 (1976).
(61 J. V. Michael, D. F. Nava, W. D. Brobst, R. P. Borkowski, and L. J. Stief, J. Phys. Chem.,
86.81 (1982).
171 P. H. Wine, N. M. Kreutter, C. A. Gump, and A. R. Ravishankara, J . Phys. Chem., 81,
2660 (1981).
[8] A. A. Westenberg and N. DeHaas, J. Chem. Phys., 59,6685 (1973).
[9] R. A. Cox and D. Sheppard, Nature, 284,330 (1980).
[lo] M. T. Leu and R. H. Smith, J. Phys. Chem., 86,73 (1982).
1111 C. L. Lin and M. T. Leu, Int. J . Chem. Kinet., in press.
[12] H. Okabe, “Photochemistry of Small Molecules,” Wiley, New York, 1978.

Received October 6, 1981


Accepted November 17,1981

You might also like