Sunteți pe pagina 1din 6

J. Phys. Chem.

1996, 100, 19911-19916 19911

Kinetics and Products of the Gas-Phase Reaction of SO3 with Water

Edward R. Lovejoy,* David R. Hanson, and L. Gregory Huey


NOAA Aeronomy Laboratory, 325 Broadway, Boulder, Colorado 80303
ReceiVed: August 8, 1996; In Final Form: October 7, 1996X

The kinetics of the gas-phase reactions of SO3 with H2O and D2O were studied over the temperature range
250-360 K in N2 with a laminar flow reactor coupled to a chemical ionization mass spectrometer. The SO3
loss is second order in the water concentration, is independent of pressure (20-80 Torr N2, 300 K), and has
a strong negative temperature dependence and a significant H/D isotope effect (kH2O 2kD2O). The yield of
sulfuric acid is 1.0 ( 0.5 per SO3 consumed. These observations are consistent with the rapid association of
SO3 and H2O to form the adduct H2OSO3 which reacts with water to produce sulfuric acid. The first-order
rate coefficients for loss of SO3 by reaction with H2O and D2O are given by kI(s-1) ) (2.26 ( 0.85) 10-43T
exp((6544 ( 106)/T)[H2O]2 and (9.45 ( 2.68) 10-44T exp((6573 ( 82)/T)[D2O]2, where T K and [H2O,
D2O] molecule cm-3. The errors are the uncertainty at the 95% confidence level for precision only. Analysis
of the temperature dependence of the SO3 loss yields an upper limit for the H2O-SO3 bond enthalpy of 13
kcal mol-1.

Introduction barrier. They postulated that the simple reaction of SO3 and
one water molecule to form H2SO4 is unlikely. Molecular beam
The gas-phase atmospheric oxidation of sulfur dioxide
experiments by Hofmann-Sievert and Castleman10 demonstrated
produces sulfur trioxide:1-3
that H2SO4 is a product of the reaction of SO3 with water
OH + SO2 + M f HOSO2 + M (1) clusters. Wang et al.11 studied the gas-phase kinetics of SO3
+ H2O by using a flow reactor with photofragment emission
HOSO2 + O2 f HO2 + SO3 (2) detection of SO3. They measured an upper limit of 6 10-15
cm3 molecule-1 s-1 for the bimolecular rate coefficient in 1-10
The dominant atmospheric loss process for SO3 is reaction with Torr of He at room temperature and concluded that the earlier
gas-phase water: kinetic studies were influenced by heterogeneous chemistry. A
smaller upper limit of 2.4 10-15 cm3 molecule-1 s-1 was
SO3 + H2O ff H2SO4 (3) measured in 85 Torr of air subsequently by Reiner and Arnold12
using a flow reactor coupled to a chemical ionization mass
Sulfuric acid is an important precursor for atmospheric aerosol.4 spectrometer. In a more detailed study, Reiner and Arnold13
SO3 also reacts efficiently with ammonia, producing a stable report a pressure independent (23-195 Torr air) bimolecular
adduct:5,6 rate coefficient of (1.2 ( 0.2) 10-15 cm3 molecule-1 s-1 at
room temperature.
SO3 + NH3 + M f H3NSO3 + M (4)
Recent high level theoretical studies by Hofmann and
It is possible that despite the low atmospheric NH3 concentra- Schleyer14 and Morokuma and Muruguma15 indicate that H2-
tions, H3NSO3 may play a role in particle formation because of OSO3 is bound by about 8 kcal mol-1 relative to SO3 + H2O,
its very low vapor pressure and strong affinity for sulfuric acid.5 and the isomerization of the H2OSO3 adduct to sulfuric acid is
The reaction of SO3 with water has been studied extensively. inhibited by a very large barrier (19 (ref 14) and 24 kcal mol-1
Castleman et al.7 performed fast-flow experiments with mass (ref 15) above SO3 + H2O). Morokuma and Muruguma15 find
spectrometric detection of SO3 and reported a second-order rate that more facile paths to sulfuric acid exist when SO3 interacts
coefficient of 9 10-13 cm3 molecule-1 s-1. Subsequent with two water molecules:
calculations by Holland and Castleman8 suggested that SO3
forms an adduct with water: SO3 + (H2O)2 f H2SO4 + H2O (7)
SO3 + H2O + M f H2OSO3 + M (5) H2OSO3 + H2O f H2SO4 + H2O (8)
which isomerizes efficiently to H2SO4:
Their calculations suggest that the reaction with the water
H2OSO3 + M f H2SO4 + M (6) dimer (7) proceeds without a barrier, and the reaction of H2-
OSO3 with water (8) is inhibited by a small barrier of about 5
Chen and Plummer9 applied a higher level of theory to the kcal mol-1. Morokuma and Muguruma15 also predict that the
SO3/H2O system and also concluded that the reaction of SO3 (H2O)2SO3 species is bound by about 6 kcal mol-1 relative to
and water forms a relatively stable adduct H2OSO3, but found H2OSO3 + H2O but is separated from H2SO4 + H2O by a barrier
that the isomerization of the adduct to H2SO4 is inhibited by a of about 10 kcal mol-1.
Recently Kolb et al.16 studied the SO3 + H2O reaction with
* To whom correspondence should be addressed. an atmospheric pressure turbulent flow reactor coupled to a
Also affiliated with Cooperative Institute for Research in Environmental
Sciences, University of Colorado, Boulder, CO 80309. chemical ionization mass spectrometer. They report that the
X Abstract published in AdVance ACS Abstracts, November 15, 1996. loss of SO3 is second order in the water concentration and
S0022-3654(96)02414-8 CCC: $12.00 1996 American Chemical Society
19912 J. Phys. Chem., Vol. 100, No. 51, 1996 Lovejoy et al.

increases significantly (>10 times) as the temperature is lowered SO2 and O2 over a red-hot Nichrome filament (0.004 in.
from 333 to 243 K. Kolb et al.16 postulate that a significant diameter) fixed about 0.5 cm from the exit of the inlet. Typical
fraction of the SO3 loss likely involves reaction with the water inlet flow rates were about 0.003 STP cm3 s-1 of SO2 and 0.1-
dimer (reaction 7). 0.5 STP cm3 s-1 of O2. An aluminum cylinder (about 3 cm
Matrix isolation studies have identified IR absorptions due long by 3 cm diameter) with a series of longitudinal passages
to the H2OSO3 complex.17-19 Bondybey and English18 report for gas transit was attached to the end of the inlet to enhance
that the complex is stable with respect to isomerization to H2- the conduction of heat from the filament to the reactor walls.
SO4 at 5 K in a Ne matrix. From observed isotope shifts, Typically, 0.5-1 W was dissipated by the filament. With the
Schriver et al.19 conclude that the complexation involves charge Al radiator in place, the hot filament caused < 3 C temperature
transfer with an O-S interaction. Phillips et al.20 have recently drop along the length of the reaction zone. The largest
measured the microwave spectrum of the H2OSO3 species in temperature drops occurred for the lowest reaction temperatures.
the gas phase and obtained accurate stuctural information which The formation of aerosol in the reactor was examined by
is in good agreement with the predictions of ab initio theory.14 measuring aerosol in the reactor effluent with an expansion
counter.22 For the conditions used in the present study no
Despite considerable theoretical and experimental effort, the aerosol was detected in the reactor effluent (<200 particle cm-3,
details of the mechanism of the reaction of SO3 and water are r < 0.1 m). These observations place an upper limit of 0.002
still not resolved. In this work, measurements of the first-order s-1 on the first-order rate coefficient for SO3 loss on aerosol,
rate coefficient for the loss of SO3 in excess H2O and D2O as which is significantly less than the measured rate coefficients.
a function of temperature (250-360 K, 50 Torr N2) and pressure The chemical ionization mass spectrometer used in the present
(20-80 Torr of N2 at 300 K) are presented. The SO3 loss is study has been described in detail previously.23 SO3 was
second order in water concentration and has a strong negative detected by the rapid fluoride transfer reaction with SiF5-:5,24
temperature dependence, little pressure dependence, and a
significant H/D isotope effect. These results are consistent with SiF5- + SO3 f FSO3- + SiF4 (9)
the rapid formation of an H2OSO3 adduct (reaction 5) followed
by a slower reaction of the adduct with water (reaction 8) to The high stability of the reagent ion SiF5- and the product
produce sulfuric acid. The implications of these results to the ion FSO3- allowed for operation of the ion-molecule reactor
understanding of the atmospheric chemistry of SO3 are dis- at elevated pressure, which enhanced the sensitivity of the
cussed. detector. In the present study the ion-molecule reactor was
operated at about 5 Torr with a He flow rate of about 110 STP
Experimental Section cm3 s-1. For these conditions the SO3/SiF5- reaction time was
on the order of 0.1 s giving SO3 detection limits (s/n 1 for 1
The kinetics of the SO3 + H2O reaction were studied by s integration) of about 105 molecules cm-3 in the ion-molecule
measuring the concentration of SO3 at the exit of a high-pressure reactor or 5 106 molecules cm-3 in the SO3/H2O reactor at
laminar flow reactor as a function of the contact distance with 50 Torr pressure. In the present study, the initial (z ) 0 cm)
water vapor. SO3 was detected in the reactor effluent by SO3 concentration in the neutral reactor was typically about 109
chemical ionization mass spectrometry (CIMS). molecules cm-3. At these levels, secondary chemistry involving
The flow reactor was a jacketed Pyrex cylinder with an SO3 reaction products is negligible. The maximum possible
internal diameter of 3.1 cm and a length of 50 cm. Temperature- first-order loss rate coefficient due to secondary chemistry is
regulated methanol (T < 270 K) or ethylene glycol (T > 270 about 0.1 s-1 (109 molecule cm-3 10-10 cm3 molecule-1 s-1)
K) was circulated through the jacket to regulate the reactor which is significantly smaller than the values measured in the
temperature. The main carrier gas N2 was added with water at present study.
the upstream end of the reactor. Typical N2 flows were about Water was detected as the cluster ion with SiF5-:
20 STP cm3 s-1 (STP ) 273 K and 760 Torr) and the pressure
ranged from 20 to 80 Torr, giving, for example, an average SiF5- + H2O + M f SiF5-H2O + M (10)
flow velocity of about 45 cm s-1 at 50 Torr and 300 K. The
entrance and mixing lengths21 were about 10 cm, and the SiF5- was generated by the rapid fluoride-transfer reaction
Reynolds number was about 40. Water was introduced into with SF6-:
the reactor by flowing N2 through a bubbler containing liquid
H2O or D2O. The bubbler was immersed in a water bath to SF6- + SiF4 f SiF5- + SF5 (11)
stabilize the temperature. The pressure was measured at the
exit of the bubbler with a capacitance manometer. The water and SF6- was generated by the reaction of SF6 with thermal
flow rate was calculated by assuming that the N2 leaving the electrons produced with a hot thoriated iridium filament.
bubbler was saturated with water. Saturation was confirmed Sulfuric acid and SO3 were detected by reactions with
by observing that the CIMS H2O signal was proportional to BF4-BF3
the calculated water flow rate. The water-saturated N2 stream
passed through a Teflon membrane filter to remove particles BF4-BF3 + H2SO4 f BF4-H2SO4 + BF3 (12)
and mixed with the main N2 flow before entering the reactor.
The reactor pressure and temperature were measured in the BF4-BF3 + SO3 f FSO3-BF3 + BF3 (13)
middle of the reaction zone with a capacitance manometer and
glass encased thermocouple, respectively. All of the flow BF4-BF3 was generated by fluoride transfer from SF6- to
reactor effluent entered the CIMS through a Teflon fitting BF3:
equipped with a variable orifice to control the reactor pressure.
SO3 was added to the reactor through a movable inlet (0.25 SF6- + BF3 f BF4- + SF5 (14)
in. o.d. Pyrex) which entered the upstream end of the reactor
on axis. SO3 was generated at the end of the inlet by passing followed by clustering of BF4- with BF3:
Gas-Phase Reaction of SO3 with Water J. Phys. Chem., Vol. 100, No. 51, 1996 19913

BF4- + BF3 + M f BF4-BF3 + M (15)

SO3 reacted rapidly with BF4- to give FSO3-, which places


the fluoride affinity of SO3 at or above the fluoride affinity of
BF3, contrary to the accepted ordering.25
H2SO4 and SO3 were also detected by reaction with NO3-
HNO3:13,26

NO3-HNO3 + H2SO4 f HSO4-HNO3 + HNO3 (16)

NO3-HNO3 + SO3 f NO3-SO3 + HNO3 (17)

NO3-HNO3 was generated by the reaction of HNO3 with


SF6-:23

SF6- + HNO3 f NO3-HF + SF5 (18)

NO3-HF + HNO3 f NO3-HNO3 + HF (19)


Figure 1. FSO3- signal as a function of the reaction distance for a
Ultrahigh-purity N2 (99.9995%) and O2 (99.99%) were used range of [H2O] in 50 Torr of N2 at 330 K. [H2O] ) 0.0, 0.51, 1.03,
1.75, and 2.36 1016 molecules cm-3. A background signal of 18 Hz
for the reactor flows and for preparing mixtures of SO2
was subtracted from the data.
(99.98%). CIMS measurements (reaction 10) indicated that the
isotopic purity of the D2O sample (100 D2O/(D2O + HDO
+ H2O) was >90%.

Results and Discussion


The variation of the SO3 signal as a function of reaction
distance and water concentration at 330 K is shown in Figure
1. The first-order rate coefficient for loss of SO3 was derived
from the measured decays by employing the Brown27 algorithm
and using an SO3/N2 diffusion coefficient5 of 87(T/295 K)1.9
cm2 Torr s-1. The first-order decay of SO3 was independent
(<10% change) of the concentrations of SO2 (factor of 5
changes at 255 and 300 K) and SO3 (factor of 5 change at 255
K and a factor of 10 change at 300 K). The first-order rate
coefficients for SO3 loss are plotted in Figure 2 as a function
of H2O and D2O for a range of temperatures.
The first-order rate coefficient for loss of SO3 increases
nonlinearly with the concentration of water. Measurements of
the loss of SO3 in the presence of ammonia confirmed that the
first-order rate coefficient for SO3 loss in the presence of excess
NH3 increases linearly with NH3 concentration as observed
previously.5 Therefore, the curvature observed in the water data
is probably not due to systematic errors in the present
experimental procedure. Kolb et al.16 have observed similar
nonlinear behavior for the water reaction. They proposed that
a significant fraction of the SO3 is lost by reaction with the
water dimer (reaction 7). In this case, the SO3 loss rate can be
approximated by

d[SO3] Figure 2. First-order rate coefficient for SO3 loss in 50 Torr N2 as a


) -k7Kd[H2O]2[SO3] (20) function of (a, top) [H2O] and (b, bottom) [D2O] for a range of
dt
temperatures. The solid lines are fits of the data to kI ) DT exp(C/T)-
where k7 is the second-order rate coefficient for reaction of SO3 [H2O,D2O]2. The dotted lines are individual fits to kI ) B[H2O,D2O]2.
with the water dimer and Kd is the equilibrium constant for water
dimerization. This equation is based on the assumption that temperature dependence associated with an activation energy
the water dimer is in equilibrium with the water monomer. of -13 kcal mol-1 was observed. The enthalpy of formation
Equation 20 predicts an effective first-order rate coefficient for of the water dimer is about - 4 kcal mol-1,28,29 which requires
SO3 loss in excess H2O given by the SO3 + (H2O)2 reaction to have an activation energy of -9
kcal mol-1. Such a large negative activation energy is not
kI ) k7Kd[H2O]2 (21) physically reasonable for an elementary bimolecular reaction.
It is possible that the SO3 + (H2O)2 reaction is not a simple
and a temperature dependence determined by the sum of the elementary reaction and proceeds via the (H2O)2SO3 complex.
activation energy for the SO3 + (H2O)2 reaction and the enthalpy However, the contribution of the water dimer mechanism is
of water dimerization. In the present work, a strong negative probably small because an SO3 + (H2O)2 rate coefficient (k7)
19914 J. Phys. Chem., Vol. 100, No. 51, 1996 Lovejoy et al.

of about 20 times the gas kinetic value is required to reproduce


the SO3 loss observed at 253 K in the present study. Even at
300 K, k7 must be approximately equal to the gas kinetic value
to reproduce the observed SO3 loss. This would be highly
unusual for a reaction which has a six-center transition state.15
The quadratic dependence of the SO3 loss on [H2O], and the
large negative temperature dependence are consistent with the
formation of the H2OSO3 adduct followed by reaction of the
adduct with water (reactions 5 and 8). For this mechanism,
the strong negative temperature dependence can be attributed
mostly to the enthalpy of adduct formation (reaction 5).
The kinetic data were analyzed based on the mechanism of
H2OSO3 adduct formation followed by adduct-water reaction
(reactions 5 and 8). When the adduct equilibration is much Figure 3. Summary of the energetics of the SO3 + H2O + H2O system.
more rapid than the adduct reaction with water, the first-order Enthalpies are estimated from the present study and calculated with
tabulated values (asterisk).30
losses of SO3 and H2OSO3 at times longer than the equilibration
time are given by

k8Ka[H2O]2
kI ) (22)
1 + Ka[H2O]

where k8 is the second-order rate coefficient for the reaction of


H2OSO3 with H2O and Ka is the equilibrium constant for H2-
OSO3 formation. The temperature dependencies of k8 and Ka
are expressed as

k8 ) Ae-Ea/RT (23)

Ka ) 1.362 10-22Te-G/RT ) 1.362 10-22TeS/Re-H/RT


(24)
where the units of Ka are cm3 molecule-1 and G is the Gibbs
free energy change for H2OSO3 adduct formation (G ) H Figure 4. k8Ka/T as a function of 1/T. H2O (filled circles), D2O (filled
squares), Kolb et al.16 (open circle).
- TS).
The temperature-dependent kI data for H2O and D2O were
fit to eq 22 to determine values for A, Ea, and H. The entropy kuma and Muguruma,15 which predict a 5 kcal mol-1 barrier.
changes for H2OSO3 and D2OSO3 formation from SO3 + H2O A small negative activation energy would suggest that the
and D2O were fixed at -0.027 kcal mol-1 K-1 based on the reaction H2OSO3 + H2O f H2SO4 + H2O is not a simple
Hofmann and Schleyer14 value for the entropy of H2OSO3, and elementary bimolecular reaction but may involve a weakly
tabulated values30 for the entropies of SO3 and H2O. The best bound intermediate such as (H2O)2SO3 (see Figure 3). It is
fits were achieved with H + Ea ) -13 kcal mol-1 and H > possible that there is significant tunneling involved in this
-11 kcal mol-1, which indicates that for the conditions of the reaction which may explain why the experimental barrier is
present study only a small fraction of the SO3 is tied up as H2- significantly lower than that predicted by ab initio methods.
OSO3 (i.e., Ka[H2O] , 1). In this case eq 22 reduces to kI Preexponential factors (A, eq 23) of 1.3 10-15 cm3
k8Ka[H2O]2, and the temperature dependence is given by the molecule-1 s-1 for the H2OSO3 + H2O reaction and 5.5 10-16
exponential of the sum H + Ea. Fits of the data to this cm3 molecule-1 s-1 for the D2OSO3 + D2O reaction are
simplified expression gave kI(s-1) ) (2.26 ( 0.85) 10-43T calculated from the fit results and the entropy changes for H2-
exp((6544 ( 106)/T)[H2O]2 and (9.45 ( 2.68) 10-44T exp- OSO3 and D2OSO3 formation (-0.027 kcal mol-1 K-1, see
((6573 ( 82)/T)[D2O]2 where T K and [H2O, D2O] discussion above). The preexponential factors are small, which
molecules cm-3. The errors are the uncertainty at the 95% is consistent with the highly constrained six-center transition
confidence level for precision only. The fits are the solid lines state predicted by Morokuma and Muguruma.15
in Figure 2. These observations suggest an upper limit for The kI vs [H2O,D2O] data (Figure 2) were also analyzed by
the H2O-SO3 bond enthalpy of about 13 kcal mol-1 which is fitting the individual temperature curves to the expression kI )
consistent with the recent ab initio values from Hofmann and B[H2O,D2O]2, and then plotting the logarithm of B/T vs 1/T to
Schleyer14 and Morokuma and Muguruma15 (about 8 kcal determine H + Ea from the slope. The dashed lines in Figure
mol-1). It should be noted that the Morokuma and Muguruma 2a and b are the individual fits to kI ) B[H2O,D2O]2, and log-
calculation underestimates the stability of H2SO4 by 8 kcal (B/T) are plotted vs 1/T in Figure 4. This analysis yields kI(s-1)
mol-1 whereas Hofmann and Schleyer are within 2 kcal mol-1 ) 3.95 10-43T exp(6397/T)[H2O]2 and 8.15 10-44T exp-
of the accepted experimental value. (6619/T)[D2O]2 where T K and [H2O,D2O] molecules
In the context of the proposed mechanism, the -13 kcal cm-3.
mol-1 activation energy is the sum of the enthalpy of H2- To quantify the possible contribution of the water dimer
OSO3 formation from SO3 + H2O and the activation energy of reaction 7, the data were also fit to the expression kI ) (a1T
the H2OSO3 + H2O reaction. The limit on the H2O-SO3 bond exp(c/T) + a2T exp(2013/T))[H2O]2 , which assumes that there
enthalpy then also implies a limit on the activation energy for are two parallel reaction channels, one with an effective
the reaction of the H2OSO3 adduct with water of Ea < 0 kcal activation energy of -4 kcal mol-1 (the H2O dimerization
mol-1. This value contradicts the theoretical results of Moro- enthalpy28,29). This expression fit the data as well as the one
Gas-Phase Reaction of SO3 with Water J. Phys. Chem., Vol. 100, No. 51, 1996 19915

Figure 5. First-order rate coefficient for SO3 loss as a function of Figure 6. Variation of FSO3-BF3 (filled symbols) and BF4-H2SO4
[H2O] at 300 K in 20 (squares), 40 (triangles), and 80 Torr N2 (circles). (open symbols) as a function of reaction distance at 300 K for a range
of [H2O] ) 0 (diamonds), 3.2 (triangles), 6.4 (squares), and 9.7 1015
channel expression, and indicated that a -4 kcal mol-1 channel molecules cm-3 (circles). Background signals (FSO3-BF3 ) 80 Hz
and BF4-H2SO4 ) 1700 Hz) were subtracted from the data. The lines
could contribute 0.2% at 250 K, 4% at 300 K, and 30% at 360 are fits to the data.
K.
The variation of the first-order loss of SO3 due to H2O at combined with a short adduct lifetime (less than 10 ms in 5
300 K as a function of pressure is shown in Figure 5. The Torr of He). The solid lines in Figure 6 are a fit to the data
first-order SO3 loss at 300 K changed by less than 10% over based on a simplified mechanism of first-order loss of SO3 to
the pressure range 20-80 Torr N2. Reiner and Arnold13 also the wall and to reaction with H2O giving H2SO4. A large
reported no significant pressure dependence for the SO3 loss fraction of these data was taken in the mixing region, and it is
over the range 23-195 Torr of air. The lack of pressure not possible to extract kinetic information from these data. At
dependence is consistent with the assumption that the SO3 + longer reaction times the alternate SO3 detection schemes
H2O T H2OSO3 equilibration is rapid relative to the H2OSO3 (BF4-BF3, NO3-HNO3) gave kinetic results similar to those
loss by reaction with water, so that the rate-limiting step is a of the SiF5- scheme. The fits to the SO3 loss and H2SO4
pressure-independent reaction of H2OSO3 with water. On the appearance yielded an H2SO4 wall loss which was about 60%
basis of the lack of a pressure dependence, a lower limit of 1 of the diffusion-limited value. The fits also yielded a value of
10-14 cm3 molecule-1 s-1 can be assigned to the rate 5.9 for the product of the H2SO4 yield and the CIMS detection
coefficient for the association reaction of SO3 and H2O in 20 sensitivity of H2SO4 relative to SO3. It is likely that BF4-H2SO4
Torr N2 at 300 K (reaction 5). This is consistent with the rapid and FSO3-BF3 were in equilibrium with the BF4-BF3, because
association observed for SO3 and NH3 (k ) 2 10-12 cm3 of the relatively high levels of BF3. In this case the relative
molecule-1 s-1 in 20 Torr of N2 at 295 K).5 detection sensitivities are difficult to estimate because the
The present measurements of the first-order rate coefficient thermodynamics of ion-molecule reactions 12 and 13 are not
for loss of SO3 due to reaction with H2O in 50 Torr of N2 known. Product experiments utilizing the NO3-HNO3 detec-
interpolated to 295 K are about 5 times smaller than the values tion scheme gave essentially identical SO3 loss and H2SO4
measured by Kolb et al.16 at 295 K and atmospheric pressure appearance profiles as the BF4-BF3 experiments. The
(see Figure 4). This discrepancy is probably not due to a NO3-HNO3 studies gave a product of the H2SO4 yield and the
pressure effect since in the present work the first-order loss relative detection sensitivity (H2SO4/SO3) equal to 3.1. In these
increased by less than 10% for a factor of 4 increase in pressure. experiments the ions were probably not in equilibrium, and the
This weak pressure dependence has also been documented by relative sensitivities should be given to a good approximation
Reiner and Arnold.13 The first-order SO3 rate coefficients by the ratio of the rate coefficients for the ion molecule detection
measured in the present work are in good agreement with the reactions (kH2SO4 3.0kSO3),13,24,26 giving an H2SO4 yield of 1.0
Reiner and Arnold13 data. For example, at 298 K with [H2O] ( 0.5/SO3 consumed. The uncertainty in the yield reflects the
) (1-10) 1015 molecules cm-3, the present work predicts estimated uncertainty in the relative CIMS sensitivities for H2-
effective bimolecular rate coefficients of (0.2-2) 10-15 cm3 SO4 and SO3. These product results are consistent with the
molecule-1 s-1 which encompass the value measured by Reiner observations by Reiner and Arnold.13
and Arnold13 for similar water concentrations ((1.2 ( 0.2)
The first-order loss of SO3 due to reaction with D2O was
10-15 cm3 molecule-1 s-1 ).
approximately 2 times slower than with H2O and exhibited the
The quadratic dependence of the SO3 loss on [H2O] first
same strong negative temperature dependence. These observa-
reported by Kolb et al.16 is reproduced in the present study. In
tions are in accord with the proposed mechanism involving
contrast, Reiner and Arnold13 fit their data to a linear [H2O]
formation of an H2OSO3 adduct (with a negligible isotope effect)
dependence. The source of this discrepancy is unknown.
followed by reaction of the adduct with water (with a significant
The variation of SO3 and H2SO4 as a function of reaction primary isotope effect). The large isotope effect for the adduct-
distance and [H2O] are shown in Figure 6. The chemical water reaction is consistent with the ab initio six-center transition
ionization schemes used in the present work probably do not state15 in which two O-H bonds are being formed and two
distinguish between the isomers H2SO4 and H2OSO3. However, O-H bonds are being broken.
on the basis of the interpretation of the present measurements,
the fraction of SO3 in the form of the adduct is small ([H2-
Summary
OSO3] < 0.25[SO3] at 253 K, <0.01[SO3] at 273 K, etc.).
Additionally, any adduct entering the CIMS will probably Measurements of kinetics of the gas-phase reactions of SO3
dissociate rapidly and be detected as SO3 due to the large with H2O and D2O were performed over the temperature range
dilution from the neutral reactor to the flowing afterglow 250-360 K in N2 by varying the contact between SO3 and water
19916 J. Phys. Chem., Vol. 100, No. 51, 1996 Lovejoy et al.

in a laminar flow reactor coupled to a chemical ionization mass and C. J. Howard. This work was supported in part by the
spectrometer. Low SO3 concentrations eliminated complications NOAA Climate and Global Change Program.
due to secondary and heterogeneous chemistry. The following
observations regarding the chemistry of the gas-phase reaction References and Notes
between SO3 and water were quantified: the loss of SO3 is first
(1) Stockwell, W. R.; Calvert, J. G. Atmos. EnViron. 1983, 17, 2231.
order in [SO3], second order in [H2O] and [D2O] and indepen-
(2) Gleason, J. F.; Sinha, A.; Howard, C. J. J. Phys. Chem. 1987, 91,
dent of pressure (20-80 Torr of N2, 300 K) and has a strong 719.
negative temperature dependence and a significant H/D isotope (3) Wine, P. H.; Thompson, R. J.; Ravishankara, A. R.; Semmes, D.
effect (kH2O 2kD2O). The yield of sulfuric acid is 1.0 ( 0.5/ H.; Gump, C. A.; Torabi, A.; Nicovich, J. M. J. Phys. Chem. 1984, 88,
SO3 consumed. The present data are consistent with the rapid 2095.
(4) Weber, R. J.; McMurry, P. H.; Eisle, F. L.; Tanner, D. J. J Atmos.
association of SO3 and H2O to form the H2OSO3 adduct which Sci. 1995, 52, 2242.
reacts with water to produce sulfuric acid. In the context of (5) Lovejoy, E. R.; Hanson, D. R. J. Phys. Chem. 1996, 100, 4459.
this proposed mechanism, analysis of the temperature depen- (6) Shen, G.; Suto, M.; Lee, L. C. J. Geophys. Res. 1990, 95, 13981.
dence of the SO3 loss yields expressions for the first-order rate (7) Castleman, A. W., Jr.; Davis, R. E.; Munkelwitz, H. R.; Tang, I.
coefficient for SO3 loss as a function of temperature, as well as N.; Wood, W. P. Int. J. Chem. Kinet. 1975, S1, 629.
an upper limit for the H2O-SO3 bond enthalpy of 13 kcal mol-1. (8) Holland, P. M.; Castleman, A. W., Jr. Chem. Phys. Lett. 1978, 56,
511.
It is also deduced that the H2OSO3 + H2O reaction may have (9) Chen, T. S.; Plummer, P. L. M. J. Phys. Chem. 1985, 89, 3689.
an apparent negative activation energy consistent with the (10) Hofmann-Sievert, R.; Castleman, A. W. Jr. J. Phys. Chem. 1984,
formation of an intermediate (H2O)2SO3 complex. The energet- 88, 3329.
ics derived from the present study are in reasonable agreement (11) Wang, X.; Jin, Y. G.; Suto, M.; Lee, L. C. J. Chem. Phys. 1988,
with recent predictions of ab initio calculations. 89, 4853.
(12) Reiner, T.; Arnold, F. Geophys. Res. Lett. 1993, 20, 2659.
(13) Reiner, T.; Arnold, F. J. Chem. Phys. 1994, 101, 7399.
Atmospheric Implications
(14) Hofmann, M.; Schleyer, P. V. J. Am. Chem. Soc. 1994, 116, 4947.
The atmospheric lifetime of SO3 with respect to reaction with (15) Morokuma, K.; Muguruma, C. J. Am. Chem. Soc. 1994, 116, 10316.
water will be highly variable due to the strong temperature (16) Kolb, C. E.; Jayne, J. T.; Worsnop, D. R.; Molina, M. J.; Meads,
R. F.; Viggiano, A. A. J. Am. Chem. Soc. 1994, 116, 10314.
dependence of the reaction and the quadratic [H2O] dependence.
(17) Tso, T.-L.; Lee, E. K. C. J. Phys. Chem. 1984, 88, 2776.
However, SO3 will generally be short lived throughout the (18) Bondybey, V. E.; English, J. H. J. Mol. Spectrosc. 1985, 109, 221.
atmosphere. In the troposphere the SO3 lifetime will be very (19) Schriver, L.; Carrere, D.; Schriver, A.; Jaeger, K. Chem. Phys. Lett.
short due to high water concentrations (e.g., a lifetime of about 1991, 181, 505.
1 10-4 s at 273 K with 2 Torr of H2O). Other SO3 loss (20) Phillips, J. A.; Canagaratna, M.; Goodfriend, H.; Leopold, K. R. J.
processes will not compete with the water reaction. For Phys. Chem. 1995, 99, 501.
(21) Keyser, L. F. J. Phys. Chem. 1984, 88, 4750.
example, the aerosol scavenging lifetime of SO3 in a dense cloud
(22) Lovejoy, E. R.; Hanson, D. R. J. Phys. Chem. 1995, 99, 2080.
is greater than 10 s. Steady-state SO3 concentrations will be (23) Huey, L. G.; Hanson, D. R.; Howard, C. J. J. Phys. Chem. 1995,
very low in the troposphere. Assuming an SO3 production rate 99, 5001.
from SO2 oxidation of 2 103 molecules cm-3 s-1 ([SO2] ) (24) Huey, L. G.; Lovejoy, E. R. Int. J. Mass Spectrom. Ion Proc., in
2 109 molecules cm-3, [OH] ) 1 106 molecules cm-3, press.
273 K, k(OH+SO2) 1 10-12 cm-3 molecule-1 s-1 3) yields (25) Bartmess, J. E. NIST NegatiVe Ion Data Base, Version 3.0; Standard
Reference Database 19B; National Institute of Standards and Technology:
a steady-state SO3 concentration of 0.2 molecule cm-3. In the Gaithersburg, MD, 1993.
stratosphere, the SO3 lifetime will be longer because of lower (26) Arnold, S.; Morris, R. A.; Viggiano, A. A.; Jayne, J. T. J. Geophys.
water levels (e.g., about 100 s for [H2O] ) 5 1012 molecules Res. 1995, 100, 14141.
cm-3 and 220 K). Aerosol scavenging will generally not be (27) Brown, R. L. J. Res. Natl. Bur. Stand. 1978, 83, 1.
able to compete (SO3 lifetime of several hours for background (28) Curtiss, L. A.; Frurip, D. J.; Blander, M. J. Chem Phys. 1979, 71,
2703.
aerosol loadings) with the water reaction in the stratosphere (29) Szalewicz, K.; Cole, S. J.; Kolos, W.; Bartlett, R. J. J. Chem. Phys.
except for volcanically perturbed conditions (lifetime of hun- 1988, 89, 3662.
dreds of seconds). (30) Chase, M. W., Jr.; Davies, C. A.; Downey, J. R., Jr.; Frurip, D. J.;
McDonald, R. A.; Syverud, A. N. JANAF Thermochemical Tables J. Phys.
Acknowledgment. The authors acknowledge useful discus- Chem. Ref. Data. Suppl.1 1985, 14.
sions with K. R. Leopold, D. Worsnop, J. T. Jayne, R. Bianco, JP962414D

S-ar putea să vă placă și