Sie sind auf Seite 1von 9

Environ. Sci. Technol.

2005, 39, 1632-1640

greatly increased the potential for accidental environmental


Atmospheric Chemistry of Hydrazoic releases and for human exposure to this highly toxic material.
Acid (HN3): UV Absorption Aqueous sodium azide is readily hydrolyzed to yield hydrazoic
acid, HN3 (pKa 4.7), a volatile substance that partitions
Spectrum, HO• Reaction Rate, and strongly to the gas phase (KH ) 12 M atm-1) under

Reactions of the •N3 Radical


atmospheric conditions (2). For example, even at concentra-
tions as low as 6.5 ppm (m/v) NaN3 in the aqueous phase
(pH 6.5) the gas-phase concentration reaches the threshold
JOHN J. ORLANDO* AND limit value of 0.11 ppmv (as hydrazoic acid gas) so there is
GEOFFREY S. TYNDALL interest in understanding the fate of atmospheric hydrazoic
acid. The problem of significant azide releases to the
Atmospheric Chemistry Division, National Center for
Atmospheric Research, Boulder, Colorado 80305
environment is not a hypothetical one. For example, the town
of Mona, UT was evacuated in 1996 afer a tanker truck hauling
ERIC A. BETTERTON 80 55-gallon drums of NaN3 overturned (1). The problem of
azide disposal will remain for decades, given the many
Department of Atmospheric Sciences, University of Arizona, millions of kilograms of NaN3 that is currently being carried
P.O. Box 210081, Tucson, Arizona 85721-0081
by the nation’s automobile fleet (1).
JOE LOWRY AND STEVE T. STEGALL
Although the tropospheric fate of HN3 has not been the
subject of systematic study, sufficient data are available to
National Enforcement Investigations Center, U.S. E.P.A., indicate that reaction with OH and photolysis are likely
Denver Federal Center, P.O. Box 25227, tropospheric removal processes. Hack and Jordan (3) studied
Denver, Colorado 80225 the reaction of HO• with HN3 via the flash photolysis of H2O2/
HN3/He mixtures, with HO• detection via pulsed LIF and
reported a value for k1 of 1.3 × 10-12 cm3 molecule-1 s-1.
Processes related to the tropospheric lifetime and fate of HO• + HN3 f products (1)
hydrazoic acid, HN3, have been studied. The ultraviolet
absorption spectrum of HN3 is shown to possess a maximum This would imply a tropospheric lifetime for HN3 of about
near 262 nm with a tail extending to at least 360 nm. 10 days (for a diurnally averaged [HO•] ) 106 molecule cm-3).
The photolysis quantum yield for HN3 is shown to be ≈1 Reaction of O(1D) with HN3 appears to occur at essentially
at 351 nm. Using the measured spectrum and assuming unity the gas-kinetic rate, k ) (3.2 ( 1.0) × 10-10 cm3 molecule-1
quantum yield throughout the actinic region, a diurnally s-1 (4), while values reported for the rate coefficient for its
averaged photolysis lifetime near the earth’s surface of 2-3 reaction with Cl-atoms lie in the range (9-15) × 10-13 cm3
days is estimated. Using a relative rate method, the rate molecule-1 s-1 (5-8). Given the relatively large value of k1
and the higher abundance of OH compared to Cl and O(1D),
coefficient for reaction of HO• with HN3 was found to be (3.9
these latter two processes are not likely to be of any
( 0.8) × 10-12 cm3 molecule-1 s-1, substantially larger atmospheric importance.
than the only previous measurement. The atmospheric HN3 The UV spectroscopy and photochemistry of HN3 has been
lifetime with respect to HO• oxidation is thus about 2-3 studied in considerable detail (e.g., refs 9-34), although key
days, assuming a diurnally averaged [HO•] of 106 molecule data for assessment of the importance of tropospheric
cm-3. Reactions of •N3, the product of the reaction of photolysis have yet to be obtained. McDonald et al. (18) have
HO• with HN3, were studied in an environmental chamber reported absorption cross sections for HN3 throughout the
using an FTIR spectrometer for end-product analysis. vacuum and near UV (100-325 nm). Vacuum UV measure-
The •N3 radical reacts efficiently with NO, producing N2O ments by Okabe (115-210 nm) (19) and a single wavelength
with 100% yield. Reaction of •N3 with NO2 appears to determination at 193 nm (24, 25) indicate that these
McDonald et al. (18) data may be low by about 20%. Although
generate both NO and N2O, although the rate coefficient
the McDonald et al. data do not extend beyond 325 nm and
for this reaction is slower than that for reaction with NO. No indeed may be systematically low, they do indicate that
evidence for reaction of •N3 with CO was observed, in tropospheric photolysis of HN3 could be significant (pho-
contrast to previous literature data. Reaction of •N3 with tolysis lifetime ≈2-3 days).
O2 was found to be extremely slow, k < 6 × 10-20 cm3 Numerous photochemistry and photodissociation dy-
molecule-1 s-1, although this upper limit does not namics studies of HN3 have been conducted at wavelengths
necessarily rule out its occurrence in the atmosphere. ranging from 308 nm into the vacuum UV (e.g., refs 13-17,
Finally, the rate coefficient for reaction of Cl• with HN3 was 19-34). The major photolysis products appear to be NH and
measured using a relative rate method, k ) (1.0 ( 0.2) N2 at all wavelengths studied (13, 14, 20-27, 29, 30, 32-34),
× 10-12 cm3 molecule-1 s-1. although a minor process to form H and N3 has also been
observed at 193, 248, and 266 nm (28, 31-33):

HN3 + hν f HN• + N2 (2a)


Introduction
Over the past decade, demand for sodium azide (NaN3), the f H• + •N3 (2b)
principal active ingredient in automobile air bag inflators,
has rapidly risen to exceed 5 million kg per year (1). This has The HN• photoproduct is formed exclusively in the excited
a1∆ electronic state at long wavelength (λ g 248 nm) (25), in
Corresponding author phone: (303)497-1486; fax: (303)497-1411; keeping with spin conservation rules, although other elec-
e-mail: orlando@acd.ucar.edu. tronic states (A,b,c) have been detected at 193 nm and below
1632 9 ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 39, NO. 6, 2005 10.1021/es048178z CCC: $30.25  2005 American Chemical Society
Published on Web 01/22/2005
(19, 21, 22, 25). A near-unity (φ ≈ 0.8) quantum yield for HN• the absorption cell, and is focused onto the entrance slit of
production near 290 nm has been reported (13), and unit a 0.3 m Czerny-Turner spectrograph (equipped with a 300
quantum yields are generally assumed at longer wavelengths. grooves/mm grating), which disperses the light onto a 1024-
Absolute quantum yields for H-atom formation have been pixel diode array detector (EG&G Model 1420). With this
made at three wavelengths (28, 32): φ193 ) 0.14; φ248 ) 0.20; configuration, each pixel is separated in wavelength by about
and φ266 ) 0.04. No H-atom quantum yield data are available 0.25 nm, providing coverage from about 220 to 450 nm with
at longer wavelengths, although H-atom production occurs a spectral resolution of 0.6 nm. The system was calibrated
to at least 280 nm (33), and the energy threshold for this in wavelength via interpolation between the positions of the
process is near 325 nm. emission lines from a low-pressure mercury lamp. Spectra
Photolysis experiments at higher [HN3] and at higher total were obtained from the summation of 100-200 exposures
pressures have also been carried out. Kodama (18), for of the diode array, each exposure being 0.2 s in duration.
example, photolyzed mixtures of HN3 (6.6 × 104 ppmv) in ≈0 Raw spectral data at each pixel, I(λ), obtained in the presence
to 0.8 atm of Xe buffer gas at 313 nm. Chain reactions were of an HN3 sample, were converted to absorbance (base e) via
observed (ΦN2 ) 4.85; λ ) 313 nm) which were thought to comparison with a spectrum, Io(λ), recorded with the
involve the reaction of the HN• photoproduct with additional absorption cell evacuated, i.e., A(λ) ) ln {Io(λ)/I(λ)}. Absorp-
HN3, an unlikely path in the atmosphere. In earlier studies tion spectra were smoothed, and the smoothed data were
of a similar nature, Beckman and Dickinson (11, 12) reported then interpolated in wavelength to obtain absorbance values
ΦHN3 ) 3.6 (λ ) 190 or 254 nm). at 0.5 nm intervals.
Reaction with OH and photolysis of HN3 will result in the Gaseous HN3 samples (for the UV measurements and all
formation of •N3 and HN• radicals, respectively. While HN• other studies conducted herein) were obtained by gently
is thought to react rapidly with O2 to give NO and HO• (35), heating a mixture of approximately 1 g each of NaN3 and a
the atmospheric fate of •N3 is less certain. The most detailed solid carboxylic acid (glycolic acid or stearic acid) (18, 19).
information regarding the reactivity of this species comes For the UV cross section measurements, the concentration
from the work of Hewett and Setser (36). Using a discharge- of HN3 in the absorption cell was determined by pressure
flow/LIF system, these authors reported that the reactions measurement, with the assumption that the gaseous samples
of •N3 with NO, NO2, and CO all occurred with similar rate contained no impurities. Measurements of HN3 samples by
coefficients (k3 ) 2.9 × 10-12 cm3 molecule-1 s-1; k4 ) 1.9 × FTIR spectroscopy revealed no measurable impurities.
10-12 cm3 molecule-1 s-1; and k5 ) 1.8 × 10-12 cm3 molecule-1 Multiple fills of the absorption cell were made, and measured
s-1). Thermodynamically accessible channels for these reac- UV cross sections from the different samples were indis-
tions (based on ∆Hf(N3) ) 99 kcal/mol (33)) are given below: tinguishable. Sample pressures were varied between 0.5 and
4.1 Torr.

N3 + NO f N2O + N2 (3a) Photolysis Quantum Yield at 351 nm. The photodisso-
ciation quantum yield for HN3 at 351 nm was obtained by

N3 + NO2 f 2NO + N2 (4a) subjecting HN3/O2 samples (about 1-1.5 Torr HN3 in 1 atm
O2 buffer gas) to multiple shots from a pulsed XeF excimer
f 2N2O (4b) laser (Lambda Physik Compex 102). The photolysis experi-
ments were conducted in a cylindrical Pyrex cell (25 cm long,
3.5 cm i.d.) equipped with quartz windows, which transmitted
f N2 + N2O + O• (4c)
about 80% of the 351 nm radiation. The laser pulse energy,
defined as the average of the energy before and after the cell
f 2N2 + O2 (4d) as measured with a pyroelectric power meter (Questek P9104),
was typically 90 mJ/pulse which corresponds to 1.6 × 1017

N3 + CO f NCO + N2 (5) photons/pulse at 351 nm. Photolysis experiments were run
for about 3 h with the excimer operating at 8 Hz, resulting
Given the relative atmospheric abundances of these three in a total of roughly 105 laser shots. Pre- and postphotolysis
reactants, reaction of •N3 with CO would thus dominate. No concentrations in the UV cell were determined by FT-IR
products of these reactions were determined in the Hewett spectroscopy, using the environmental chamber/FT-IR spec-
and Setser study, although likely possibilities (channels 3a, trometer system described below. For prephotolysis mea-
4a, 4b, and 5) were suggested. Reaction of •N3 with O2 was surements, a calibrated (1 L) bulb was filled with the same
found to be slow (36), k < 5 × 10-13 cm3 molecule-1 s-1, HN3 sample as the UV cell, and the contents of the calibrated
although this upper limit is not nearly low enough to rule out bulb were then swept into the chamber for analysis. Following
its importance in the atmosphere. irradiation, the concentration of HN3 and photoproducts were
In this work, studies of processes related to the atmo- determined by first expanding the contents of the UV cell
spheric destruction and ultimate fate of hydrazoic acid are into the calibrated bulb and then sweeping the bulb contents
reported, including measurement of (a) the UV absorption into the chamber. Photolysis of HN3 leads to the production
spectrum of HN3 from 215 to 365 nm; (b) its photolysis of reactive NH radical. The effects of the subsequent
quantum yield at 351 nm; (c) the rate coefficient for reaction chemistry of these species (in particular, the formation of
of HN3 with HO•; and (d) the end-products of the reactions OH and resulting consumption of HN3) was accounted for
of •N3 radical with O2, NO, NO2, and CO. Experiments showed in the analysis, as described in the results section. Control
that the diurnally averaged tropospheric lifetime of HN3 is experiments were also conducted in which the UV cell was
about 1-2 days, with both solar photolysis and reaction with filled with HN3/O2 and left to stand for 3 h without irradiation;
HO• contributing about equally to its removal. no loss of HN3 was noted in these experiments.
HO• and Cl• Rate Coefficient Measurements and •N3
Materials and Methods Reaction Product Studies. Rate coefficient measurements
UV Absorption Measurements. The UV absorption spectrum and end-product studies were all carried out in an envi-
for HN3 was determined using a diode array spectrometer ronmental chamber/Fourier transform infrared spectrometer
system that has been described in detail previously (37, 38). system, which has been described previously (39). The
Measurements were made in a 90-cm long Pyrex absorption chamber is 2 m long, is constructed of stainless steel, and
cell equipped with quartz windows. The output from a has a volume of 47 L. It is interfaced via a set of modified
broadband D2 lamp is first collimated, then passes through Hanst-type multipass optics to a Fourier transform spec-

VOL. 39, NO. 6, 2005 / ENVIRONMENTAL SCIENCE & TECHNOLOGY 9 1633


FIGURE 1. Infrared absorption cross sections for HN3 measured in FIGURE 2. UV absorption spectrum (solid line) for HN3 measured
this work. HN3 concentrations were determined manometrically in in this work. Data from ref 18 are given at selected wavelengths
a 1 L calibrated volume. Trace amounts of CO2 are evident in the (open circles). The action spectrum for HN3 near the earth’s surface
spectrum near 2350 cm-1. is shown in arbitrary units as the dashed line.
trometer (Bomem DA3) operating in the infrared. The optics
Control experiments showed that photolysis of HN3 was
were adjusted to allow sixteen traverses of the cell, thus
negligible on the time scale of a typical kinetics run.
providing an observational path length of 32.6 m. Infrared
Similar methodologies were employed to determine the
spectra were obtained at a spectral resolution of 1 cm-1 over
Cl• rate coefficient, with Cl2 photolysis as the Cl• source and
the range 800-3900 cm-1 from the coaddition of 150-200
both acetone and methyl chloride as the reference com-
interferograms, which required 2.5-3.5 min acquisition time.
pounds. These experiments were conducted in 700-710 Torr
Reaction mixtures were photolyzed through a quartz window
synthetic air, with initial concentrations in the chamber as
located at one end of the chamber using a Xe-arc lamp,
follows: [Cl2] ) (3.8-5.8) × 1015 molecule cm-3; [HN3] )
equipped with a Corning 7-54 filter, which provided radiation
(2.1-3.5) × 1014 molecule cm-3; [acetone] ) (4-5) × 1014
in the 250-400 nm range.
molecule cm-3; or [CH3Cl] ) (1.4-1.8) × 1015 molecule cm-3.
For the HO• and Cl• rate coefficient measurements,
Acetone and methyl chloride were monitored primarily near
standard relative rate methodologies were employed (39).
1220 and 1350 cm-1, respectively. For these experiments,
For the HO• studies, ethene was used as the reference
which could be conducted on shorter time scales than the
compound, while HO• generation was from the photolysis of
OH experiments, wall losses were found to be negligibly slow.
methyl nitrite in the presence of added NO:
For studies of the products of the reactions of •N3, reaction
of Cl• with HN3 was used as the •N3 source reaction in most
CH3ONO + hν f CH3O• + NO (6) cases:

CH3O• + O2 f CH2O + HO2• (7) Cl2 + hν f Cl• + Cl• (9)

HO2• + NO f HO• + NO2 (8) Cl• + HN3 f •N3 + HCl (10)

Experiments were carried out in synthetic air at 700-720 In these experiments, mixtures of Cl2 (typically ≈4 × 1015
Torr total pressure, with initial concentrations of species in molecule cm-3) and HN3 (≈3 × 1014 molecule cm-3) were
the chamber as follows: [CH3ONO] ≈ 4 × 1015 molecule cm-3; photolyzed in 1 atm N2 in the presence of one or more
[NO] ≈ 5.5 ×1014 molecule cm-3; [C2H4] ≈ 3.5 × 1014 molecule reactants (NO, NO2, CO, or O2). HO•-initiated oxidations of
cm-3; [HN3] ≈ (2-4) × 1014 molecule cm-3. Quantification HN3 were also carried out, involving the photolysis of
of ethene was done primarily using the strong absorption CH3ONO (≈4 × 1015 molecule cm-3), NO (3-5 × 1014 molecule
feature centered at 950 cm-1. Quantification of HN3 was cm-3), and HN3 (2-4 × 1014 molecule cm-3). Further details
accomplished using the absorption features at 2140 and 1150 regarding reaction conditions for various experiments are
cm-1. A calibrated infrared absorption spectrum for HN3 is provided in the results section of the manuscript.
shown in Figure 1. Chemicals were obtained from the following sources:
The heterogeneous loss of HN3 in the chamber was found glycolic acid (99%, Aldrich), stearic acid (Chem. Service),
to be significant on the time scale of a typical relative rate sodium azide (Sigma), Cl2 (Matheson, UHP), ethene (Linde,
determination, and thus this process had to be corrected for. C.P.), acetone (Sigma-Aldrich, 99.9+%, HPLC grade), methyl
Upon filling the chamber with the mixture just described, chloride (Matheson), CO (Linde, CP grade), NO (Linde), NO2
the decay of HN3 and C2H4 was monitored in the dark for a (from reaction of NO with O2), N2 (boil-off from a liquid N2
period of about 30 min, in the presence of UV light for another Dewar), O2 (U.S. Welding, UHP). Methyl nitrite, CH3ONO,
25-30 min, and then again in the dark for a further 20-30 was synthesized from the dropwise addition of sulfuric acid
min. While C2H4 was found to be stable during the dark to saturated solutions of sodium nitrite in methanol (40) and
periods, HN3 loss occurred with a first-order rate coefficient stored in dry ice between uses. Gases were used as received,
in the range of (1-3) × 10-5 s-1. Some conditioning of the while acetone was degassed by several freeze-pump-thaw
cell with time was evident, as HN3 decays were typically faster cycles before use.
before photolysis than after. For determination of the relative
rate coefficient, HN3 loss during the photolysis period was Results and Discussion
corrected by the average of the loss rate before and after UV Absorption Spectrum of HN3 Our HN3 UV absorption
photolysis. The magnitude of the correction was about 20%. spectrum is plotted in Figure 2, and the data are tabulated

1634 9 ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 39, NO. 6, 2005


TABLE 1: Ultraviolet Absorption Cross Sections for HN3
Measured in This Work
absorption absorption
wavelength cross section wavelength cross section
(nm) (cm2 molecule-1) (nm) (cm2 molecule-1)
215 3.61E-19 291 3.40E-20
217 2.89E-19 293 2.96E-20
219 1.97E-19 295 2.56E-20
221 1.39E-19 297 2.22E-20
223 1.01E-19 299 1.90E-20
225 7.49E-20 301 1.64E-20
227 5.75E-20 303 1.40E-20
229 4.88E-20 305 1.19E-20
231 4.60E-20 307 1.02E-20
233 4.61E-20 309 8.64E-21
235 4.63E-20 311 7.35E-21
237 4.73E-20 313 6.22E-21
239 5.09E-20 315 5.22E-21
241 5.62E-20 317 4.48E-21 FIGURE 3. Relative rate of decay of HN3 (after correction for
243 5.95E-20 319 3.74E-21 heterogeneous loss) versus that of ethene in the presence of HO•.
245 6.12E-20 321 3.12E-21
247 6.35E-20 323 2.60E-21 Some diffuse structure is evident in both our measurement
249 6.93E-20 325 2.17E-21 and that of McDonald et al. (18) in the 240-290 nm region.
251 7.54E-20 327 1.85E-21 As noted by McDonald et al., this structure can be attributed
253 7.65E-20 329 1.53E-21 to two 1600 cm-1 progressions offset from each other by
255 7.73E-20 331 1.26E-21 about 600 cm-1. They assigned the 1600 cm-1 mode to the
257 7.82E-20 333 1.08E-21 ν2 N-N-N asymmetric stretch in the upper electronic state
259 8.01E-20 335 8.72E-22 and the 600 cm-1 frequency to an upper state N-N-N
261 8.59E-20 337 7.43E-22
263 8.69E-20 339 6.57E-22 bending mode (most likely ν6).
265 8.34E-20 341 5.00E-22 HN3 Quantum Yield Determination at 351 nm. Using
267 8.21E-20 343 3.81E-22 the 351 nm absorption cross section measured in this work
269 8.06E-20 345 3.53E-22 (2.03 × 10-22 cm2 molecule-1) and the fractional loss of HN3
271 7.86E-20 347 2.82E-22 upon exposure to excimer irradiation (46% loss for 9 × 104
273 7.78E-20 349 2.35E-22 excimer pulses), the quantum yield for HN3 loss was
275 7.47E-20 351 2.03E-22 determined to be 2.14 ( 0.25 (uncertainties given throughout
277 7.00E-20 353 1.61E-22 this paper are 1σ). N2O, the only measurable photoproduct,
279 6.70E-20 355 1.08E-22
had an appearance quantum yield of 0.86 ( 0.15. These
281 6.34E-20 357 9.02E-23
283 5.74E-20 359 9.34E-23 observations are consistent with a near-unity (1.07 ( 0.15)
285 5.07E-20 361 5.27E-23 quantum yield for the primary photodissociation process
287 4.45E-20 363 4.76E-23 yielding HN•, coupled with subsequent loss of a second HN3
289 3.89E-20 365 3.57E-23 via reaction with OH:

HN3 + hνfHN• + N2 (2a)


in 2 nm intervals in Table 1. Uncertainties are estimated to
be (7% near the maximum of the spectrum (dominated by
the uncertainty in the measurement of the HN3 partial HN• + O2 f HO• + NO (11)
pressure, with lesser contributions from uncertainty in

temperature and path length). Uncertainties increase at HO + HN3 f •N3 + H2O (1)
longer wavelength (e.g., to about (20% near 360 nm) due to
increasing uncertainty in the absorbance measurement. The •
N3 + NO f N2 + N2O (3a)
spectrum shows a broad local maximum near 262 nm, and
the onset of a stronger maximum at wavelengths shorter
net: 2HN3 + hν + O2 f 2N2 + N2O + H2O
than 215 nm. Measurable absorption extends beyond 360
nm. The spectrum of HN3 has previously been measured
The near-quantitative production of N2O during photolysis
quantitatively by McDonald et al. (18) over the range 100-
(i.e., one N2O produced for every two HN3 molecules
325 nm. Their data (estimated from their Figure 3) are shown
consumed) is strong evidence for the near-quantitative
at a few wavelengths in Figure 2. However, measurements
conversion of HN radicals into NO, since R3a appears to be
by Okabe (115-210 nm) (19) and Rohrer and Stuhl (193 nm the only logical source of N2O in the system.
only) (25) suggest that the McDonald et al. data may be low Our spectrum and quantum yield determination can be
by about 15-20%. A similar discrepancy between our data convolved with solar flux data to calculate a theoretical
and those of McDonald et al. is also evident in Figure 2. The tropospheric photolysis rate constant for HN3 (jHN3, s-1). For
scaling of the McDonald et al. (18) data suggested by Rohrer these calculations, diurnally averaged sea-level surface solar
and Stuhl (25) (by the ratio of the two 193 nm cross sections) flux data corresponding to a 40 °N, mid-summer day were
provides a cross section at 248.5 nm (6.8 × 10-20 cm2 used (41). The retrieved action spectrum (dashed line, Figure
molecule-1) that is within 2% of our data point at this 2) shows that maximum atmospheric photolysis is centered
wavelength. Thus, a combination of the Okabe (19) data in near 320 nm. Integration under this curve, assuming a
the vacuum UV (which agrees with the Rohrer and Stuhl (25) quantum yield of unity for the entire actinic region, yields
data at 193 nm) with our longer wavelength data appears to an approximate 1/jHN3 lifetime of 2-3 days.
provide an accurate representation of the HN3 absorption Preliminary experiments to directly determine the HN3
spectrum over the range 110-360 nm. solar photolysis rate were carried out at the Denver Federal

VOL. 39, NO. 6, 2005 / ENVIRONMENTAL SCIENCE & TECHNOLOGY 9 1635


Center. In these experiments, HN3/air mixtures (contained
in a Teflon bag) were exposed to solar radiation over the
course of 2 days, and the HN3 temporal profile was monitored
via FTIR spectroscopy. Although HN3 decay was clearly
observed during sunlight hours, these experiments are not
considered quantitative at this time due to the presence of
patchy cloud cover and the potential for HN3 loss via
unwanted processes (i.e, heterogeneous loss, reaction with
OH). Further experiments of this type, to provide a quantita-
tive comparison with the laboratory cross section and
quantum yield data, are planned.
Rate Coefficient for Reaction of HO• with HN3. The rate
coefficient for reaction of HO• with HN3 was determined
relative to the well-established rate coefficient for reaction
of HO• with C2H4 (8.2 × 10-12 cm3 molecule-1 s-1, 298 K, 1
atm total pressure (42, 43)).

HO• + HN3 f •N3 + H2O (1)


FIGURE 4. Relative rate of decay of HN3 versus those of acetone
(solid circles) and methyl chloride (open circles) in the presence
HO• + CH2dCH2 + M f HOCH2CH2 + M (12) of Cl-atoms.
Relative rate data, after correction for heterogeneous loss k10 of (0.97 ( 0.20) × 10-12 cm3 molecule-1 s-1 and (1.07 (
of HN3 as described earlier, are displayed in Figure 3 and 0.20) × 10-12 cm3 molecule-1 s-1, from which a final value
yield a rate coefficient ratio k1/k12 ) 0.48 ( 0.05. The k10 ) (1.02 ( 0.20) × 10-12 cm3 molecule-1 s-1 can be obtained.
magnitude of the heterogeneous correction was on the order There are a number of determinations of k10 in the literature
of 20%. Incorporating a 10% uncertainty in k12, a value of k1 (5-8), including one temperature-dependent study (8), all
) (3.9 ( 0.8) × 10-12 cm3 molecule-1 s-1 is determined. We obtained using flow tube methodologies. While there is very
also note that other species potentially reactive with HN3 are reasonable agreement between the various room-temper-
generated in these experiments, namely O-atoms (via pho- ature measurements, which range from (9-13) × 10-13 cm3
tolysis of NO2) and O3 (via recombination of O with O2). Both molecule-1 s-1, these measurements were of an indirect
of these species react slowly with HN3 (4, 35) and will not nature, involved substantial occurrence of, or correction for,
contribute significantly to its loss. secondary reactions, and/or required modeling of fairly
Hack and Jordan (3) reported the only other measurement complex reaction systems, and thus uncertainties on the order
of k1. They monitored the decay of HO• via pulsed laser- of (25% typically apply. Our measurement of k10 via a
induced fluorescence, following the flash photolysis of completely different methodology, one that should be free
mixtures of H2O2 and HN3 in He buffer gas. Their value, (1.3 of any complication from secondary reactions, provides
( 0.2) × 10-12 cm3 molecule-1 s-1, is considerably smaller confirming evidence for a value of k10 of (1.0 ( 0.2) × 10-12
than ours; reasons for this discrepancy are not obvious. cm3 molecule-1 s-1.
Similarities in the reactivity of HN3 and HBr might be As discussed above, similarities in reactivity between HN3
expected, given the near identical H-X bond strengths (369 and HBr might be expected. As in the case of the reaction
kJ mol-1) in the two species (33, 42, 43) and the reasonable of HO• with these two species, Cl• reacts more slowly with
correlation between HO• rate coefficients and H-X bond HN3 than with HBr (5, 35), in this case by about a factor of
strengths in the hydrogen halides (and pseudohalides). This 5.
holds true in a qualitative way, with k1 being about a factor Reactions of the •N3 Radical. A series of experiments was
of 2 lower than the rate coefficient for reaction of HO• with carried out to determine the products of the reactions of •N3
HBr (42). with various species (itself, NO, NO2, CO, and O2) and to
Using our value for k1 and a diurnally averaged tropo- determine semiquantitatively the rate coefficients or at least
spheric HO• concentration of 1 × 106 molecule cm-3, the the relative rates for these reactions. For convenience, results
lifetime for HN3 with respect to HO• reaction can be estimated obtained in our work and in previous studies are summarized
to be ≈2-3 days. in Table 2.
Rate Coefficient for Reaction of Cl• with HN3. Relative The simplest experiments involved the photolysis of Cl2
rate data for R10 versus the two reference reactions, R13 and (≈4 × 1015 molecule cm-3)/HN3 (≈2 × 1014 molecule cm-3)
R14, are shown in Figure 4. mixtures in the presence of 700 Torr N2. Although HN3 was
efficiently destroyed in these experiments, the only product
Cl• + HN3 f •N3 + HCl (10) observed in the infrared was N2O with a molar yield of only
(3.1 ( 0.4)% (Figure 5, open circles). The small N2O yield
Cl• + CH3C(O)CH3 f CH3C(O)CH2 + HCl (13) observed possibly arises from the presence of small NOx
impurities in the chamber (reaction of NO2 or NO with •N3
Cl• + CH3Cl f CH2Cl + HCl (14) would eventually lead to N2O, see below). The lack of large
yields of observable products suggests the involvement of
•N self-reaction as an important removal process for N ,
3 3
Least-squares analysis of these data yield the following
resulting in the formation of N2, which is undetectable with
rate coefficient ratios, k10/k13 ) 0.44 ( 0.06 and k10/k14 ) 2.19
our apparatus:
( 0.25. Although there is some discrepancy in the literature
regarding the value of k13 (values range from about (1.8-3.1) •
× 10-12 cm3 molecule-1 s-1), recent data (39, 44, 45) are N3 + •N3 f 3N2 (15)
centered at (2.2 ( 0.4) × 10-12 cm3 molecule-1 s-1. The value
of k14 ) (4.9 ( 0.8) × 10-13 cm3 molecule-1 s-1 seems to be Although the rate coefficient for R15 has not been firmly
well established (42, 43). Combining these reference rate established, it appears to be e 2 × 10-12 cm3 molecule-1 s-1
coefficient data with our measured ratios yields values for (36). Box model simulations of the experiments (conducted

1636 9 ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 39, NO. 6, 2005


TABLE 2: Summary of Available Data Regarding Rate Coefficients and Products for Reactions of N3 Radicals with Atmospherically
Relevant Species
reaction
reactant number rate coefficient (cm3 molecule-1 s-1) products
NO R3 2.9 × 10-12 (ref 36) N2O + N2 (≈100%), this work
NO2 R4 1.9 × 10-12 (ref 36); less than k3/3 (this work) 2NO + N2 (major, this work); 2N2O (<30%, this work);
N2 + N2O + O (<60%, this work); 2N2 + O2
(minor/negligible, this work)
CO R5 1.8 × 10-12 (ref 36); less than k3/100 (this work) NCO + N2 (speculative, ref 36)
O2 R25 <1 × 10-13 (ref 36); < 6 × 10-20 (this work) N2O + NO, N2 + NO2, and/or N2 + NO + O (speculative)

Losses of NO were also measured. In N2 buffer, the loss


of NO was found to be identical to the N2O production (Figure
5, open squares), again consistent with the quantitative
occurrence of R10 and R3a. When air was used as the buffer
gas, losses of NO exceeded those of N2O, and NO2 production
was also observed. This extra conversion of NO to NO2 likely
results from the occurrence of the well-known (35) three-
body reaction, R19:

NO + NO + O2 f NO2 + NO2 (19)

Reasonable correspondence was obtained between the


loss of NO in air, corrected by the amount of NO2 formed,
and the observed N2O. Hewett and Setser (36) determined
a rate coefficient of k3 ) (2.9 ( 0.3) × 10-12 cm3 molecule-1
s-1. They surmised that N2 and N2O were the most likely
products of the reaction, a result now confirmed by the
FIGURE 5. Product yields observed from the Cl-atom-initiated experiments conducted herein.
oxidation of HN3, plotted as a function of HN3 consumption. Filled Another of the reactions studied by Hewett and Setser
circles, N2O produced in experiments with NO added; open squares, (36) was the reaction of •N3 with CO, for which they reported
loss of NO in experiments with NO added; filled triangles, N2O a rate coefficient of (1.8 ( 0.2) × 10-12 cm3 molecule-1 s-1
produced in experiments with NO2 added; open circles, N2O and proposed NCO and N2 as the most reasonable reaction
production in the absence of added NOx. See text for detailed products. When standard Cl2/HN3/N2 reaction mixtures were
experimental conditions. photolyzed in our chamber in the presence of CO, (3.5-7)
× 1014 molecule cm-3, only a very small (≈4%) yield of N2O
using the Acuchem software package (46)) indicate that reac-
was observed. Although HN3 was efficiently consumed, no
tion of Cl• with •N3 cannot be ignored under these conditions:
change in [CO] was observed nor was any other product
detected. These findings are suggestive of a CO-catalyzed
Cl• + •N3 f NCl• + N2 (16) recombination of •N3, via R5a, and R20 and/or R21, resulting
in large yields of undetectable N2:
The fate of NCl• is not known in this system but may involve
formation of NCl3 a compound whose major IR absorption •
N3 + CO f NCO + N2 (5a)
bands (<800 cm-1) likely occur outside our accessible range.
NCO + NCO f CO + CO + N2 (20)
NCl• + Cl2 f NCl2 + Cl• (17)
NCO + •N3 f CO + 2N2 (21)
NCl2 + Cl2 f NCl3 + Cl• (18)
However, we present evidence below that suggests that
Finally, we note that in this case, where no reactive species
R5 may be considerably slower than reported by Hewett and
(NOx, CO) have been added to the chamber, the lifetime of
Setser (36). Thus, in these Cl2/HN3/CO/N2 experiments, it is
N3 may be sufficiently long (perhaps a few seconds, as
possible that NCO is not formed at all, and •N3 loss is
indicated by box modeling studies) to allow for some
controlled by its self-reaction and/or its reaction with Cl•,
heterogeneous removal at the chamber surface.
again leading to large yields of N2 or other undetectable
We conducted numerous experiments with NO, (0.7-
products. The trace yield of N2O in these experiments is again
4.2) × 1014 molecule cm-3, added to the standard Cl2/HN3
likely the result of the presence of minor NOx impurities in
mixtures. Experiments were conducted with either N2 or
the chamber.
synthetic air as the buffer gas at a total pressure of 700-720
To assess the relative rate of the reaction of •N3 with NO
Torr. The major product observed in these experiments was
and CO, mixtures of Cl2 (≈3.5 × 1015 molecule cm-3), HN3
N2O, with a molar yield of 91 ( 12%, independent of the
(≈3 × 1014 molecule cm-3), CO ((1-35) × 1014 molecule cm-3),
nature of the buffer gas (solid circles, Figure 5). This essentially
and NO (≈1 × 1014) were photolyzed in the chamber. The
1:1 conversion of HN3 to N2O, observed under all conditions
initial [NO]:[CO] ratio was varied from 1:1 to 1:25. Given the
in which NO was present, is consistent with the quantitative
similar values for k3 and k5 reported by Hewett and Setser
occurrence of R3a:
(36), one would expect significant production of NCO via
R5a under all of these conditions, and essentially exclusive
Cl• + HN3 f •N3 + HCl (10)
NCO formation at the highest [CO]:[NO] ratios employed.

Reaction of NCO with NO, R22, is thought (47-50) to generate
N3 + NO f N2 + N2O (3a) near-equal yields of N2O and CO2.

VOL. 39, NO. 6, 2005 / ENVIRONMENTAL SCIENCE & TECHNOLOGY 9 1637


NCO + NO f N2O + CO (22a) Second, although NO2 to NO conversion is dominated by
NO2 photolysis in many cases, some experiments suggest a
NCO + NO f CO2 + N2 (22b) large yield of NO must be obtained from R4, and thus a large
branching to channel (4a) and/or (4c) is likely. (Note that
Thus, in our Cl2/HN3/CO/NO experiments, increasing the R4c can lead indirectly to NO via production of O-atoms
ratio of [CO] to [NO] should result in an increase in the yield followed by the occurrence of R24). Third, nitrogen mass
of CO2, via R5a and R22b, and a concomitant decrease in the balance arguments, which do not allow for large yields of N2
yield of N2O. However, in no case was any CO2 production from R4, can be used to preclude a large contribution from
or CO consumption observed in our experiments. Further- channel (4d). Finally, we conclude that the total rate
more, yields of N2O remained high (g85%), indistinguishable coefficient k4 is likely about a factor of 3 or more smaller
from those observed when only NO was added to the system. than k3, given that a unity yield of N2O was obtained in “type
Thus, we find no evidence for the production of NCO in this 2” experiments, even when the initial [NO2]:[NO] ratio was
system, and thus no evidence for the occurrence of a reaction as high as 4:1. Although not a unique solution, all experiments
between •N3 and CO. From the lack of CO2 production in the conducted can be simulated reasonably well with k4a ≈ 6 ×
experiment with the highest [CO]:[NO] ratio, we estimate 10-13 cm3 molecule-1 s-1 and k4b ≈ 2 × 10-13 cm3 molecule-1
that the value for k5 to produce NCO is at least 100 times s-1. Inclusion of a minor occurrence of R4c in the box model
slower than the value of k3. It should be noted that this finding with a compensatory reduction in k4a and k4b also provides
is obtained under the assumption of approximately equal adequate simulations of the experimental results.
rate coefficients for the two branches of R22, as reported in The last reaction studied was that of •N3 with O2, a reaction
refs 47-50. Should higher yields of N2O relative to CO2 be for which Hewett and Setser (36) reported a rate coefficient
the case, the picture would be blurred considerably (note upper limit of 1 × 10-13 cm3 molecule-1 s-1. The reaction
that R5a followed by R22a is indistinguishable from R3a). At appears to have three energetically favorable reaction chan-
the very least then, our data are not consistent with the nels
combined results of refs 36 and 47-50. Given the substantial •
evidence for CO2 production in R22 (47-50), obtained by N3 + O2 f N2O + NO (25a)
various independent investigators using different techniques,
it seems most likely that the value of k5 has been overesti- f N2 + NO2 (25b)
mated in ref 36.
The next reaction studied was that of •N3 with NO2, for f N2 + NO + O (25c)
which Hewett and Setser (36) reported a rate coefficient of
(1.9 ( 0.2) × 10-12 cm3 molecule-1 s-1. As noted in the Our experiments consisted of the photolysis of mixtures
Introduction, multiple reaction channels are possible, leading of Cl2 (≈4 × 1015 molecule cm-3)/HN3 (≈1014 molecule cm-3)
to the production of NO, N2O, N2, O2, and/or O-atoms. We in 750-1050 Torr O2/N2 diluent, with the O2 partial pressure
conducted three types of experiments which can provide varied from 10 to 1000 Torr. The only product observed in
information regarding the rate coefficient and/or mechanism these experiments was N2O. Although its yield did increase
of R4. First, standard Cl2/HN3 mixtures were photolyzed in with O2 partial pressure (from about 8% at 10 Torr O2 to 30%
the presence of (1.7-2.8) × 1014 molecule cm-3 NO2 (“type at 1000 Torr), this increase was less than linear in [O2],
1” experiments). Second, experiments were conducted in suggesting that reaction of •N3 with O2 was not the source of
which both NO and NO2 were added to the standard Cl2/ all of the observed N2O. From box model simulations of the
HN3 reaction mixtures (“type 2”). Finally, experiments were high [O2] experiments, under the assumption that all of the
conducted in which HO• was used to initiate the oxidation N2O formed derived from R25, an upper limit of 6 × 10-20
(“type 3”). While NO2 is not initially present in these “type cm3 molecule-1 s-1 was found for k25. This upper limit is not
3” experiments, it is generated in the HO• source chemistry dependent on the mechanism for R25, as each of the three
(see R6-8 above) and thus its concentration builds up as an reaction channels listed above gave similar N2O yields in the
experiment progresses. As shown in Figure 5 (solid triangles), model simulations. The derived upper limit is however
N2O was clearly a major product in the “type 1” experiments, somewhat dependent on the value used in the model for k15,
although its appearance profile is curved and its initial yield the rate coefficient for self-reaction of •N3 (faster removal of
•N via R15 requires a larger value for k
is clearly less than unity. The production of NO was also 3 25 to generate the
noted in these experiments, which could arise from the observed N2O). While Hewett and Setser have reported an
photolysis of NO2 or from R4. upper limit for k15 of 1.5 × 10-12 cm3 molecule-1 s-1, the
upper limit we report for k25 still applies for values of k15 as
NO2 + hν f NO + O• (23) high as 10-11 cm3 molecule-1 s-1.
Atmospheric Fate of HN3. Due to its low Henry’s law
constant and moderate pKa, aqueous azide is easily trans-
O• + NO2 f NO + O2 (24) ferred to the gas phase as hydrazoic acid (2). The data
obtained in this study show that removal of HN3 from the
In both the “type 2” and “type 3” experiments (i.e., when NO lower troposphere will be controlled by photolysis and by
is present in the initial mixture), the yield of N2O was near reaction with HO• with an overall lifetime of about 1-2 days.
unity, and its appearance profiles were found to be linear. This implies that a hydrazoic acid plume could be sufficiently
Because of the multiple pathways for conversion of NO2 long-lived to be advected many kilometers downwind from
to NO and the multiple potential sources of N2O, i.e., R3, a spill, assuming that other more rapid sinks are not available.
R4b, and R4c, a quantitative assessment of the rate coefficient Solar photolysis of HN3 and reaction with HO• will lead
and branching ratios for R4 is not possible. However, through to the production of the HN• and •N3 radicals, respectively.
box model simulations of the experiments, some conclusions Although these radicals are likely to have little further
can be drawn. First, the less than unity yield of N2O early in atmospheric impact, it is interesting to consider their ultimate
the “type 1” experiments clearly precludes channel (4b) from fate, and thus that of the parent HN3. The HN• radical has
being a major channel and precludes a dominant contribu- been shown to react rapidly with O2 to generate HO• and NO.
tion from R4c. Simulations suggest an N2O yield from R4 of (35) The kinetic database for •N3 is, however, not sufficiently
about 40 ( 20%si.e., a branching ratio of no more than 30% mature to allow a confident assessment. The work conducted
to (4b), or 60% via (4c), or some combination of the two. herein, in conjunction with rate data from Hewett and Setser

1638 9 ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 39, NO. 6, 2005


(36), shows that reactions of •N3 with NO or NO2 (leading in (18) McDonald, J. R.; Rabalais, J. W.; McGlynn, S. P. Electronic spectra
large part to N2O) may contribute. However, we are as yet of the azide ion, hydrazoic acid, and azido molecules. J. Chem.
unable to rule out reaction of •N3 with either CO or O2 as Phys. 1970, 52, 1332-1340.
being of atmospheric relevance. Our work indicates that k5 (19) Okabe, H. Photodissociation of HN3 in the vacuum-ultraviolet,
production and reactivity of electronically excited NH*. J. Chem.
may be considerably slower than measured by Hewett and Phys. 1968, 49, 2726-2733.
Setser, but given the typical atmospheric predominance of (20) Paur, R. J.; Bair, E. J. The isothermal flash photolysis of hydrazoic
[CO] over [NOx], R5 may still contribute. Even our upper acid. Int. J. Chem. Kinet. 1976, 8, 139-152.
limit for k25 leads to a lifetime for •N3 with respect to O2 (21) Welge, K. H. Formation of N2 (A3Σu+) and N(2D,2P) by photo-
reaction of a few seconds, a time scale that is comparable dissociation of HN3 and N2O and their reactions with NO and
to •N3 removal by a few ppbv of NOx. Given the existence of N2O*. J. Chem. Phys. 1966, 45, 166-170.
some discrepancies between our work and that of Hewett (22) Welge, K. H. Formation of NH (c1Π) and NH (A3Πi) in the
and Setser (36), further direct (time-resolved) kinetic studies vacuum-UV photolysis of HN3. J. Chem. Phys. 1966, 45, 4373-
of the •N3 radical, perhaps using tunable diode laser 4374.
techniques for product detection, would be productive. (23) McDonald, J. R.; Miller, R. G.; Baronavski, A. V. Photofragmen-
tation energy distributions and reaction rates of NH from
photodissociation of HN3 at 266 nm. Chem. Phys. Lett. 1977, 51,
Acknowledgments 57-60.
The National Center for Atmospheric Research is operated (24) Kenner, R. D.; Rohrer, F.; Stuhl, F. Generation of NH(a1∆) in the
by the University Corporation for Atmospheric Research 193 nm photolysis of ammonia. J. Chem. Phys. 1987, 86, 2036-
under the sponsorship of the National Science Foundation. 2043.
The authors are grateful to Richard Ross and Jim Hoban (25) Rohrer, F.; Stuhl, F. The 193 (and 248) nm photolysis of HN3:
Formation and internal energy distributions of the HN(a1∆, b1Σ+,
(U.S. EPA NEIC) for their assistance with the solar photolysis A3Π, and c1Π) states. J. Chem. Phys. 1988, 88, 4788-4799.
experiments and to J.-F. Lamarque and David Hanson of (26) Gericke, K.-H.; Theinl, R.; Comes, F. J. Photofragment energy
NCAR for their critical reading of the manuscript. One of us, distribution and rotational anisotropy from excitation of HN3
E.A.B., would like to thank to Ms. Diana A. Love, Director, at 266 nm. Chem. Phys. Lett. 1989, 164, 605-611.
U.S. EPA NEIC, for providing financial and technical support (27) Gericke, K.-H.; Theinl, R.; Comes, F. J. Vector correlation in the
that made possible his sabbatical leave from the University photofragmentation of HN3. J. Chem. Phys. 1990, 92, 6548-
of Arizona. 6555.
(28) Gericke, K.-H.; Lock, M.; Comes, F. J. Photodissociation of HN3:
Direct formation of hydrogen atoms. Chem. Phys. Lett. 1991,
Literature Cited 186, 427-430.
(1) Betterton, E. A. Environmental fate of sodium azide derived (29) Gericke, K.-H.; Lock, M.; Fasold, R.; Comes, F. J. Influence of
from automobile airbags. Crit. Rev. Environ. Sci. Technol. 2003, the electronic asymmetry in NH (1∆) state Λ doublets on the
33, 423-458. photodissociation dynamics of HN3 and DN3. J. Chem. Phys.
(2) Betterton, E. A.; Robinson, J. L. Henry’s Law coefficient for 1992, 96, 422-432.
hydrazoic acid. J. Air Waste Manage. Assoc. 1997, 47, 1216- (30) Bohn, B.; Stuhl, F. NH/ND (a1∆, v′′) vibrational distributions in
1219. the UV photolyses of HN3/DN3 and HNCO/DNCO. J. Phys. Chem.
(3) Hack, W.; Jordan, R. Reaction rates of the reactions of OH(X2Π) 1993, 97, 4891-4898.
with NH(a1∆) and HN3(X). Chem. Phys. Lett. 1999, 306, 111- (31) Haas, T.; Gericke, K.-H.; Maul, C.; Comes, F. J. Photodissociation
116. dynamics of HN3. The N3 fragment internal energy distribution.
(4) Ongstad, A. P.; Coombe, R. D.; Neumann, D. K.; Stech, D. J. Chem. Phys. Lett. 1993, 202, 108-118.
Photochemistry of O3/HN3 mixtures. J. Phys. Chem. 1989, 93, (32) Lock, M.; Gericke, K.-H.; Comes, F. J. Photodissociation dynamics
549-552. of HN3 (DN3) + hν f H(D) + N3. Chem Phys. 1996, 213, 385-
(5) Manke, G. C., II; Setser, D. W. Measuring gas-phase chlorine 396.
atom concentrations: Rate constants for Cl + HN3, CF3I, and
(33) Cook, P. A.; Langford, S. R.; Ashfold, M. N. R. The ultraviolet
C2F5I. J. Phys. Chem. A 1998, 102, 153-159.
photodissociation of HN3: The H + N3 product channel. Phys.
(6) Manke, G. C., II; Henshaw, T. L.; Madden, T. J.; Hager, G. D.
Chem. Chem. Phys. 1999, 1, 45-55.
Temperature dependence of the Cl + HN3 reaction from 300
to 480 K. Chem. Phys. Lett. 1999, 310, 111-120. (34) Barnes, R. J.; Gross, A.; Lock, M.; Sinha, A. Using vector
(7) Yamasaki, K.; Fueno, T.; Kajimoto, O. Studies on the reaction correlation to probe the influence of vibrational state selection
N + N3 f N2 (B3Πg) + N2(X1Σg). Chem. Phys. Lett. 1983, 94, on the photodissociation dynamics of HN3. J. Phys. Chem. A
425-429. 1997, 101, 6133-6137.
(8) Pritt, A. T., Jr.; Coombe, R. D. Azide mechanisms for the (35) NIST Chemical Kinetics Database, http://kinetics.nist.gov.
production of NCl metastables. Int. J. Chem. Kinet. 1980, 12, (36) Hewett, K. B.; Setser, D. W. Chemical kinetics of the azide
741-753. radical: Rate constants for reactions with Cl, NO, NO2, O2, CO,
(9) Shapira, D.; Treinin, A. Photolysis of hydrazoic acid in aqueous CO2, Cl2, and C3H6. J. Phys. Chem. A 1998, 102, 6274-6281.
solution. J. Phys. Chem. 1973, 77, 1195-1198. (37) Staffelbach, T. A.; Orlando, J. J.; Tyndall, G. S.; Calvert, J. G. The
(10) Reiser, A.; Wagner, H. M. In Chemistry of the Azido Group; Patai, UV-visible absorption spectrum and photolysis quantum yields
S., Ed.; Interscience: London, 1971; pp 441-501. of methylglyoxal. J. Geophys. Res. 1995, 100, 14189-14198.
(11) Beckman, A. O.; Dickinson, R. G. The products of the photo- (38) Orlando, J. J.; Tyndall, G. S. Gas-phase UV absorption spectra
chemical decomposition of hydrogen azide. J. Am. Chem. Soc. for peracetic acid, and for acetic acid monomer and dimer.
1928, 50, 1870-1875. J. Photochem. Photobiol. A: Photochem. 2003, 157, 161-
(12) Beckman, A. O.; Dickinson, R. G. The quantum yield in the 166.
photochemical decomposition of hydrogen azide. J. Am. Chem.
(39) Orlando, J. J.; Tyndall, G. S.; Vereecken, L.; Peeters, J. The
Soc. 1930, 52, 124-132.
atmospheric chemistry of the acetonoxy radical. J. Phys. Chem.
(13) Piper, L. G.; Krech, R. H.; Taylor, R. L. The UV photolysis of
A 2000, 104, 11578-11588.
hydrazoic acid. J. Chem. Phys. 1980, 73, 791-800.
(14) Kodama, S. Reactions of NH radicals. 1. Photolysis of HN3 vapor (40) Taylor, W. D.; Allston, T. D.; Moscato, M. J.; Fazekas, G. B.;
at 313 nm. Bull. Chem. Soc. Jpn. 1983, 56, 2348-2354. Kozlowski, R.; Takacs, G. A. Atmospheric photodissociation
(15) Kodama, S. Reactions of NH radicals. 2. Photolysis of HN3 in lifetimes for nitromethane, methyl nitrite, and methyl nitrate.
the presence of C2H6 at 313 nm. Bull. Chem. Soc. Jpn. 1983, 56, Int. J. Chem. Kinet. 1980, 12, 231-240.
2355-2362. (41) McKeen, S. NOAA Aeronomy Laboratory, personal com-
(16) Kodama, S. Reactions of NH radicals. 3. Photolysis of HN3 in munication.
the presence of C2H4 at 313 nm. Bull. Chem. Soc. Jpn. 1983, 56, (42) DeMore, W. B.; Sander, S. P.; Golden, D. M.; Hampson, R. F.;
2363-2370. Kurylo, M. J.; Howard, C. J.; Ravishankara, A. R.; Kolb, C. E.;
(17) Fang, W.-H. Photodissociation of HN3 at 248 nm and longer Molina, M. J. Chemical kinetics and photochemical data for use
wavelength: A CASSCF study. J. Phys. Chem. A 2000, 104, 4045- in stratospheric modeling, Evaluation No. 12, NASA JPL Publ.
4050. 97-4 1997.

VOL. 39, NO. 6, 2005 / ENVIRONMENTAL SCIENCE & TECHNOLOGY 9 1639


(43) Atkinson, R.; Baulch, D. L.; Cox, R. A.; Hampson, R. F.; Kerr, J. (48) Cooper, W. F.; Park, J.; Hershberger, J. F. Product channel
A.; Troe, J. Evaluated kinetic and photochemical data for dynamics of the NCO + NO reaction. J. Phys. Chem. 1993, 97,
atmospheric chemistry: Supplement VI, IUPAC subcommittee 3283-3290.
on gas kinetic data evaluation for atmospheric chemistry. J. (49) Brownsword, R. A.; Hancock, G. Time-resolved FTIR emission
Phys. Chem. Ref. Data 1997, 26, 1326-1499. study of product dynamics in the NO + NCO reaction. J. Chem.
(44) Wallington, T. J.; Andino, J. M.; Ball, J. C.; Japar, S. M. Fourier Soc., Faraday Trans. 1997, 93, 1279-1286.
transform infrared studies of the reaction of Cl atoms wtih PAN,
PPN, CH3OOH, HCOOH, CH3COCH3 and CH3COC2H5 at 295 ( (50) Becker, K. H.; Kurtenbach, R.; Wiesen, P. Investigation of N2O
2 K. J. Atmos. Chem. 1990, 10, 301-313. formation in the NCO + NO reaction by Fourier transform
(45) Christensen, L. K.; Ball, J. C.; Wallington, T. J. Atmospheric infrared spectroscopy. Chem. Phys. Lett. 1992, 198, 424-
oxidation mechanism of methyl acetate. J. Phys. Chem. A 2000, 428.
104, 345-351.
(46) Braun, W.; Herron, J. T.; Kahaner, D. K. Acuchem: A computer
program for modeling complex chemical reaction systems. Int. Received for review November 19, 2004. Accepted November
J. Chem. Kinet. 1988, 20, 51-62. 23, 2004.
(47) Wategaonkar, S.; Setser, D. W. The F + HNCO reaction system:
A flow reactor source of NCO (X2Π) and NF (X3Σ-). J. Phys.
Chem. 1993, 97, 10028-10034. ES048178Z

1640 9 ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 39, NO. 6, 2005

Das könnte Ihnen auch gefallen