Download Quantifying the contribution to atmospheric iodine from

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Ocean wikipedia , lookup

Ocean acidification wikipedia , lookup

Ecosystem of the North Pacific Subtropical Gyre wikipedia , lookup

Physical oceanography wikipedia , lookup

Critical Depth wikipedia , lookup

Atmospheric convection wikipedia , lookup

Transcript
1
Quantifying the contribution of marine organic gases to atmospheric
2
iodine: Supporting Material - further method details
3
4
Further details of ambient VOIC analysis
5
Both GC/MS instruments incorporated a 50 m x 0.32 mm internal diameter SGE BPX5
6
capillary column, and detection by electron impact ionisation quadrupole mass
7
spectrometry. Calibration of the GC/MS instruments for ambient VOIC analysis was
8
achieved based upon a permeation oven based technique described fully in Wevill and
9
Carpenter [2004]. Permeation tubes containing pure liquid halocarbons (>97%, Aldrich)
10
maintained in a thermostated oven with a constant nitrogen flow rate were used to deliver
11
a constant calibration gas output containing dilute levels of halocarbons. Sequential 10 μl
12
volumes of permeation gas were injected into a stream of nitrogen (BOC, CP grade) in
13
order to generate a linear calibration of each halocarbon at pptv levels. Full permeation
14
oven calibrations were performed at the beginning and end of each cruise, whilst day-to
15
day variability in instrument sensitivity was accounted for using an in-house prepared gas
16
standard, containing pptv-levels of all VOICs studied.
17
During the MAP cruise the air sample inlet was located on the foredeck to the port side of
18
the ship and during RHaMBLe the inlet was sited on the port side. During both cruises
19
ambient air was pumped through a PFA Teflon line (50 m length for the MAP cruise, 20
20
m length for RHaMBLe, both ½ i.d) using a diaphragm pump at a rate of ~30 l min-1,
21
and clean air was sub-sampled diverted upstream of the diaphragm pump via a metal
22
bellows pump (Senior Aerospace Limited) and delivered to a cold trap at -30 °C within
23
the thermal desorption unit to pre-concentrate volatile components prior to GC/MS
24
analysis.
25
26
Sea-air flux calculations
27
Sea-air fluxes were calculated as a function of the concentration difference across the sea-
28
air boundary (ΔC, mol cm-3) and the total gas transfer velocity k (cm h-1, which
29
incorporates both the waterside and airside transfer velocities - kw and ka, respectively),
30
according to Equation (1). We used the Nightingale et al. [2000] parameterization for the
31
waterside transfer velocity, (kw = {0.222u2 + 0.333u}{SC/660}-1/2) and the McGillis et al.
32
[2000] expression to include the airside resistance (k = kw {1 – γa}) (where γa is the
33
airside gradient function), with minute averaged wind speeds u (in m s-1), and the
34
dimensionless Henry’s law coefficients (H) from Moore et al. [1995]. The Khalil et al.
35
[1999] approximation was used to derive the temperature dependent Schmidt numbers
36
(Sc) for each gas (Equation 2), where T is seawater temperature and M is the relative
37
molecular mass. Sea-air fluxes were also calculated according to the Liss and Merlivat
38
[1986] and Wanninkhof [1992]
39
parameterizations resulted in mean average CH3I, CH2ICl and CH2I2 fluxes for both
40
cruises which differed by ±25-30%.
parameterizations for kw, and the three different kw
41
42
F = kΔC
(1)
Sc = 335.6M1/2 (1-0.065T + 0.002043T2 - (2.6x10-5)T3)
(2)
43
44
45
46
47
Further details of the Ocean Mixed Layer Model
48
The ocean mixed layer (OML) model contains explicit parameterisations of the
49
mechanisms which drive mixing within the surface ocean. Mixing in surface waters
50
occurs as a response to two broad categories of forcing: wind-driven forced convection
51
(caused by surface and internal waves, shear instabilities, secondary circulations - e.g.
52
Langmuir cells - and various interactions between them) and density instability-driven
53
free convection. Buoyancy instabilities driving pure free convection may be a result of
54
surface cooling or evaporation. Free convection is usually combined with some degree of
55
forced convection. A model aiming to accurately reproduce the behaviour of physical
56
properties in oceanic surface waters must therefore explicitly describe or parameterise the
57
above mechanisms adequately. The OML model is based on Large et al. [1994] and uses
58
a parameterisation for the depth-resolved eddy diffusivity profile based on criteria
59
consistent with the conservation equations in their primitive form. The model uses a non-
60
local eddy diffusivity (K) profile parameterisation (KPP) after Troen & Mahrt [1986]
61
with local parameterisations for the three mixing processes in the interior added by Large
62
et al. [1994], as well as adaptations of the rules for matching boundary layer properties at
63
the ocean interior, and of the boundary conditions at the surface and bottom.
64
The model was validated for physical property prediction by running with a standard set
65
of initialisations and external forcings, and comparing predicted physical properties such
66
as the mixed layer depth with those obtained by six other turbulent vertical mixing
67
models reported in Martin [1985, 1986] under the same synthetic forcing conditions.
68
Having validated the model skill in reproducing physical properties, it was modified and
69
extended to integrate the vertical production, destruction and transport of dissolved
70
gaseous CH2I2 and CH2ICl.
The depth-dependent biological production terms were
71
equivalent in structure to a typical Chl a profile (based on in-situ measured depth profiles
72
of Chl a for MAP) maximising at the bottom of the predicted mixed layer, and were
73
simply scaled to approximately reproduce the observed CH2ICl and CH2I2 concentrations
74
at 2 or 6 m depth. Maximum production rates (at the Chl a maximum) were ~6 pmol dm-
75
3
76
photolysis rates were calculated using wavelength-resolved solar irradiance from
77
NASA’s
78
(http://snowdog.larc.nasa.gov/jin/rtset.html) combined with absorption cross-sections and
79
quantum yields for photolysis of dihalomethanes in seawater from Jones and Carpenter
80
[2006]. Sea–air exchange losses are parameterised using the wind speed dependent gas
81
transfer velocities used in sea-air flux calculations (described above). Initialising the
82
model with vertical salinity, temperature and velocity profiles (temperature and salinity
83
from in-situ profile measurements during MAP and from archived BODC data from the
84
same region and time of year for RHaMBLe; initial velocity profiles were based on
85
acoustic doppler current profiler (ADCP) measurements from previous studies in similar
86
regions) and forcing it with in-situ ship-board meteorological observations (~10 m above
87
sea level), the physical properties and VOIC vertical profile evolution were predicted.
d-1 and ~1.5 pmol dm-3 d-1 for CH2I2 and CH2ICl, respectively. Depth-dependent
Coupled
Ocean
Atmosphere
Radiative
Transfer
(COART)
model
88
89
Further details of the MISTRA 1D Atmospheric Model
90
The one-dimensional Lagrangian MISTRA model, which contains detailed treatment of
91
the thermodynamics and microphysics and explicit gas and aqueous phase chemical
92
mechanisms, was used to simulate iodine sources to the MBL. The gas-phase chemical
93
mechanism used in the MISTRA model was updated to the latest IUPAC
94
recommendation (June 2006), with a revised DMS mechanism mainly based upon Barnes
95
et al. [2006]. The chemistry of hydrocarbons and halogenated hydrocarbons was revised
96
by including more explicit treatment of the intermediates, based on the Master Chemical
97
Mechanism protocol [Saunders et al., 2003]. Besides chlorine and bromine inorganic
98
chemistry, the model includes a detailed iodine mechanism, similar to the one described
99
in Pechtl et al. [2006], with some minor revisions and updates.
100
The model is initialised with ozone mixing ratios typical of clean sub-tropical MBL air
101
that has travelled for at least 3 days in the boundary layer from the mid-latitude to the
102
sub-tropical Atlantic Ocean (30-40 ppbv, Read et al., [2008]). The model was initially
103
run for 1.5 days to spin-off the meteorology and the microphysics, after which the
104
chemistry was reinitialized and the model run for 3 days.
105
106
Although the total organic iodine flux derived from our measurements is similar to that of
107
the Vogt et al. [1999] base case fluxes used in the model (which were based on coastal
108
studies), the relative abundance of each gas is different, with lower CH2I2 and higher
109
CH3I fluxes from the measurements; the net effect being much slower release of reactive
110
iodine using the measured fluxes.
111
112
Sensitivity tests on MISTRA model
113
Sensitivity tests were carried out on the iodine mechanism in the model, in an attempt to
114
resolve the discrepancy between the modeled and measured concentrations of IO
115
(average measured IO was ~1.4 pptv, compared to the maximum modelled value of ~0.3
116
pptv). Rate coefficients and/or products and branching ratios of selected reactions were
117
altered, including: mass accommodation coefficients of HOI and HIO3, photolysis rates
118
of OIO, I2O2 and CH2I2, products of the decomposition reaction of I2O2, products of the
119
OH+IBr, I+INO3 reactions and of the IO self-reaction, formation of HIO3 and I2O3,
120
photolysis of iodinated oxygenates formed by peroxy radicals self- and cross-reactions
121
and addition of a speculative HIO3 loss term. However, none of these reactions impacted
122
the modelled concentrations of IO by more than a few percent, except when HIO3
123
formation was set to zero or when a loss term for HIO3 (with a pseudo first-order rate
124
coefficient of 1.0x10-3 s-1) was added to the mechanism, but even in this case the model
125
still underestimated the observed IO mixing ratios by up to a factor of 3. Even doubling
126
the measured VOIC fluxes did not provide an iodine source sufficient to produce the
127
observed IO concentrations (doubling the upwelling VOIC fluxes resulted in IO
128
concentrations of ~0.6 pptv).
129
The model was also run with an iodine mechanism similar to that used by Gomez-Martin
130
et al. [2009], which resulted in higher modeled IO (~0.6 pptv compared to ~0.3 pptv with
131
the initial mechanism). Although a direct comparison between the two models is not
132
straightforward, we believe that the main reasons for the differences are (i) the Gomez-
133
Martin et al. [2009] mechanism does not include the IO+CH3O2 reaction, (ii) the rate
134
coefficient of the OIO+OH reaction used by Gomez-Martin et al. [2009] is 6x10-12 cm-3
135
molecule-1 s-1 and the product is HOI, while in our mechanism the rate coefficient is
136
5x10-10 cm-3 molecule-1 s-1 (at 298 K, Plane et al. [2006]) and the product is HIO3, and
137
(iii) mass accommodation coefficients for inorganic iodine species are a factor of 2 (HI)
138
and up to a factor of 50 (HOI) lower than those used in the present work. The differences
139
in the gas-phase mechanism may or may not be significant (the reaction IO+CH3O2 is, in
140
fact, extremely uncertain), although the model is very sensitive to HIO3 formation and
141
loss. The ultimate fate of iodine is accumulation in particles in the form of iodide and
142
iodate [Pechtl et al., 2006] and this typically occurs by uptake of inorganic iodine (mostly
143
INOx, HIO3 and IxOy); therefore all processes that reduce the formation rate of these
144
species and/or their uptake onto particles (such as OIO+OH and the mass accommodation
145
coefficients) will result in higher concentrations of inorganic iodine in the gas-phase.
146
147
148
Supporting references
149
Barnes, I., J. Hjorth and N. Mihalopoulos (2006), Dimethyl sulfide and dimethyl
150
sulfoxide and their oxidation in the atmosphere, Chemical Reviews, 106, 940-975.
151
Gomez-Martin, J. C., S. H. Ashworth, A. S. Mahajan and J. M. C. Plane (2009),
152
Photochemistry of OIO: laboratory study and atmospheric implications, Geophys.
153
Res. Lett., 36, L09802.
154
Jones, C. E. and L. J. Carpenter (2006), Solar photolysis of CH2I2, CH2ICI, and
155
CH2IBr in water, saltwater, and seawater, Environ. Sci. Technol., 40(4), 1372,
156
doi:10.1021/es058022e.
157
Khalil, M. A. K., R. M. Moore, D. B. Harper, J. M. Lobert, D. J. Erickson, V.
158
Koropalov, W. T. Sturges and W. C. Keene (1999), Natural emissions of chlorine-
159
containing gases: Reactive Chlorine Emissions Inventory, J. Geophys. Res. - Atmos.,
160
104(D7), 8333-8346.
161
Large, W. G., J. C. McWilliams and S. C. Doney (1994), Oceanic vertical mixing: a
162
review and a model with a non-local boundary layer parameterization, Rev. of
163
Geophys., 32, 363-403.
164
Liss, P. S. and L. Merlivat (1986), Air-sea gas exchange rates: Introduction and
165
synthesis, in The Role of Air-Sea Exchange in Geochemical Cycling, edited by P.
166
Buat-Menard, pp. 113 – 127 D. Reidel, Norwell, Mass.
167
Martin, P. J. (1985), Simulation of the mixed layer at OWS N and P with several
168
models, J. Geophys. Res., 90, 903-916.
169
Martin, P. J. (1986), Testing and comparison of several mixed-layer models, U. S.
170
Naval Ocean Research and Development Activity Report 143.
171
McGillis, W. R., J. W. H. Dacey, N. M. Frew, E. J. Bock and R. K. Nelson (2000),
172
Water-air flux of dimethylsulfide, J. Geophys. Res., 105(C1), 1187-1193.
173
Moore, R. M., C. E. Geen and V. K. Tait (1995), Determination of Henry Law
174
Constants for a suite of naturally-occurring halogenated methanes in seawater,
175
Chemosphere, 30(6), 1183-1191.
176
Mössinger, J. C., D. E. Shallcross and R. A. Cox (1998), UV-VIS absorption cross-
177
sections and atmospheric lifetimes of CH2Br2, CH2I2 and CH2BrI, J. Chem. Soc.
178
Faraday Trans., 94, 1391-1396.
179
Nightingale, P. D., G. Malin, C. S. Law, A. J. Watson, P. S. Liss, M. I. Liddicoat, J.
180
Boutin and R. C. Upstill-Goddard (2000), In situ evaluation of air-sea gas exchange
181
parameterizations using novel conservative and volatile tracers, Global Biogeochem.
182
Cy., 14(1) 373-387.
183
Pechtl, S., E. R. Lovejoy, J. B. Burkholder and R. von Glasow (2006), Modeling the
184
possible role of iodine oxides in atmospheric new particle formation, Atmos. Chem.
185
Phys., 6, 505-523.
186
Plane, J. M. C., D. M. Joseph, B. J. Allan, S. H. Ashworth and J. S. Francisco (2006),
187
An experimental and theoretical study of the reactions OIO + NO and OH + OH, J.
188
Phys. Chem. A, 110(1), 93-100 doi: 10.1021/jp055364y.
189
Rattigan, O. V., D. E. Shallcross and R. A. Cox (1997), UV absorption cross-sections
190
and atmospheric photolysis rates of CF3I, CH3I, C2H5I and CH2ICl, J. Chem. Soc.
191
Faraday Trans., 93, 2839-2846.
192
Read, K. A., et al. (2008), Extensive halogen-mediated ozone destruction over the
193
tropical Atlantic Ocean, Nature, 453, 1232-1235 doi:10.1038/nature07035.
194
Saunders, S. M., M. E. Jenkin, R. G. Derwent and M. J. Pilling (2003), Protocol for
195
the development of the Master Chemical Mechanism, MCM v3 (Part A): tropospheric
196
degradation of non-aromatic volatile organic compounds, Atmos. Chem. Phys., 3,
197
161-180.
198
Vogt, R., R. Sander, R. von Glasow and P. J. Crutzen (1999), Iodine chemistry and its
199
role in halogen activation and ozone loss in the marine boundary layer: A model
200
study, J. Atmos. Chem., 32(3), 375-395.
201
Troen, I. and L. Mahrt (1986), A simple model of the atmospheric boundary layer -
202
sensitivity to surface evaporation, Boundary-Layer Meteorology, 37, 1-2, 129-148.
203
Wanninkhof, R. (1992), Relationship between wind-speed and gas-exchange over the
204
ocean, J. Geophys. Res. - Oceans, 97(C5), 7373-7382.
205
Wevill, D. J. and L. J. Carpenter (2004), Automated measurement and calibration of
206
reactive volatile halogenated organic compounds in the atmosphere, The Analyst, 129,
207
634-638.
208
209
210
Supporting Material Figure S1. Rate of iodine atom release from VOICs within the
211
atmospheric boundary layer, based upon the open ocean sea-air fluxes (ocean) and
212
upwelling fluxes (upwell). Total VOICs corresponds to iodine atom production from all
213
measured VOICS (CH3I, C2H5I, 1-C3H7I, CH2IBr, CH2ICl and CH2I2).
214
215
216