Interpreting the Isotopic Composition of Sulfate and...

1
Overview Atmospheric oxidant concentrations impact the lifetime of reduced trace gases (CH4, SO2, NOx, VOCs) with implications for climate forcing and air pollutants, but there is little direct evidence of paleo-oxidant concentrations. The oxygen isotopic composition 17 O = δ 17 O - 0.52×δ 18 O) of sulfate and nitrate reflect their formation pathways in the atmosphere, and measurements from ice cores in conjunction with atmospheric chemistry modeling have the potential to constrain past oxidant concentrations. Objective The first continuous record of Δ 17 O of both sulfate and nitrate has been constructed for the past 2300 years from the WAIS-Divide ice core. Features of the record include: Increase in Δ 17 O(SO4 2- ) and decrease in Δ 17 O(NO3 - ) between ~1700 CE and present. Coincident decreases in Δ 17 O(SO4 2- ), Δ 17 O(NO3 - ), and δ 15 N at ~350 C.E. that exceeds 200 BCE - 1600 CE variability. The goal of this work is to interpret this new ice core record using sulfate and nitrate box models constrained by GEOS-Chem. Δ 17 O of Sulfate Δ 17 O(SO4 2- ) depends on: Oxidant concentrations Cloud liquid water content Cloud water pH Δ 17 O of Nitrate Δ 17 O(NO3 - ) depends on: Oxidant concentrations Aerosol surface area DMS concentration Post-depositional processing Combined measurements and modeling of both sulfate and nitrate Δ 17 O better constrains oxidant changes than if only one were considered, as all non-oxidant factors impact only sulfate or nitrate, but not both. SO 2 SO 3 2- SO 2 ·H 2 O HSO 3 - SO 4 2- aqueous gas heterogeneous O 3 (ss, dust) OH O 3 O 2 (Fe, Mn) H 2 O 2 NO RO2 (R3) HO2 (R2) O3 (R1) hX(R4) OH (R5) O 3 (R6) DMS, HC (R7) NO2 PAN decomposition + other NO2 sources Stratospheric denitrification Emission HNO3 NO3 NO 2 , H 2 O (R8) Acknowledgments This work was supported by NSF grants NSF-OPP 0538049 and NSF-AGS 0704169. The authors appreciate the support of the WAIS Divide Science Coordination Office at the Desert Research Institute of Reno Nevada for the collection and distribution of the WAIS Divide ice core and related tasks (Kendrick Taylor, NSF Grants 0440817 and 0230396). The National Science Foundation Office of Polar Programs also funds the Ice Drilling Program Office and Ice Drilling Design and Operations group for coring activities; Raytheon Polar Services for logistics support in Antarctica; and the 109th New York Air National Guard for airlift in Antarctica. The National Ice Core Laboratory, which curated the core and preformed core processing, is jointly funded by the National Science Foundation and the United States Geological Survey. WAIS-Divide Ice Core Sulfate Δ 17 O, Nitrate Δ 17 O, and Nitrate δ 15 N Preliminary Interpretation •Long-term: Changing post-depositional processing (due to changing snow accumulation) can explain the trend in nitrate Δ 17 O and δ 15 N. 1700 CE-Present: Ice core indicators of non-oxidant factors (LWC, pH, aerosol surface area, DMS, post-depositional processing) do not show changes consistent with the changes in the Δ 17 O of sulfate and nitrate. The changes in the Δ 17 O of sulfate and nitrate may reflect changes in the oxidant concentrations in the different source regions of sulfate and nitrate. Model-Measurement Comparison Comparison of GEOS-Chem to measurements of present-day sulfate and nitrate Δ 17 O from aerosol, cloud, and firn samples. Preindustrial Oxidants and Δ 17 O(SO4 2- ) GEOS-Chem modeled changes in annual mean tropospheric (a) O3, (b) OH, and (c) H2O2 concentrations and the corresponding change in (d) Δ 17 O(SO 4 2-) between the preindustrial and present. In the Arctic, increased metal-catalyzed oxidation and decreased cloud pH render Δ 17 O(SO 4 2-) insensitive to changing oxidant concentrations. In the Antarctic, Δ 17 O(SO 4 2-) is sensitive to oxidants, but there is little change in Δ 17 O(SO 4 2-) because the relative changes in oxidant concentrations have opposing isotopic effects [Sofen et al., 2011]. GEOS-Chem does not reproduce the observed ~1‰ 1600 CE-present increase in Δ 17 O(SO 4 2-) using standard preindustrial assumptions. Box Model Monte Carlo Simulations Box models of sulfate and nitrate formation will be run with a range of conditions representative of present-day and preindustrial conditions based on GEOS-Chem simulations or relevant literature. Δ 17 O(SO4 2- ) and Δ 17 O(NO3 - ) are calculated and compared to the ice core record to assess which sets of conditions (e.g. oxidants, LWC, pH, DMS, aerosol surface area) are consistent with the observations. 2Aerosol SA, [DMS] Nitrate Box Model 4 ) O(NO 3 - ) = References Alexander, B.A., M.G. Hastings, D.J. Allman, J. Dachs, J.A. Thornton, S.A. Kunasek (2009), Quanifying atmospheric nitrate formation pathways based on a global model of the oxygen isotopic composition (Δ 17 O) of atmospheric nitrate. Atmos. Chem. Phys., 9, 5043-5056, doi:10.5194/acp-9-5043-2009. Kunasek, S.A., B. Alexander, E.J. Steig, M.G. Hastings, D.J. Gleason, J.C. Jarvis (2008), Measurements and modeling of Δ 17 O of nitrate in snowpits from Summit, Greenland. J. Geophys. Res., 113, D24302, doi: 10.1029/2008JD010103. Marlon, J.R., P.J. Bartlein, C. Carcaillet, D.G. Gavin, S.P. Harrison, P.E. Higuera, F. Joos, M.J. Power, I.C. Prentice (2008), Climate and human influences on global biomass burning over the past two millennia. Nat. Geosci., doi: 10.1038/ngeo313. Sofen, E.D., B. Alexander, S.A. Kunasek (2011), The impact of anthropogenic emissions on atmospheric sulfate production pathways, oxidants, and ice core Δ 17 O(SO4 2– ). Atmos. Chem Phys., 11, 3565-3578, doi:10.5194/acp-11-3565-2011. Interpreting the Isotopic Composition of Sulfate and Nitrate from an Antarctic Ice Core Using GEOS-Chem Eric Sofen 1 , Becky Alexander 1 , Eric Steig 2 , Mark Thiemens 3 , Shelley Kunasek 2,4 , Andrew Schauer 2 , Meredith Hastings 1,5 , Janet Bautista 2 , Terri Jackson 3 , Laura Vogel 6 , Helen Amos 1,7 1 Dept. of Atmospheric Sciences, University of Washington; 2 Dept. of Earth and Space Sciences, UW; 3 Dept. of Chemistry and Biochemistry, UCSD; 5 Now at Brown U.; 4 Now at Highline Community College; 6 JISAO, UW, Now at Columbia U.; 7 Now at Harvard U. SO2 + OH → SO4 Δ 17 OSO4 = 0.0 ‰ S(IV) + H2O2 → SO4 Δ 17 OSO4 = 0.65‰ S(IV) + O3 → SO4 Δ 17 OSO4 = 8.75‰ S(IV) + O2 → SO4 Δ 17 OSO4 = -0.09‰ [Alexander et al., 2009] Sulfate [Sofen et al., 2011] Nitrate [Alexander et al., 2009] ! " #$# % &'( )* ! " #$# % &'( + +, + +, + +, -+ + +, + +, + +, -+ . / +,01 2 +3 / + [Helen Amos] Nitrate • Sources: Biomass burning, lightning, soil NOx emissions; stratospheric denitrification. • Nitrate forms in NOx emission regions. • Increased NOx increases OH locally. Decreases Δ 17 O(NO3 - ). Alternative: Increased PAN transport imparts local summer Δ 17 O(NO3 - ). Sulfate • Source: Southern Ocean DMS emissions. • Sulfate forms in remote ocean areas. • OH in remote regions has decreased since the preindustrial due to increased CH4 concentrations. Increases Δ 17 O(SO4 2- ).

Transcript of Interpreting the Isotopic Composition of Sulfate and...

OverviewAtmospheric oxidant concentrations impact the lifetime of reduced trace gases (CH4, SO2, NOx, VOCs) with implications for climate forcing and air pollutants, but there is little direct evidence of paleo-oxidant concentrations. The oxygen isotopic composition (Δ17O = δ17O - 0.52×δ18O) of sulfate and nitrate reflect their formation pathways in the atmosphere, and measurements from ice cores in conjunction with atmospheric chemistry modeling have the potential to constrain past oxidant concentrations.

ObjectiveThe first continuous record of Δ17O of both sulfate and nitrate has been constructed for the past 2300 years from the WAIS-Divide ice core. Features of the record include:• Increase in Δ17O(SO4

2-) and decrease in Δ17O(NO3-) between

~1700 CE and present.• Coincident decreases in Δ17O(SO4

2-), Δ17O(NO3-), and δ15N at

~350 C.E. that exceeds 200 BCE - 1600 CE variability. The goal of this work is to interpret this new ice core record using sulfate and nitrate box models constrained by GEOS-Chem.

Δ17O of Sulfate Δ17O(SO4

2-) depends on:•Oxidant concentrations•Cloud liquid water content•Cloud water pH

Δ17O of Nitrate Δ17O(NO3

-) depends on:•Oxidant concentrations•Aerosol surface area•DMS concentration•Post-depositional processing

Combined measurements and modeling of both sulfate and nitrate Δ17O better constrains oxidant changes than if only one were considered, as all non-oxidant factors impact

only sulfate or nitrate, but not both.

SO2

SO32-

SO2·H2O

HSO3- SO4

2-aqueous

gas

heterogeneous

O 3 (s

s, dust)

OH

O3

O2

(Fe, Mn)

H2O2

5044 B. Alexander et al.: Model nitrate �17O

NO

RO2 (R3)

HO2 (R2)

O3 (R1)

h

(R4)

OH (R5)

O3 (R6)

DMS, HC (R7)

NO2

PAN decomposition

+ other NO2 sources

Stratospheric denitrification

Emission

HNO3

NO3NO2, H2O (R8)

Fig. 1. Simplified chemistry leading to inorganic nitrate formationin the model.

NO is oxidized to NO2 by ozone (O3) (Reaction R1) and per-oxy radicals (HO2 and RO2) (Reactions R2–R3). NO2 is alsoproduced through decomposition or oxidation of peroxyacylnitrates (PANs), reactions of peroxy radicals (RO2) with per-oxynitrates (RO2NO2) and alkyl nitrates (RONO2), and ox-idation or decomposition of nitrous acid (HONO), pernitricacid (HNO4), nitrate radicals (NO3) and dinitrogen pentox-ide (N2O5). NO2 is lost from the atmosphere through ox-idation and dry deposition to the surface. NO2 is oxidizedby OH (daytime) to form HNO3 (Reaction R5) or by O3 toform NO3 (Reaction R6). NO3 is rapidly photolyzed duringthe daytime, so that NO3 concentrations are only significantat night. At night, NO3 reacts with dimethylsulfide (DMS)or hydrocarbons (HC) (Reaction R7) or with NO2 to formN2O5 followed by hydrolysis on the surface of aerosols (Re-action R8) to form HNO3. The lifetime of NOx against con-version to HNO3 in the boundary layer varies from ∼1 day(tropics and in summer in the mid-to-high latitudes) to ∼3days (high latitudes in winter) (Levy et al., 1999). Nitrate islost from the atmosphere mainly through wet and dry depo-sition.The oxygen isotopic composition of nitrate reflects the rel-

ative importance of different oxidants in NOx cycling andnitrate formation (Michalski et al., 2003; Hastings et al.,2003). Atmospheric nitrate exhibits an anomalous (“mass-independent”) oxygen isotopic composition. The isotopiccomposition of atmospheric nitrate is considered anomalousdue to enrichment in 17O relative to 18O over the expected re-lationship (δ17O≈0.5×δ18O) that results from purely mass-dependent fractionation processes (Matsuhisa et al., 1978).Isotope ratios are expressed in “delta notation”,

δxO =�Ox/O16

�sample�

Ox/O16�standard

− 1 (1)

where x=17 or 18 and the standard used for oxygen iso-topic analysis is Standard Mean Ocean Water (SMOW).Generally, δ values are expressed in per mil (‰). Theoxygen isotopic anomaly in nitrate, quantified here as�17O=δ17O−0.52×δ18O, results mainly from the transfer ofan isotopic anomaly in atmospheric ozone during oxidationof NO and NO2 (Reactions R1 and R6 in Fig. 1).

Observations of the mean �17O value of troposphericozone (�17O(O3)) at different locations range from 25–35‰(Johnston and Thiemens, 1997; Krankowsky et al., 1995);the absolute variability is much larger (6–54‰) (Morin et al.,2007). This large range in observed�17O(O3) is unexpectedbased on the pressure and temperature dependence of the iso-topic enrichment measured in laboratory studies (Morton etal., 1990). This discrepancy may be due to a bias (most likelylow) in the observational data from this difficult measure-ment and/or uncertainties in the laboratory data. A photo-chemical equilibrium model constrained with laboratory data(Janssen et al., 1999; Mauersberger et al., 1999) calculates�17O=35‰ for surface ozone (Lyons, 2001) Michalski andBhattacharya (2009) calculated �17O(O3)=33–37‰ using aquadratic fit of data fromMorton et al. (1990) assuming tem-peratures and pressures typical of mid-latitudes. Other oxi-dants (OH, RO2, HO2) involved in NOx cycling and nitrateformation have �17O values at or near zero (Dubey et al.,1997; Savarino and Thiemens, 1999a; Lyons, 2001; Savarinoand Thiemens, 1999b). Observations of the �17O value ofatmospheric nitrate (�17O(nitrate)) range from ∼10–40‰(Kaiser et al., 2007; McCabe et al., 2007; Michalski et al.,2003; Morin et al., 2007, 2008; Savarino et al., 2007; Broth-ers et al., 2008) highlighting the importance of ozone for re-active nitrogen chemistry in the atmosphere.Reactive bromine (BrO) can also play a role in both NOx

cycling and nitrate formation in polar regions (Evans et al.,2003; Saiz-Lopez et al., 2008). BrO participates in NOx cy-cling and nitrate formation through the following reactions:

NO+ BrO → NO2 + Br (R1)

NO2 + BrO+ M → BrONO2 + M (R2)

BrONO2 + H2O(aq) → HOBr+ HNO3 (R3)

BrO is expected to have large �17O values due to the in-volvement of O3 in BrO formation (Morin et al., 2007). Theformation of nitrate through BrO oxidation and BrONO2 hy-drolysis is thus expected to lead to large �17O (nitrate) val-ues on the upper end of the range of observations. The globalimportance of reactive bromine in NOx cycling and nitrateformation remains to be quantified, but is thought to havesignificant regional impacts on both tropospheric NOx con-centrations and ozone production (Yang et al., 2005).Understanding the importance of different oxidants for

NOx cycling and the pathways of nitrate formation is criticalfor understanding the budget of reactive nitrogen in the atmo-sphere. Here we present the first global chemical transportmodel of the oxygen isotopic composition of atmosphericnitrate. Comparison with available observations sheds lighton previous assumptions used in box model studies regard-ing the isotopic composition of ozone and the isotopic trans-fer function during NOx oxidation reactions, and provides ameans to test and validate the model’s representation of re-active nitrogen chemistry.

Atmos. Chem. Phys., 9, 5043–5056, 2009 www.atmos-chem-phys.net/9/5043/2009/

Acknowledgments This work was supported by NSF grants NSF-OPP 0538049 and NSF-AGS 0704169. The authors appreciate the support of the WAIS Divide Science Coordination Office at the Desert Research Institute of Reno Nevada for the collection and distribution of the WAIS Divide ice core and related tasks (Kendrick Taylor, NSF Grants 0440817 and 0230396). The National Science Foundation Office of Polar Programs also funds the Ice Drilling Program Office and Ice Drilling Design and Operations group for coring activities; Raytheon Polar Services for logistics support in Antarctica; and the 109th New York Air National Guard for airlift in Antarctica. The National Ice Core Laboratory, which curated the core and preformed core processing, is jointly

funded by the National Science Foundation and the United States Geological Survey.

WAIS-Divide Ice Core Sulfate Δ17O, Nitrate Δ17O, and Nitrate δ15N

Preliminary Interpretation•Long-term: Changing post-depositional processing (due to changing snow accumulation) can explain the trend in nitrate Δ17O and δ15N.•1700 CE-Present: Ice core indicators of non-oxidant factors (LWC, pH, aerosol surface area, DMS, post-depositional processing) do not show changes consistent with the changes in the Δ17O of sulfate and nitrate. The changes in the Δ17O of sulfate and nitrate may reflect changes in the oxidant concentrations in the different source regions of sulfate and nitrate.

Model-Measurement ComparisonComparison of GEOS-Chem to measurements of present-day sulfate and nitrate Δ17O from aerosol, cloud, and firn samples.

Preindustrial Oxidants and Δ17O(SO42-)GEOS-Chem modeled changes in annual mean tropospheric (a) O3, (b) OH, and (c) H2O2 concentrations and the corresponding change in (d) Δ17O(SO4

2-) between the preindustrial and present. In the Arctic, increased metal-catalyzed oxidation and decreased cloud pH render Δ17O(SO4

2-) insensitive to changing oxidant concentrations. In the Antarctic, Δ17O(SO4

2-) is sensitive to oxidants, but there is little change in Δ17O(SO4

2-) because the relative changes in oxidant concentrations have opposing isotopic effects [Sofen et al., 2011]. GEOS-Chem does not reproduce the observed ~1‰ 1600 CE-present increase in Δ17O(SO4

2-) using standard preindustrial assumptions.

Box Model Monte Carlo SimulationsBox models of sulfate and nitrate formation will be run with a range of conditions representative of present-day and preindustrial conditions based on GEOS-Chem simulations or relevant literature. Δ17O(SO4

2-) and Δ17O(NO3-) are calculated and

compared to the ice core record to assess which sets of conditions (e.g. oxidants, LWC, pH, DMS, aerosol surface area) are consistent with the observations.

Oxidants[SO2], pH,

LWC...Aerosol SA,[DMS]

Nitrate Box Model

!

!"#

$

$"#

%

%"#

&'()*+,-!./-012

$3

%3

43

56+7*+,-

!./-012

8$33 3 $33 433 933 :33 !333 !$33 !433 !933 !:33 $3333

#

!3

!#

56+7*+,-

!#5-012

;,*7-0<"="2>6?,@A*(,@-BC<39D8!E-3#D8!-*FG-39H-0IA<JFF,((83:34!32

Sulfate Box Model

∆17O

(SO4 2-)

∆17O

(NO3 -)

?=

ReferencesAlexander, B.A., M.G. Hastings, D.J. Allman, J. Dachs, J.A. Thornton, S.A. Kunasek (2009), Quanifying atmospheric nitrate

formation pathways based on a global model of the oxygen isotopic composition (Δ17O) of atmospheric nitrate. Atmos. Chem. Phys., 9, 5043-5056, doi:10.5194/acp-9-5043-2009.

Kunasek, S.A., B. Alexander, E.J. Steig, M.G. Hastings, D.J. Gleason, J.C. Jarvis (2008), Measurements and modeling of Δ17O of nitrate in snowpits from Summit, Greenland. J. Geophys. Res., 113, D24302, doi: 10.1029/2008JD010103.

Marlon, J.R., P.J. Bartlein, C. Carcaillet, D.G. Gavin, S.P. Harrison, P.E. Higuera, F. Joos, M.J. Power, I.C. Prentice (2008), Climate and human influences on global biomass burning over the past two millennia. Nat. Geosci., doi: 10.1038/ngeo313.

Sofen, E.D., B. Alexander, S.A. Kunasek (2011), The impact of anthropogenic emissions on atmospheric sulfate production pathways, oxidants, and ice core Δ17O(SO42–). Atmos. Chem Phys., 11, 3565-3578, doi:10.5194/acp-11-3565-2011.

Interpreting the Isotopic Composition of Sulfate and Nitrate from an Antarctic Ice Core Using GEOS-ChemEric Sofen1, Becky Alexander1, Eric Steig2, Mark Thiemens3, Shelley Kunasek2,4, Andrew Schauer2, Meredith Hastings1,5, Janet Bautista2, Terri Jackson3, Laura Vogel6, Helen Amos1,7

1 Dept. of Atmospheric Sciences, University of Washington; 2 Dept. of Earth and Space Sciences, UW; 3 Dept. of Chemistry and Biochemistry, UCSD; 5 Now at Brown U.; 4 Now at Highline Community College; 6 JISAO, UW, Now at Columbia U.; 7 Now at Harvard U.

SO2 + OH → SO4 Δ17OSO4 = 0.0 ‰ S(IV) + H2O2 → SO4 Δ17OSO4 = 0.65‰

S(IV) + O3 → SO4 Δ17OSO4 = 8.75‰S(IV) + O2 → SO4 Δ17OSO4 = -0.09‰

[Alexander et al., 2009]

Sulfate [Sofen et al., 2011] Nitrate [Alexander et al., 2009]

5048 B. Alexander et al.: Model nitrate �17O

Fig. 2. Scatter plot of monthly-mean model calculations at thesurface versus observations. Observations include: Bermuda(black circles), La Jolla (red triangles), South Pole (green tri-angles), DDU (blue diamonds), Chile (purple diamond), Prince-ton (purple open circles), Summit (orange squares), Alert (redopen squares), Ecuador (purple square), COCA (blue open trian-gles), and an Atlantic cruise (black open squares). Each paneluses different assumptions regarding the bulk isotopic composi-tion of ozone and the isotopic transfer mechanism during ozoneoxidation: (a) �17O(O3)=35‰ and statistical �17O transfer, (b)�17O(O3)=25‰ and statistical�17O transfer (c)�17O(O3)=35‰and Savarino et al. (2008) transfer applied to NO+O3 and NO2+O3,(d) �17O(O3)=25‰ and Savarino et al. (2008) transfer applied toNO+O3 and NO2+O3. The linear least-squares regression (thick),y=x (solid), y=2x and y=0.5x (dashed) lines are shown. Inset is theregression equation and R2 values.

NO2+O3 reactions. Although the comparison in Fig. 2 sug-gests that �17O(O3)=35‰ is the appropriate value to usein calculations of �17O(nitrate), it is no substitute for di-rect observations of �17O(O3), as model biases will impactour results. Model errors in transport may impact for ex-ample the vertical distribution of NOx, which will in turninfluence �17O(nitrate) due to changing A values (gener-ally decreasing with height) or changes in the importance ofdifferent nitrate formation pathways with altitude. Globalbiases in oxidant (O3, NO3, OH, RO2, HO2) concentra-tions or the formation rate of nitrate for specific pathways(e.g. N2O5 hydrolysis) will also impact the isotopic cal-culations. For example, if the true �17O(O3)=25‰, thenour finding that �17O(O3)=35‰ provides the best agree-ment between observed and modeled �17O(nitrate) may in-dicate that the model globally and systematically under-estimates (overestimates) O3 and/or NO3 (OH and/or per-oxy radical) concentrations. The resulting magnitude of thebias in �17O(nitrate) will be dependent upon the magni-tude of the bias in O3, NO3, OH, or peroxy radical con-centrations. Our calculated global mean OH concentra-tion (10.7×105 molecules cm−3) compares well with (Wang

et al., 2008) (10.6×105 molecules cm−3) who optimizedthe global abundance of OH by interpreting observationsof methyl chloroform using an inverse technique and theGEOS-Chem model. This suggests no global, systematicbias in OH. Calculated O3 concentrations in the GEOS-Chem model has been extensively evaluated against obser-vations (Hudman et al., 2009; Zhang et al., 2008; Teraoet al., 2008; Wu et al., 2007; Wang et al., 2009), suggest-ing no systematic global bias. Peroxy radical concentrationswould have to be overestimated globally by a factor of ∼3to account for a 10‰ difference in �17O(NOx). Althoughwe may expect significant regional biases in these radicalconcentrations, a systematic global bias seems unlikely forspecies (or nitrate formation pathways) with a high degree ofspatial variability.We note that Michalski et al. (2003) and Kunasek

et al. (2008) achieved good agreement with observa-tions of �17O(nitrate) using a box models that assumed�17O(O3)=35‰ with the statistical transfer function. Intheir analysis of observed �17O(nitrate) in La Jolla, CA,Michalski et al. (2003) estimated a much larger fraction ofnitrate formed via N2O5 hydrolysis (30–90%) than we do(5–50%), revealing the sensitivity of �17O(nitrate) to thefractional contribution of N2O5 hyrolysis to total nitrate for-mation. We expect that Michalski et al. (2003) significantlyoverestimated the contribution of N2O5 hydrolysis in theirmodel due to the use of a reaction rate constant that did notaccount for variability of factors such as aerosol surface area,chemical composition, humidity and temperature. Also, it isgenerally thought that models overestimate the importanceof nitrate formation via N2O5 hydrolysis through overesti-mates of the reaction probability (Davis et al., 2008; Mc-Neill et al., 2006; Brown et al., 2009). Kunasek et al. (2008)also obtained good agreement with snowpit observations of�17O(nitrate) in Summit, Greenland with a box model thatassumed �17O(O3)=35‰ with the statistical isotopic trans-fer function. Their box model used boundary conditionsfrom the same global model used here (GEOS-Chem), butinherently did not account for transport of nitrate formed atlower latitudes (with lower values of �17O(nitrate)). Thisdemonstrates the importance of accounting for transport ofnitrate from different regions with potentially different val-ues of �17O(nitrate), particularly for remote polar locationsfar from major sources of NOx.

4.2 Global variability of tropospheric nitrate formationpathways and �17O(nitrate)

Figure 3 shows global plots of modeled June-July-Augustand December-January-February mean �17O(nitrate) inthe lowest model layer (0–200m above the surface).�17O(nitrate) ranges from 7–41‰, withAday andAnight gen-erally ranging from 0.2–1.0 and 0.9–1.0, respectively. Thespatial variability of �17O(nitrate) is largely determined bythe importance of O3 in NOx cycling (A) due to the fact that

Atmos. Chem. Phys., 9, 5043–5056, 2009 www.atmos-chem-phys.net/9/5043/2009/

[Helen Amos]

Nitrate• Sources: Biomass burning, lightning, soil NOx

emissions; stratospheric denitrification.• Nitrate forms in NOx emission regions.• Increased NOx increases OH locally.➡Decreases Δ17O(NO3

-).➡Alternative: Increased PAN transport imparts

local summer Δ17O(NO3-).

Sulfate• Source: Southern Ocean DMS emissions.• Sulfate forms in remote ocean areas.• OH in remote regions has decreased since the

preindustrial due to increased CH4 concentrations.

➡ Increases Δ17O(SO42-).