Download as pdf or txt
Download as pdf or txt
You are on page 1of 4

LETTERS

PUBLISHED ONLINE: 18 DECEMBER 2011 | DOI: 10.1038/NGEO1353

Formation and fate of oxidized mercury in the upper troposphere and lower stratosphere
Seth N. Lyman * and Daniel A. Jaffe
Mercury contamination affects many aquatic ecosystems1 . The atmosphere is the main transport route for this toxicant2 . According to aircraft measurements, the upper troposphere and lower stratosphere are depleted in gaseous elemental mercury3,4 but enriched in oxidized, particle-bound mercury5,6 . It is therefore assumed that mercury is oxidized in the stratosphere, and then incorporated into stratospheric aerosols6 . However, direct evidence for mercury oxidation in the stratosphere is missing. Here, we present simultaneous measurements of elemental and oxidized mercury concentrations in air of stratospheric origin, collected during an aircraft campaign over North America and Europe in 2010. We show that levels of oxidized mercury are strongly correlated with tracers of stratospheric air. Concentrations of total and elemental mercury, in contrast, are negatively correlated with these tracers. Together, the ndings indicate that elemental mercury is oxidized in stratospheric air masses. We develop a numerical model of atmospheric mercury, based on the assumption that mercury is oxidized in the upper troposphere and lower stratosphere. The resultant vertical proleswhich depict a rapid decline in mercury concentrations with increasing stratospheric heightresemble those seen in other studies, and indicate that mercury has a relatively short stratospheric lifetime. We suggest that following oxidation, mercury is removed from the stratosphere by sedimentation and entrainment processes common to all stratospheric particles. Although previous studies have measured atmospheric mercury by aircraft, our instrument is the first to differentiate between elemental and oxidized forms at a high time resolution. The National Center for Atmospheric Research (NCAR) C-130 has a typical cruising altitude of 6,0007,000 m and is usually well below the tropopause. However, during the OctoberNovember 2010 Western Airborne Mercury Observations campaign aboard the NCAR C-130, we intercepted stratosphere-influenced air on a few occasions. The strongest of these events was during a flight from Bangor, Maine, to Broomfield, Colorado, on 5 November (Fig. 1). During this flight, oxidized mercury (Hg(ii); gaseous + particle-bound) was positively correlated with stratospheric tracers (ozone and potential vorticity (PV)) and negatively correlated with CO (R2 = 0.91), indicating that Hg(ii) increased with increasing stratospheric influence. The Global Forecast System model confirms that the flight altitude was at or near the tropopause during these events (see Supplementary Information). We observed a negative linear correlation between gaseous elemental mercury (Hg(0)) and Hg(ii) in stratosphere-influenced air (Fig. 2), providing evidence that Hg(0) is converted to Hg(ii) in the stratosphere3,4,6 . The slope for this relationship was 0.530.08 (95% confidence interval; R2 = 0.89; that is an Hg(0) loss of 1 pg m3 was associated with an Hg(ii) gain of 0.53 pg m3 ),
a
10
0 5 10 Potential vorticity (PVU)

Altitude (km)

8 3
0 300 600 Hg(II) (pg m3)

6 1 4

1,800

250

THg and Hg(II) (pg m3)

1,350

THg

200

Ozone (ppb)

150 900 100 450 Ozone Hg(II) 50

0 17.00

18.00

19.00 20.00 UTC hour on 5 November 2010 80 85 90 Longitude ( W)

21.00

75

95

100

Figure 1 | Mercury and ozone during a 5 November 2011 ight through stratosphere-inuenced air. a, Flight track, coloured by Hg(II) concentration, along 40 to 44 latitude. The background is coloured by potential vorticity. Solid red lines indicate stratosphere-inuenced air (PV > 1.5 PVU; Events 1 and 2) and the solid green line indicates high-Hg(II) air that was not recently inuenced by the stratosphere (Event 3). b, Concentrations of THg, Hg(II) and ozone. The x axes correspond to a and b. The data gap from 18:00 to 19:00 exists owing to diplomatic restrictions on sampling in Canadian airspace.

greater than the slope of approximately 1 that would be expected if no loss mechanism existed for Hg(II) (ref. 7). An air mass with elevated Hg(ii) that did not show recent stratospheric influence (O3 < 80 ppb, PV < 1PVU; green in Figs 1a and 2) was encountered on the same flight and had a more negative slope of 0.93 0.25 (R2 = 0.62). These two slopes were significantly different (p = 0.003).

University of Washington, Bothell, 18115 Campus Way NE, Bothell, Washington 98011, USA. *e-mail: slyman@uw.edu.
NATURE GEOSCIENCE | ADVANCE ONLINE PUBLICATION | www.nature.com/naturegeoscience
2011 Macmillan Publishers Limited. All rights reserved.

LETTERS
600
20

NATURE GEOSCIENCE DOI: 10.1038/NGEO1353

Ozone

Geopotential height (km)

Stratosphere-influenced y = 0.53x + 771 Hg(II) (pg m3)

Tropopause 15

300

THg Hg(II) 10 Hg(0)

Upper troposphere y = 0.93x + 1,109 0 300


5

600

900 Hg(0) (pg m3)

1,200

500

1,000

1,500

Hg (pg m3) or ozone (ppb)

Figure 2 | Hg(0) versus Hg(II) on 5 November 2011. Air with PV > 1.5 PVU is labelled as stratosphere-inuenced (Events 1 and 2 from Fig. 1), and air with ozone <80 ppb and PV < 1 PVU is labelled as being from the upper troposphere (Event 3 from Fig. 1). The colour scheme corresponds with that of Fig. 1.

Figure 3 | Modelled vertical prole of mercury forms. The ozone vertical prole and average thermal tropopause height are from Wallops Island, Virginia, ozonesonde data. THg is calculated from the observed inverse relationship with ozone. Hg(II) concentrations are calculated from the observed Hg(II)/THg ratio in stratosphere-inuenced air. Hg(0) was calculated as the difference between THg and Hg(II).

The Hg(0)Hg(ii) slope in stratosphere-influenced air was probably greater than 1 because total atmospheric mercury (THg) decreased as the Hg(ii)/Hg(0) ratio increased with increasing stratospheric influence (see Fig. 1). In other words, THg decreases with stratospheric depth, and the proportion of THg that is Hg(ii) increases. In contrast, THg does not decrease with height in the troposphere8 . Hg(0)Hg(ii) slopes close to 1 have been reported for air with upper troposphere/lower stratosphere influence in other studies7 , and Hg(ii) in these cases may have been formed within the troposphere, not in the stratosphere. On the other hand, a slope much greater than 1 in the troposphere9 could be due to Hg(ii) loss processes occurring within the middle or lower troposphere, and is probably not a robust indicator of stratospheric influence. We developed a simple model of the vertical distribution of mercury forms in the upper troposphere and lower stratosphere based on our measurements (Fig. 3). We obtained ozonesonde data for October and November 2010 from Wallops Island, Virginia, and used these to determine an average ozone vertical profile. We then used the linear relationships of ozone with THg (2.59 pg m3 ppb1 O3 (ppb) + 1,556 pg m3 ; R2 = 0.78) and the unitless Hg(ii)/THg ratio (0.003 ppb1 O3 (ppb) 0.15; R2 = 0.94) in stratosphere-influenced air to calculate THg and Hg(ii) concentrations, and calculated Hg(0) as the difference between these two (see Supplementary Information for regression confidence intervals). This model implicitly assumes that THg is mostly Hg(0) in the lower and middle troposphere, and that Hg(0) is oxidized to Hg(ii) with increasing height in the stratosphere. The decrease in THg with height implies a loss mechanism for stratospheric mercury, which is discussed below. Relationships with ozone were used because ozone is a convenient tracer for stratospheric air, and the model is not meant to imply a causal relationship between ozone and the distribution of mercury forms. Although it is unlikely that this model is quantitatively representative of the stratosphere as a whole, it does exhibit the qualitative features of mercury distribution in the stratosphere that have been demonstrated in this and other studies. The Hg(0)Hg(ii) slope in the model (0.52 0.08; 95% confidence interval; R2 = 0.99) is not different from the measured slope
2

for stratosphere-influenced air (p = 0.85). Moreover, the vertical profile of Hg(ii) in this model is qualitatively similar to the profile for particle-bound mercury measured previously6 , which showed a maximum in particle-bound mercury just above the tropopause and a decrease with further depth into the stratosphere. Measurements of Hg(ii) gas-particle partitioning indicate that most Hg(ii) probably exists in the particulate phase for the temperatures encountered during Events 1 and 2 (20 to 30 C; ref. 10). Whereas the model shows a total depletion of mercury about 5 km above the tropopause, previous measurements6 show some mercury even 8 km above the tropopause. This discrepancy probably exists because the linear relationships used in the model become less applicable deep in the stratosphere (note that the ozonesonde data used in the model exhibited higher ozone mixing ratios at the thermal tropopause than is typical; see Supplementary Information). We compared our data against mercury measurements from the CARIBIC aircraft4 and from the INTEX-B (ref. 3) and ARCTAS (ref. 11) aircraft campaigns (Fig. 4). All of these data include extended flight periods in stratosphere-influenced air. CARIBIC measurements are for the same months as our data whereas INTEX-B and ARCTAS data are from 2006 and 2008, respectively. Data with CO > 100 ppb and data below detection were excluded (see Supplementary Information), and data were binned by ozone mixing ratio. Although the CARIBIC, INTEXB and ARCTAS mercury instruments probably captured Hg(0) efficiently, the extent to which they captured Hg(ii) is unclear. The fraction of Hg(ii) captured may have been different for CARIBIC versus INTEX-B and ARCTAS owing to differences in system configuration (see Supplementary Information). In Fig. 4, CARIBIC, INTEX-B and ARCTAS observations binned by ozone concentration are most similar to our observations of Hg(0) (rather than THg). In spite of the temporal and spatial differences among these measurements, they all show a similar trend of decreasing mercury with increasing stratospheric influence. INTEX-B data were 32% lower than Hg(0) measurements from this study (average difference for binned values in Fig. 4). An in-flight intercomparison during INTEX-B showed that INTEX-B mercury measurements were 3040% lower than an aircraft that compared

NATURE GEOSCIENCE | ADVANCE ONLINE PUBLICATION | www.nature.com/naturegeoscience


2011 Macmillan Publishers Limited. All rights reserved.

NATURE GEOSCIENCE DOI: 10.1038/NGEO1353


1,500

LETTERS
the lifetime of lower stratospheric particles can be as short as a few weeks18 . As stratospheric Hg(ii) is concentrated near the tropopause, it can be expected to have a relatively short lifetime. Oxidation of carbonyl sulphide (OCS) is a major source of stratospheric sulphate aerosol20,21 . OCS does not exhibit a strong vertical gradient in the troposphere, but it decreases with height in the stratosphere22 . Sulphate particles produced through OCS oxidation slowly descend owing to sedimentation, and stratospheric sulphate concentrations, like OCS mixing ratios, decrease with height21 . Thus, OCS provides an analogue for Hg(0). Hg(0) is depleted as Hg(ii) is produced, and particulate Hg(ii) is continually removed from the stratosphere by sedimentation and other processes, limiting the stratospheric concentration of THg. Hg(0) is probably depleted much lower in the stratosphere relative to OCS because it has a shorter lifetime. The lifetime of stratospheric OCS is 65 years22 , whereas the lifetime of Hg(0) in the stratosphere is thought to be <1yr (ref. 13). Evidence exists that precipitation associated with deep convective clouds may be enriched in Hg(ii) from the upper troposphere/lower stratosphere23 . Furthermore, a number of studies have reported that high Hg(ii) is associated with tracers of upper troposphere/lower stratosphere air7,24 . Concurrent measurements of ozone and the 10 Be/7 Be ratio indicate that only 115% of surface ozone is of stratospheric origin, with the rest being produced in the troposphere25,26 . Similar measurements may elucidate the contribution of upper troposphere/lower stratosphere Hg(ii) to surface Hg(ii) concentrations. The results of this study, coupled with the findings of others, indicate that the stratosphere is depleted in THg and enriched in Hg(ii). Our model of the vertical profile of mercury forms in the upper troposphere and lower stratosphere matches previous results and amounts to a prediction that can be tested in future campaigns. The rapid decrease in mercury with increasing depth in the stratosphere implies a relatively short lifetime for stratospheric Hg(ii). Thus, the stratosphere is probably a source of Hg(ii) to the troposphere, either by entrainment of stratospheric air or through scavenging by deep convective clouds. However, we expect this source to be small when compared with Hg(ii) production in the troposphere. Using a downward mass flux from the stratosphere of 3.8 1017 kg yr1 (ref. 27) and a tropopause Hg(ii) concentration of 600 pg m3 , we calculate that the stratosphere contributes 176 MgHg(II) yr1 to the troposphere, compared with an estimated global Hg(ii) net deposition rate to the surface of 4,300 Mg yr1 (ref. 13).

THg (this study) 1,000 ARCTAS CARIBIC INTEX-B 500 Hg(0) (this study)

Hg (pg m3)

100

200 Ozone (ppb)

300

400

Figure 4 | Mercury versus ozone measured in this study and during other aircraft campaigns. CARIBIC data are from October and November 2010, and INTEX-B and ARCTAS data are from 2006 and 2008, respectively. Data are binned by ozone mixing ratio in 50 ppb increments (050, 50100 and so on). The squares indicate means, and the error bars show 95% condence intervals. Only data with CO < 100 ppb are shown. Data below detection and CARIBIC data >3,000 pg m3 (<1% of all data) are not shown.

well against ground-based measurements12 , and in a mercury modelling study INTEX-B measurements were >40% lower than model results, whereas the model matched measurements from ARCTAS and CARIBIC reasonably well13 . ARCTAS measurements were only 6% lower than Hg(0) in this study, and the mercuryozone slopes of these data sets were not significantly different from each other (unbinned data; P = 0.52). The 6% difference between the two studies could easily be due to a real difference in mercury concentrations, especially because the two studies were carried out more than two years apart and in different locations. CARIBIC measurements were not significantly different from Hg(0) in this study, except for the 200250 ppb ozone bin (Fig. 4). The CARIBIC data follow a slope that is qualitatively similar to the ARCTAS slope and the Hg(0) slope from this study for lower ozone values, but not at high ozone values. One explanation for this could be that CARIBIC measurements include some fraction of Hg(ii), especially if the fraction of Hg(ii) measured increases with increasing stratospheric influence. The Hg(ii) maximum near the tropopause probably exists because Hg(ii) production becomes limited by the decreasing Hg(0) concentration with increasing stratospheric depth (Fig. 3), and because of removal of Hg(ii) from the stratosphere. It has been shown14 that the Hg(ii) maximum is concurrent with a maximum in particle-bound bromine, providing evidence that reactive bromine compounds may be involved in the oxidation process. Reactive bromine is abundant in the stratosphere15 and oxidizes Hg(0) efficiently16 . However, it has been shown13 that reactive bromine chemistry alone cannot account for the observed steep decline in Hg(0) above the tropopause, and it was reasoned that Cl, Cl2 and/or BrCl may also play a role. The decrease in THg with increasing stratospheric depth provides strong evidence that a loss mechanism exists for stratospheric Hg(ii). As stratospheric Hg(ii) is probably particle bound, loss mechanisms should be the same as those for stratospheric particles in general. These include gravitational sedimentation, mass exchange and scavenging by clouds17,18 . The average lifetime for particles in the stratosphere is about 1 year17 , but particles injected deep in the stratosphere by large volcanic eruptions can persist for several years19 , and under some conditions

Methods
Mercury instrumentation. Atmospheric mercury was quantified with the University of Washington Dual-channel Oxidized Hg System (UW DOhGS). The DOhGS measured Hg(0), THg and Hg(ii) in simultaneous 2.5 min intervals. The system pulled outside air through a heated (>50 C) rear-facing polytetrafluoroethylene (PTFE) inlet and a heated (110 C) perfluoroalkoxy (PFA) 1.0-cm-inner-diameter sample line. A diaphragm pump pulled a constant volumetric flow of 20 l min1 through the sample line, and two Tekran 2537B mercury vapour analysers each pulled 1 standard litre per minute (0 C and 1 atm) from the sample line. The 50% particle cut size for the inlet was 0.9 m (see Supplementary Information). One 2537B analyser pulled sample air through a 650 C quartz pyrolyser tube packed with quartz wool. The heated area of the pyrolyser was 2.2 cm diameter 17.2 cm long. The pyrolyser thermally reduced gas-phase Hg(ii) to Hg(0) with 95% efficiency, reduced particle-bound mercury to Hg(0) and passed existing Hg(0) intact (see Supplementary Information). Thus, all mercury in the sample line was captured by this 2537B analyser, giving a measurement of THg. The other 2537B analyser pulled sample air through a 1.0-cm-inner-diameter 25-cm-long PFA tube packed with 3 g quartz wool. Quartz wool was pre-blanked by heating at 500 C for 2 h. The downstream ends of quartz wool traps were plugged with 2.5-mm-thick PTFE frit material with a 30 m pore size. The system employed up to four traps and switched between them sequentially using PTFE valves to maintain a good collection efficiency. The traps were air cooled, and new traps were used for each flight. They retained gas-phase Hg(ii) with 97% efficiency and retained particle-bound mercury, allowing only Hg(0) to pass through to the
3

NATURE GEOSCIENCE | ADVANCE ONLINE PUBLICATION | www.nature.com/naturegeoscience


2011 Macmillan Publishers Limited. All rights reserved.

LETTERS
analyser (see Supplementary Information) . Thus, this 2537B analyser measured Hg(0). The difference between Hg(0) and THg is assumed to equal to the amount of Hg(ii) retained on the quartz wool. All mercury data are reported in units of picograms per standard cubic metre, with standard conditions of 0 C and 1 atm. For Hg(0), 100 pg m3 = 11.2 ppqv . Detection limits (2.5 min) were calculated as two times the standard deviation of measured values of THg, Hg(0) and Hg(ii) in mercury-free air, and were 49 14 (mean standard deviation of detection limit determined on different days), 69 38 and 68 20 pg m3 , respectively. Hg(0) calibration checks carried out in ambient air in flight and on the ground showed 98 10% recovery for the THg analyser and 958% recovery for the Hg(0) analyser (n = 65). More information about the UW DOhGS is available in the Supplementary Information, including information about particle collection and measurement efficiency. Other measurements. Ozone was measured using a fast-response chemiluminescence instrument28 . Carbon monoxide was measured with an Aero-Laser Model 5002. Aerosol size distributions were measured with an FSSP-300 (0.320 m size range; ref. 29). Meteorological and state measurements aboard the C-130 are described at http://www.eol.ucar.edu/raf/Bulletins/. Potential vorticity was derived from Global Forecast System 0.5 degree global analysis data (http://dss.ucar.edu/datasets/ds335.0/docs/) using Unidatas Integrated Data Viewer. CARIBIC mercury measurements are from flights FrankfurtCaracasFrankfurtOsakaFrankfurt on 2022 October, FrankfurtJohannesburg on 14 November and FrankfurtBogota on 16 November 2010. INTEX-B and ARCTAS data were taken from http://www.espo.nasa.gov/intex-b and http://www.espo.nasa.gov/arctas, respectively. Ozonesonde data for Wallops Island, Virginia, were obtained from http://uairp.wff.nasa.gov/browse/dir/ozone/wallops/index.html.

NATURE GEOSCIENCE DOI: 10.1038/NGEO1353


13. Holmes, C. D. et al. Global atmospheric model for mercury including oxidation by bromine atoms. Atmos. Chem. Phys. 10, 1203712057 (2010). 14. Murphy, D. & Thomson, D. Halogen ions and NO in the mass spectra of aerosols in the upper troposphere and lower stratosphere. Geophys. Res. Lett. 27, 32173220 (2000). 15. Dorf, M. et al. Bromine in the tropical troposphere and stratosphere as derived from balloon-borne BrO observations. Atmos. Chem. Phys. 8, 72657271 (2008). 16. Skov, H. et al. Fate of elemental mercury in the Arctic during atmospheric mercury depletion episodes and the load of atmospheric mercury to the Arctic. Environ. Sci. Technol. 38, 23732382 (2004). 17. Rasch, P. J. et al. An overview of geoengineering of climate using stratospheric sulphate aerosols. Phil. Trans. R. Soc. A 366, 40074037 (2008). 18. Menzies, R. T. & Tratt, D. M. Evidence of seasonally dependent stratospheretroposphere exchange and purging of lower stratospheric aerosol from a multiyear lidar data set. J. Geophys. Res. 100, 31393148 (1995). 19. Zhao, J., Turco, R. P. & Toon, O. B. A model simulation of Pinatubo volcanic aerosols in the stratosphere. J. Geophys. Res. 100, 73157328 (1995). 20. Notholt, J. et al. Enhanced upper tropical tropospheric COS: Impact on the stratospheric aerosol layer. Science 300, 307310 (2003). 21. Wilson, J. C. et al. Steady-state aerosol distributions in the extra-tropical, lower stratosphere and the processes that maintain them. Atmos. Chem. Phys. 8, 66176626 (2008). 22. Barkley, M. P., Palmer, P. I., Boone, C. D., Bernath, P. F. & Suntharalingam, P. Global distributions of carbonyl sulfide in the upper troposphere and stratosphere. Geophys. Res. Lett. 35, L14810 (2008). 23. Selin, N. E. & Jacob, D. J. Seasonal and spatial patterns of mercury wet deposition in the United States: Constraints on the contribution from North American anthropogenic sources. Atmos. Environ. 42, 51935204 (2008). 24. Weiss-Penzias, P., Gustin, M. S. & Lyman, S. N. Observations of speciated atmospheric mercury at three sites in Nevada: Evidence for a free tropospheric source of reactive gaseous mercury. J. Geophys. Res. 114, D14302 (2009). 25. Dibb, J. E. et al. Estimation of stratospheric input to the Arctic troposphere: 7 Be and 10 Be in aerosols at Alert, Canada. J. Geophys. Res. 99, 1285512864 (1994). 26. Zanis, P. et al. An estimate of the impact of stratosphere-to-troposphere transport (STT) on the lower free tropospheric ozone over the Alps using 10 Be and 7 Be measurements. J. Geophys. Res. 108, 85208529 (2003). 27. Seo, K. & Bowman, K. P. Lagrangian estimate of global stratospheretroposphere mass exchange. J. Geophys. Res. 107, 45554563 (2002). 28. Ridley, B. A., Grahek, F. E. & Walega, J. G. A small high-sensitivity, medium-response ozone detector suitable for measurements from light aircraft. J. Atmos. Oceanic Technol. 9, 142148 (1992). 29. Gass, S. & Hegg, D. A. Comparison of columnar aerosol optical properties measured by the MODIS airborne simulator with in situ measurements: A case study. Remote Sens. Environ. 66, 138152 (1998).

Received 19 July 2011; accepted 15 November 2011; published online 18 December 2011

References
1. Selin, N. E. Global biogeochemical cycling of mercury: A review. Annu. Rev. Environ. Res. 34, 4363 (2009). 2. Fitzgerald, W. F., Engstrom, D. R., Mason, R. P. & Nater, E. A. The case for atmospheric mercury contamination in remote areas. Environ. Sci. Technol. 32, 17 (1998). 3. Talbot, R., Mao, H., Scheuer, E., Dibb, J. & Avery, M. Total depletion of Hg in the upper tropospherelower stratosphere. Geophys. Res. Lett. 34, L23804 (2007). 4. Slemr, F. et al. Gaseous mercury distribution in the upper troposphere and lower stratosphere observed onboard the CARIBIC passenger aircraft. Atmos. Chem. Phys. 9, 19571969 (2009). 5. Murphy, D., Thomson, D. & Mahoney, M. In situ measurements of organics, meteoritic material, mercury, and other elements in aerosols at 5 to 19 km. Science 282, 16641669 (1998). 6. Murphy, D., Hudson, P., Thomson, D., Sheridan, P. & Wilson, J. Observations of mercury-containing aerosols. Environ. Sci. Technol. 40, 31633167 (2006). 7. Swartzendruber, P. C. et al. Observations of reactive gaseous mercury in the free troposphere at the Mt. Bachelor Observatory. J. Geophys. Res. 111, D24301 (2006). 8. Swartzendruber, P. C., Jaffe, D. A. & Finley, B. Development and first results of an aircraft-based, high time resolution technique for gaseous elemental and reactive (oxidized) gaseous mercury. Environ. Sci. Technol. 43, 74847489 (2009). 9. Fain, X., Obrist, D., Hallar, A. G., McCubbin, I. & Rahn, T. High levels of reactive gaseous mercury observed at a high elevation research laboratory in the Rocky Mountains. Atmos. Chem. Phys. 9, 80498060 (2009). 10. Rutter, A. P. & Schauer, J. J. The effect of temperature on the gas-particle partitioning of reactive mercury in atmospheric aerosols. Atmos. Environ. 41, 86478657 (2007). 11. Mao, H. et al. Arctic mercury depletion and its quantitative link with halogens. J. Atmos. Chem. 65, 126 (2010). 12. Swartzendruber, P. C. et al. Vertical distribution of mercury, CO, ozone, and aerosol scattering coefficient in the Pacific Northwest during the spring 2006 INTEX-B campaign. J. Geophys. Res. 113, D10305 (2008).

Acknowledgements
This study was financially supported by the US National Science Foundation. We thank the CARIBIC team (especially F. Slemr and R. Ebinghaus) for providing us with CARIBIC data (www.caribic-atmospheric.com). We thank R. Talbot at the University of Houston for access to data from INTEX-B and ARCTAS. We are grateful for the efforts of numerous technicians, engineers and scientists at NCARs Research Aviation Facility who were essential to the success of this work.

Author contributions
S.N.L. participated in study design, collected and analysed data and is the primary author of the manuscript. D.A.J. led the study and contributed substantially to all aspects of it, including manuscript preparation.

Additional information
The authors declare no competing financial interests. Supplementary information accompanies this paper on www.nature.com/naturegeoscience. Reprints and permissions information is available online at http://www.nature.com/reprints. Correspondence and requests for materials should be addressed to S.N.L.

NATURE GEOSCIENCE | ADVANCE ONLINE PUBLICATION | www.nature.com/naturegeoscience


2011 Macmillan Publishers Limited. All rights reserved.

You might also like