ArticlePDF Available

Tropospheric ozone change from 1980 to 2010 dominated by equatorward redistribution of emissions

Authors:

Abstract

Ozone is an important air pollutant at the surface, and the third most important anthropogenic greenhouse gas in the troposphere. Since 1980, anthropogenic emissions of ozone precursors - methane, non-methane volatile organic compounds, carbon monoxide and nitrogen oxides (NO x) - have shifted from developed to developing regions. Emissions have thereby been redistributed equatorwards, where they are expected to have a stronger effect on the tropospheric ozone burden due to greater convection, reaction rates and NO x sensitivity. Here we use a global chemical transport model to simulate changes in tropospheric ozone concentrations from 1980 to 2010, and to separate the influences of changes in the spatial distribution of global anthropogenic emissions of short-lived pollutants, the magnitude of these emissions, and the global atmospheric methane concentration. We estimate that the increase in ozone burden due to the spatial distribution change slightly exceeds the combined influences of the increased emission magnitude and global methane. Emission increases in Southeast, East and South Asia may be most important for the ozone change, supported by an analysis of statistically significant increases in observed ozone above these regions. The spatial distribution of emissions dominates global tropospheric ozone, suggesting that the future ozone burden will be determined mainly by emissions from low latitudes.
LETTERS
PUBLISHED ONLINE: 7 NOVEMBER 2016 | DOI: 10.1038/NGEO2827
Tropospheric ozone change from 1980 to 2010
dominated by equatorward redistribution
of emissions
Yuqiang Zhang1, Owen R. Cooper2,3, Audrey Gaudel2,3, Anne M. Thompson4, Philippe Nédélec5,
Shin-Ya Ogino6and J. Jason West1*
Ozone is an important air pollutant at the surface1, and
the third most important anthropogenic greenhouse gas in
the troposphere2. Since 1980, anthropogenic emissions of
ozone precursors—methane, non-methane volatile organic
compounds, carbon monoxide and nitrogen oxides (NOx)—
have shifted from developed to developing regions. Emissions
have thereby been redistributed equatorwards3–6, where they
are expected to have a stronger eect on the tropospheric
ozone burden due to greater convection,reaction rates and NOx
sensitivity7–11. Here we use a global chemical transport model
to simulate changes in tropospheric ozone concentrations
from 1980 to 2010, and to separate the influences of
changes in the spatial distribution of global anthropogenic
emissions of short-lived pollutants, the magnitude of these
emissions, and the global atmospheric methane concentration.
We estimate that the increase in ozone burden due to the
spatial distribution change slightly exceeds the combined
influences of the increased emission magnitude and global
methane. Emission increases in Southeast, East and South Asia
may be most important for the ozone change, supported by an
analysis of statistically significant increases in observed ozone
above these regions. The spatial distribution of emissions
dominates global tropospheric ozone, suggesting that the
future ozone burden will be determined mainly by emissions
from low latitudes.
Ozone (O3) production in the troposphere, by the oxidation of
carbon monoxide (CO), non-methane volatile organic compounds
(NMVOCs) and methane (CH4) in the presence of nitrogen
oxides (NOx) and sunlight, exceeds the stratosphere-to-troposphere
exchange by a factor of 5–7 (ref. 12). O3is an urban and regional
air pollutant, but is also sufficiently long-lived (22 days globally
averaged12) that its baseline concentrations are elevated over the
entire Northern Hemisphere13 (NH). Observations14 and models15,16
have associated emission increases in Asia with increasing O3above
western North America. The tropospheric O3burden (BO3) is an
important quantity that is related to radiative forcing (RF), as O3
is more effective as a greenhouse gas in the middle and upper
troposphere than near the surface2, and to surface air quality because
it influences both urban and rural baseline ozone.
From 1940 to 1980, global anthropogenic emissions of O3
precursors increased, but the spatial distribution remained fairly
unchanged, with the greatest emissions in the NH middle and high
latitudes. Starting in 1980, emissions began to shift equatorward as
China and low-latitude nations became more industrialized. From
1980 to 2010, global anthropogenic emissions of CO, NOxand
NMVOCs are estimated to have increased by 6.4%, 21.2% and 6.0%
(ref. 5,17), and the global CH4mixing ratio increased by 14.7%
(231 ppbv, Methods). During the same period, emissions of CO
and NMVOCs increased south of 30N but decreased north of
this latitude, while NOxemissions increased south of 40N but
decreased to the north (Supplementary Figs 1 and 2).
Here we investigate the influences of global emission changes
from 1980 to 2010 on BO3and surface O3, separating the influences
of changes in: the spatial distribution of anthropogenic short-
lived emissions; the global magnitude of emissions; and the
global CH4mixing ratio. Simulations are conducted with the
CAM-chem18 global chemical transport model for 1980 and 2010,
and sensitivity simulations alter these three parameters individually
to 1980 conditions, relative to the 2010 simulation (Methods and
Supplementary Table 1).
The global BO3is modelled to have increased by 28.12 Tg
(8.9%) from 1980 to 2010 (four-year averages), with 57% of the
total increase in the NH (Fig. 1). The largest BO3increases are
over 30S–30N (17.93 Tg, Figs 1 and 2). The influence of the
change in the spatial distribution of global anthropogenic emissions
contributes 16.39Tg of the total tropospheric O3burden change
(1BO3), also with a greater influence in the NH than in the Southern
Hemisphere (SH), slightly greater than the combined influences of
the change in emission magnitude (8.59Tg) and the global CH4
change (7.48 Tg) (Fig. 2). The influence of the change in spatial
distribution is greater than the sum of the other two influences
in three of four years modelled, with the interannual variability in
1BO3being much smaller than the overall change (Supplementary
Table 2). The sensitivity of BO3to CH4here (0.123 Tg BO3per Tg
CH4a1) is within the range of other models (0.11–0.16 Tg BO3per
Tg CH4a1, ref. 19). Note that the total 1BO3from the sum of the
three sensitivity simulations (32.46Tg) is larger than the difference
between S_2010 (Supplementary Table 1) and S_1980 (28.12 Tg), as
only one variable is changed to 1980 conditions in each simulation.
Over 30S–30N, the 1BO3from the emission spatial distribution
change is much greater than the other influences. In extratropical
regions, the 1BO3from the emission spatial distribution change
is only slightly greater or comparable to 1BO3from the other
influences. North of 60N, the 1BO3due to the emission spatial
1Environmental Sciences and Engineering Department, University of North Carolina at Chapel Hill, Chapel Hill, North Carolina 27599, USA.
2Cooperative Institute for Research in Environmental Sciences, University of Colorado, Boulder, Colorado 80309, USA. 3Chemical Sciences Division, NOAA
Earth System Research Laboratory, Boulder, Colorado 80305, USA. 4NASA Goddard Space Flight Center, Greenbelt, Maryland 20771, USA. 5Laboratoire
d’Aérologie, CNRS, Université Paul Sabatier Toulouse III, FR-31062 Toulouse, France. 6Japan Agency for Marine-Earth Science and Technology, Yokosuka
237-0061, Japan. Present address: Environmental Protection Agency, Research Triangle Park, North Carolina 27709, USA. *e-mail: jjwest@email.unc.edu
NATURE GEOSCIENCE | ADVANCE ONLINE PUBLICATION | www.nature.com/naturegeoscience 1
© ƐƎƏƖɥMacmillan Publishers LimitedƦɥ/13ɥ.$ɥ/1(-%#1ɥ341#. All rights reservedƥ
LETTERS NATURE GEOSCIENCE DOI: 10.1038/NGEO2827
28.1216.398.597.48
0 6 12 18 24 30
Burden change (Tg O3)
Burden change (Tg O3)
Total
Spatial pattern
Magnitude
Global CH4
Total
Spatial pattern
Magnitude
Global CH4
Global
NH
SH
0246810
>60° N
30°–60° N
0°–30° N
30°–0° S
60°–30° S
90°–60° S
Figure 1 | Tropospheric O3burden change (1BO3) from 1980 to 2010.
1BO3is shown globally, in each hemisphere, and in dierent latitudinal
bands. The estimated components of 1BO3due to the emission spatial
distribution change (red square), magnitude change (blue triangle) and
global CH4change (purple circle) are also seen.
distribution change is lowest, as lower amounts of O3and its
precursors are transported from North America and Europe.
Between 1980 and 2010, the greatest modelled increases in
O3burden are over 10–35N from the surface to the upper
troposphere (Fig. 3). Sixty-eight per cent of 1BO3is below 500 hPa,
although the greatest changes in mixing ratio are in the middle
and upper troposphere (Supplementary Fig. 4). Notable O3
increases are also seen over 30S–10N. Over 35–60N, O3
increases at all altitudes, even though anthropogenic emissions
from North America and Europe decreased between 1980 and 2010
(Supplementary Figs 1–3). The influences of the global emission
magnitude change and the global CH4change both increase O3
over 30S–35N, but the emission spatial distribution change best
explains the overall O3change, particularly the regions with greatest
ozone increases. Increases in O3precursor emissions south of 35N
are transported efficiently to the middle and upper troposphere,
from strong convection in the Hadley cell, whereas emission
decreases north of 35N stay at high latitudes and low elevation in
Ferrell cell circulation (Fig. 4). When global emissions shift equator-
ward, strong convective mixing over the tropics and subtropics lifts
O3and its precursor NOyto higher altitudes (Figs 3b and 4b and
Supplementary Figs 4 and 5), where the O3lifetime is longer, favour-
ing O3accumulation. When emission increases occur in NH mid-
latitudes, less NOyis lofted to high altitudes (Fig. 4c). O3increases
at high altitudes over middle and high latitudes are affected by the
transport of pollutants from the tropics and subtropics16,20 (Fig. 3b).
In addition to strong convection, the tropics and subtropics have
faster chemical reaction rates than other regions (Supplementary
Fig. 6), due to the strong sunlight and warm temperatures.
Therefore, changes in the O3chemical production (PO3), and loss
(LO3) rates are greater for the spatial distribution change than
for the magnitude change, due to greater low-latitude emissions
(Supplementary Fig. 7). Similarly, the distribution change increases
the global ozone production efficiency, whereas the magnitude
change decreases it. In addition, strong NOxsensitivity prevails
over the tropics and subtropics, especially in the middle and upper
troposphere (Supplementary Figs 8 and 9), and emission trends
show greater increases of NOxthan of NMVOCs (Supplementary
Fig. 1). Finally, O3lifetime is lower over the tropics, due to
destruction from water vapour and dry deposition to vegetated
surfaces20. However, this effect is clearly not dominant as we see
larger O3increases over the tropics.
In Fig. 2, the largest modelled 1BO3occurs over South and
Southeast Asia, suggesting a strong influence of emission increases
in these regions. We estimate the importance of different regions
for BO3by multiplying the change in NOxemissions in each of nine
90° N
a
60° N
30° N
0°
30° S
60° S
90° S
180°120° W60° W
Longitude
Latitude
0°
28.12 Tg 16.39 Tg
8.59 Tg 7.48 Tg
60° E 120° E 180°
90° N
60° N
30° N
0°
30° S
60° S
90° S
180°120° W60° W
Longitude
Latitude
0°60° E 120° E 180°
90° N
60° N
30° N
0°
30° S
60° S
90° S
180°120° W60° W
Longitude
Latitude
0°
−0.10 −0.06 −0.02
Tons km−2
0.02 0.06 0.10
−0.08 −0.04 0.00 0.04 0.08
60° E 120° E 180°
90° N
60° N
30° N
0°
30° S
60° S
90° S
180°120° W60° W
Longitude
Latitude
0°60° E 120° E 180°
b
cd
Figure 2 | Spatial distributions for 1BO3(tons km2) from 1980 to 2010. a, Total changes from 1980 to 2010. bd, Influences of changes in the global
emissions spatial distribution (b), the global emissions magnitude (c), and global CH4mixing ratio (d).
2
© ƐƎƏƖɥMacmillan Publishers LimitedƦɥ/13ɥ.$ɥ/1(-%#1ɥ341#. All rights reservedƥ
NATURE GEOSCIENCE | ADVANCE ONLINE PUBLICATION | www.nature.com/naturegeoscience
NATURE GEOSCIENCE DOI: 10.1038/NGEO2827 LETTERS
1,000
850
700
500
400
300
Pressure (mb)
Height (km)
250
200
150
100
a
2
2
2
2
2
4
4
4
6
1,000
850
0°30° S60° S90° S30° N
−9 −7 −5 −3 −1 1 3 5 7 9−8 −6 −4 −2 0 2 4 6 8
60° N90° N0°30° S60° S90° S30° N60° N90° N
0°
Latitude Latitude
Latitude Latitude
30° S60° S90° S30° N60° N90° N0°30° S60° S90° S
4
8
12
16
Height (km)
4
8
12
16
30° N60° N90° N
700
500
400
300
Pressure (mb)
250
200
150
100
Tons km−3
b
cd
Figure 3 | Zonal annual average O3change. a, Total change from 1980 to 2010. bd, Influences of changes in the global emissions spatial distribution (b),
the global emissions magnitude (c), and global CH4mixing ratio (d).
world regions by the sensitivity of BO3per unit NOxemissions from
ref. 9 (Supplementary Table 3). Doing so, we find that emissions
changes from Southeast Asia are most important for the 1980–2010
global 1BO3, followed by East Asia and South Asia. Southeast Asia
emerges as most important—although its emission increase is only
22% that from East Asia—because of the very large sensitivity of
BO3to NOx. Strong convection over Southeast Asia contributes
to this large sensitivity. Convection over the Himalayas in the
NH summer, when the intertropical convergence zone is farthest
north, also suggests a pathway for South Asian emissions to the
upper troposphere20 (Supplementary Figs 10 and 11). Future work
should model the contributions of emissions changes from each
region individually.
Surface O3changes, using the three-month O3-season maximum
daily 8-h average O3(MDA8) (Supplementary Fig. 12), are
dominated by regional emission trends: decreases within Europe
and North America, and increases over East and Southeast Asia,
consistent with observations1,21. Similar regional variations of
MDA8 O3are also seen from the influence of the global emission
spatial pattern change. The MDA8 change between 1980 and
2010 is therefore also dominated by the global emission spatial
pattern change, with smaller contributions from the global emission
magnitude and CH4changes.
These modelled ozone changes are broadly consistent with ob-
served changes over recent decades that show strong increases in
South, East and Southeast Asia (Supplementary Information)21.
In particular, our analyses of ozone observations from IAGOS
commercial aircraft22 and SHADOZ ozonesondes23 in these re-
gions show good agreement with the model, and changes from
1994–2004 to 2005–2014 that are similar to the modelled 1980–2010
changes (Supplementary Figs 13–15 and Supplementary Table 4);
over Southeast Asia and southern India, we show for the first time
statistically significant ozone increases at most elevations, whereas
for northeastern China, we show that the positive trends detected
earlier24 have continued. We also find that S_1980 compares well
with the global ozonesonde climatology of ref. 25, and S_2010 with
that of ref. 26 (Supplementary Figs 16 and 17), but is biased high
at 200 mb in S_1980 between 30S and the Equator, and that bias
increases in S_2010 between 30S and 30N. Given that most of the
modelled 1BO3is below 500 mb, the bias at 200 mb is probably not
very important for our main findings. Comparing the 1980–2010
ozone trends at long-term observation sites, the model overesti-
mates the 1980–2010 ozone change at two of three NH-midlatitude
long-term ozonesonde sites (Supplementary Fig. 18), and captures
well the 1980–2010 O3trend at five of six rural or remote surface
sites, although tending to overestimate the trend and to overestimate
NATURE GEOSCIENCE | ADVANCE ONLINE PUBLICATION | www.nature.com/naturegeoscience
© ƐƎƏƖɥMacmillan Publishers LimitedƦɥ/13ɥ.$ɥ/1(-%#1ɥ341#. All rights reservedƥ
3
LETTERS NATURE GEOSCIENCE DOI: 10.1038/NGEO2827
−0.5 −0.4 −0.3 −0.2 −0.1 0.0 0.1 0.2 0.3 0.4 0.5
Tons km−3
1,000
850
700
500
400
300
Pressure (mb)
250
200
150
100
a
0°
Latitude Latitude
Latitude Latitude
30° S60° S90° S30° N60° N90° N
1,000
850
700
500
400
300
Pressure (mb)
250
200
150
100
0°30° S60° S90° S30° N60° N90° N 0°30° S60° S90° S30° N60° N90° N
90° N0°30° S60° S90° S
4
8
12
16
4
8
12
16
30° N60° N
0.1
0.1
0.1
0.1
0.1
b
cd
Figure 4 | Zonal annual average NOychange. a, Total changes from 1980 to 2010. bd, Influences of changes in the global emissions spatial
distribution (b), the global emissions magnitude (c), and global CH4mixing ratio (d). See Methods for the NOydefinition.
O3in the NH and underestimate in the SH (Supplementary Fig. 19
and Supplementary Table 5). Compared with OMI/MLS satellite ob-
servations27, S_2010 has high biases for BO3, particularly in the trop-
ics and subtropics including over Africa (Supplementary Fig. 20),
that are comparable to the simulated 1980–2010 burden change
(Supplementary Tables 6 and 7), and thus the model may overstate
the magnitude of the tropical and subtropical ozone changes. How-
ever, OMI/MLS trends in ozone columns from 2004 to 2015 show
the greatest growth over South and Southeast Asia (Supplementary
Fig. 21), consistent with the model (Fig. 2). Despite model biases,
these observations provide strong evidence for ozone increases
where the model predicts the greatest increases.
Observations have also shown that the ozone peak has shifted
earlier in the year at rural NH sites, and emissions moving equator-
ward has been hypothesized as an explanation21,28. However, S_2010
does not show the observed shift in the timing of ozone peaks
relative to S_1980, nor does S_Distribution (Supplementary Fig. 22).
Uncertainty in historical emissions and seasonal distributions, or
inaccuracies in model chemistry or physics29, may be the reasons
for our inability to explain these observations.
By using the same meteorology in 1980 and 2010, we neglect the
possible effects of climate change or climate variability on 1BO3.
We also evaluate the contributions of each parameter by simulating
1980 conditions for each parameter individually, relative to the 2010
simulation, and 1BO3would probably be smaller had we evaluated
relative to 1980. However, we expect that the relative contributions
would be similar.
These results are expected to have important implications for
the RF of O3. However, increasing NOxmay cause a negative RF,
due mainly to decreases in CH4, and regions with high sensitivity
of BO3to NOxemissions also have a high sensitivity of CH4to
NOx, with the two RFs roughly cancelling over all source regions7,10.
CO is sufficiently long-lived that the sensitivity of BO3does not
vary strongly with the location of CO emissions30. The effect of the
equatorward emission shift on RF should be investigated further.
The change in the spatial distribution of the global anthropogenic
emissions from 1980 to 2010 dominates the BO3change, and is
slightly greater than the combined effects of changes in the global
emission magnitude and global CH4. In particular, increases in
O3precursor emissions in the tropics and subtropics significantly
influence the global BO3, and our findings suggest that emissions
increases from Southeast, East and South Asia have been most
important for the BO3increase. This can be attributed to the strong
photochemical reaction rates, convection and NOxsensitivity in the
4
© ƐƎƏƖɥMacmillan Publishers LimitedƦɥ/13ɥ.$ɥ/1(-%#1ɥ341#. All rights reservedƥ
NATURE GEOSCIENCE | ADVANCE ONLINE PUBLICATION | www.nature.com/naturegeoscience
NATURE GEOSCIENCE DOI: 10.1038/NGEO2827 LETTERS
tropics and subtropics. As a result, the global BO3might continue to
increase due to a continued equatorward shift of emissions, even if
global anthropogenic emissions remain unchanged or decrease. The
location of emissions and the dominant role of emissions from the
tropics and subtropics deserve greater emphasis in future research
and projections of global tropospheric ozone.
Methods
Methods, including statements of data availability and any
associated accession codes and references, are available in the
online version of this paper.
Received 29 March 2016; accepted 27 September 2016;
published online 7 November 2016
References
1. The Royal Society Ground-Level Ozone in the 21st Century: Future Trends,
Impacts and Policy Implications (The Royal Society Science Policy, 2008).
2. Myhre, G. et al. in Climate Change 2013: The Physical Science Basis
(eds Stocker, T. F. et al.) Ch. 8 (IPCC, Cambridge Univ. Press, 2013).
3. Richter, A., Burrows, J. P., Nüss, H., Granier, C. & Niemeier, U. Increase in
tropospheric nitrogen dioxide over China observed from space. Nature 437,
129–132 (2005).
4. Duncan, B. N. et al. A space-based, high-resolution view of notable changes in
urban NOxpollution around the world (2005–2014). J. Geophys. Res. 121,
976–996 (2016).
5. Lamarque, J. F. et al . Historical (1850–2000) gridded anthropogenic and
biomass burning emissions of reactive gases and aerosols: methodology and
application. Atmos. Chem. Phys. 10, 7017–7039 (2010).
6. Granier, C. et al. Evolution of anthropogenic and biomass burning emissions of
air pollutants at global and regional scales during the 1980–2010 period.
Climatic Change 109, 163–190 (2011).
7. Naik, V. et al. Net radiative forcing due to changes in regional emissions of
tropospheric ozone precursors. J. Geophys. Res. 110, D24306 (2005).
8. Derwent, R. G. et al. Radiative forcing from surface NOxemissions: spatial and
seasonal variations. Climatic Change 88, 385–401 (2008).
9. West, J. J., Naik, V., Horowitz, L. W. & Fiore, A. M. Effect of regional precursor
emission controls on long-range ozone transport—Part 1: short-term changes
in ozone air quality. Atmos. Chem. Phys. 9, 6077–6093 (2009).
10. Fry, M. M. et al. The influence of ozone precursor emissions from four world
regions on tropospheric composition and radiative climate forcing. J. Geophys.
Res. 117, D07306 (2012).
11. Gupta, M. L., Cicerone, R. J. & Elliott, S. Perturbation to global tropospheric
oxidizing capacity due to latitudinal redistribution of surface sources of NOx,
CH4and CO. Geophys. Res. Lett. 25, 3931–3934 (1998).
12. Young, P. J. et al. Pre-industrial to end 21st century projections of tropospheric
ozone from the Atmospheric Chemistry and Climate Model Intercomparison
Project (ACCMIP). Atmos. Chem. Phys. 13, 2063–2090 (2013).
13. Hemispheric Transport of Air Pollution 2010—Part A Ozone and Particulate
Matter (United Nations Economic Commission for Europe, 2010).
14. Cooper, O. R., Gao, R.-S., Tarasick, D., Leblanc, T. & Sweeney, C. Long-term
ozone trends at rural ozone monitoring sites across the United States,
1990–2010. J. Geophys. Res. 117, D22307 (2012).
15. Cooper, O. R., Andrew, O., Parrish, D. D. & Fahey, D. W. Challenges of a
lowered U.S. ozone standard. Science 348, 1096–1097 (2015).
16. Verstraeten, W. W. et al. Rapid increases in tropospheric ozone production and
export from China. Nat. Geosci. 8, 690–695 (2015).
17. Riahi, K. et al. RCP 8.5-A scenario of comparatively high greenhouse gas
emissions. Climatic Change 109, 33–57 (2011).
18. Lamarque, J.-F. et al. CAM-chem: description and evaluation of interactive
atmospheric chemistry in the community Earth system model. Geosci. Model
Dev. 5, 369–411 (2012).
19. Fiore, A. M., West, J. J., Horowitz, L. W., Naik, V. & Schwarzkopf, M. D.
Characterizing the tropospheric ozone response to methane emission
controls and the benefits to climate and air quality. J. Geophys. Res. 113,
D08307 (2008).
20. Lawrence, M. G., von Kuhlmann, R., Salzmann, M. & Rasch, P. J. The balance
of effects of deep convective mixing on tropospheric ozone. Geophys. Res. Lett.
30, 3–6 (2003).
21. Cooper, O. R. et al. Global distribution and trends of tropospheric ozone:
an observation-based review. Elementa 2, 000029 (2014).
22. Petzold, A. et al. Global-scale atmosphere monitoring by in-service
aircraft—current achievements and future prospects of the European Research
Infrastructure IAGOS. Tellus B 67, 28452 (2015).
23. Thompson, A. M. et al. Southern Hemisphere Additional Ozonesondes
(SHADOZ) 1998–2004 tropical ozone climatology: 3. Instrumentation,
station-to-station variability, and evaluation with simulated flight profiles.
J. Geophys. Res. 112, D03304 (2007).
24. Ding, A. J. et al. Tropospheric ozone climatology over Beijing: analysis of
aircraft data from the MOZAIC program. Atmos. Chem. Phys. 8, 1–13 (2008).
25. Logan, J. A. An analysis of ozonesonde data for the troposphere:
recommendations for testing 3-D models and development of a gridded
climatology for tropospheric ozone. J. Geophys. Res. 104, 16115–16149 (1999).
26. Tilmes, S. et al. Technical note: ozonesonde climatology between 1995 and
2011: description, evaluation and applications. Atmos. Chem. Phys. 12,
7475–7497 (2012).
27. Ziemke, J. R. et al. A global climatology of tropospheric and stratospheric
ozone derived from Aura OMI and MLS measurements. Atmos. Chem. Phys.
11, 9237–9251 (2011).
28. Parrish, D. et al. Lower tropospheric ozone at northern midlatitudes: changing
seasonal cycle. Geophys. Res. Lett. 40, 1631–1636 (2013).
29. Parrish, D. D. et al. Long-term changes in lower tropospheric baseline ozone
concentrations: comparing chemistry-climate models and observations at
northern midlatitudes. J. Geophys. Res. 119, 719–5736 (2014).
30. Fry, M. M. et al. Net radiative forcing and air quality responses to regional CO
emission reductions. Atmos. Chem. Phys. 13, 5381–5399 (2013).
Acknowledgements
Y.Z. and J.J.W. were funded by National Institute of Environmental Health Sciences grant
no. 1 R21 ES022600-01 and Environmental Protection Agency STAR grants no. 834285
and RD83587801, and O.R.C. and A.G. were funded by NOAAs Health of the
Atmosphere and Atmospheric Chemistry and Climate Programs. The contents are solely
the responsibility of the grantee and do not necessarily represent the official views of the
US EPA or other funding sources. We thank the NCAR AMWG for developing and
maintaining the diagnostic package for the model evaluation. We acknowledge the free
use of O3observation data from NOAA GMD for the remote sites of Barrow, Mauna Loa,
Samoa and South Pole; Global Atmosphere Watch World Data Centre for Greenhouse
Gases for Hohenpeissenberg, J. Schwab from University at Albany-SUNY for Whiteface
Mountain, and P. Young of Lancaster University for processed ozonesonde climatology
of ref. 25.
Author contributions
Y.Z., J.J.W. and O.R.C. designed the study and Y.Z. and J.J.W. planned the model
experiments. Y.Z. prepared the emission inputs, performed the model simulations, and
prepared the figures. Y.Z. and A.G. conducted data analysis for observations, and J.J.W.
and O.R.C. assisted with the data analysis. P.N., S.-Y.O. and A.M.T. provided
observational data. Y.Z., J.J.W. and O.R.C. wrote the paper with comments from A.M.T
and A.G.
Additional information
Supplementary information is available in the online version of the paper. Reprints and
permissions information is available online at www.nature.com/reprints.
Correspondence and requests for materials should be addressed to J.J.W.
Competing financial interests
The authors declare no competing financial interests.
NATURE GEOSCIENCE | ADVANCE ONLINE PUBLICATION | www.nature.com/naturegeoscience
© ƐƎƏƖɥMacmillan Publishers LimitedƦɥ/13ɥ.$ɥ/1(-%#1ɥ341#. All rights reservedƥ
5
LETTERS NATURE GEOSCIENCE DOI: 10.1038/NGEO2827
Methods
Global emissions spatial pattern change. Here we use anthropogenic emissions
including biomass burning in 1980 from the Atmospheric Chemistry and Climate
Model Intercomparison Project (ACCMIP, ref. 5), and in 2010 from the
representative concentration pathways 8.5 (RCP8.5) scenario17,31, which are
downloaded from the RCP database (http://tntcat.iiasa.ac.at:8787/RcpDb/dsd?
Action=htmlpage&page=download, accessed 10/31/2014) to analyse emission
trends and drive the global model (Supplementary Figs 1 and 2). RCP8.5 2010
emissions are self-consistent with the ACCMIP historical emissions, and are
considered to be the most reasonable scenario to extend ACCMIP emissions
beyond 2000 (ref. 6) as they track the GAINS current legislation scenario for
several decades17. The 2010 global total anthropogenic emissions of CO, NOxand
non-methane volatile organic compounds (NMVOCs) are 1030 Tg, 82 Tg NO and
180 Tg, respectively. As with other global emission estimates for short-lived species,
emissions are uncertain and may be especially uncertain in some developing
regions29,32–34. The rapid growth of emissions in the tropics and subtropics is seen in
global emission inventories5,6 as well as regional ones35,36. These inventories are
supported by observations from satellites3,4,37,38 and from surface and airborne
observations29,39 that show that NO2in developed regions, such as Europe and
North America, have greatly diminished emissions since 1980, but the emissions
are increasing in developing countries, especially China and India, shifting global
emissions equatorward.
CAM-chem model configuration. We use the global chemistry–climate model
CAM-chem, which is based on the global Community Atmosphere Model (CAM)
version 4, the atmospheric component of the Community Earth System Model
(CESM, v1.2.2) (refs 18,40). The model has a horizontal resolution of 1.9(latitude)
×2.5(longitude), and 56 vertical levels between the surface and 4 hPa (40 km)
with a time step of 1,800 s. We use the NASA Global Modeling and Assimilation
Office GEOS-5 meteorology to drive the model as a chemical transport model. For
all simulations, we run CAM-chem for five consecutive years, with the first year as
spin-up and results are shown as a four-year average. Monthly mean distributions
of chemically active stratospheric species (such as O3, NO, NO2and N2O5) are
prescribed using the climatology from the Whole Atmospheric Community
Climate Model simulations41, following ref. 18. We use a single global chemical
transport model, and results with other models may differ, particularly due to
different reaction mechanisms.
NMVOC species from ACCMIP (1980) and RCP8.5 (2010) are both
re-speciated to match the CAM-chem VOC categories following previous
studies30,42,43. Monthly temporal variations for all anthropogenic emission sectors
are also added based on the monthly time dependence of emissions from
RETRO30,42–44, except for aircraft, shipping and biomass burning for which seasonal
variations were provided. The Model of Emissions of Gases and Aerosols from
Nature (MEGAN-v2.1, ref. 45) simulates biogenic emissions for 150 compounds
online within CAM-chem, yielding four-year average global biogenic emissions of
isoprene, monoterpene, methanol and acetone of 420.69 Tgyr1, 133.23 Tg yr1,
91.99 Tg yr1and 42.67 Tg yr1. Lightning NOxemissions are calculated online as
3.21 TgN yr1(four-year average), which is lower than the average of ACCMIP
models for 2000, but within the range12; lower lightning NOxemissions may affect
the sensitivity of ozone to NOxand VOCs, particularly in the tropics and
mid-troposphere where lightning emissions are greatest. Other natural emissions
(ocean, volcano and soil) are from the standard CAM-chem emission files (for
2000), and remain the same for all of the simulations, with soil NOxat 8.0 TgN yr1
(refs 18,46). The CH4volume mixing ratio (ppb) is fixed at uniform global values of
1,567 and 1,798 ppbv for 1980 and 2010 (ref. 47).
In addition to simulating 1980 and 2010, we conduct three sensitivity
simulations in which the spatial distribution of global anthropogenic emissions
(S_Distribution), the magnitude of the global emissions (S_Magnitude), and the
global CH4mixing ratio (S_CH4) are individually set to 1980 levels and all other
parameters stay the same as the 2010 simulation (Supplementary Table 1). Here
global anthropogenic emissions refer to all short-lived species, including ozone
precursors and other species such as aerosols, from anthropogenic sources
including biomass burning. The differences between S_2010 and S_1980 reflect the
total emission changes from 1980 to 2010. Each of the other three simulations is
subtracted from S_2010 to isolate individual influences. We use meteorology from
2009–2012 with 2008 as a spin-up for all simulations, isolating the effects of
changes in emissions and neglecting possible effects of climate variability or change
from 1980 to 2010.
Tropospheric O3burden (BO3) is defined as the total below the chemical
tropopause of 150 ppbv ozone in the S_2010 simulation, with the same tropopause
applied to all simulations. The four-year average BO3in S_2010 is 342.7 Tg, within
the range of ACCMIP models (337 ±23 Tg for 1995–2005), and the 1980–2010
1BO3of 28.12 Tg is similar to the ACCMIP 15 ±6 Tg for 1980–2000, and 41 ±12
for 1980–2030 (RCP8.5) (ref. 12). The three-month ozone season average MDA8 is
found for the consecutive three-month period with the highest MDA8 in each grid
cell. NOy(total reactive nitrogen) is calculated as NO +NO2+NO3+HNO3+
HO2NO2+2×N2O5+CH3CO3NO2(PAN) +CH2CCH3CO3NO2(MPAN,
methacryloyl peroxynitrate) +CH2CHCCH3OOCH2ONO2(ISOPNO3, peroxy
radical from NO3+isoprene) from CAM-chem output.
CAM-chem evaluation. A comprehensive evaluation of S_2010 is performed using
a present-day climatology of O3data from multi-year observations from
ozonesondes, satellites, aircraft campaigns, and ground-based observations,
compared with the four-year average of CAM-chem output. Our model captures
the vertical distribution of O3in ozonesondes26 very well, although it is biased high
between 30S and 30N, particularly in the upper troposphere (Supplementary
Fig. 23 and Supplementary Table 8). The seasonal cycles of O3at specific pressure
levels from the ozonesonde data are also captured well by the model
(Supplementary Fig. 24), with a correlation coefficient between the observed and
simulated monthly regional O3average that is usually greater than 0.8
(Supplementary Fig. 25). When evaluating model performance with aircraft
campaign observations, we focus on the regional average over the campaign areas,
and analyse the data at different altitudes. Generally, the model performs better in
the NH than the SH (Supplementary Fig. 26 and Supplementary Table 9). When
evaluated with multi-year satellite data27, the model overestimates O3between
20S and 40N, which is common for global models12, with a modelled global BO3
that is 23.8 Tg higher (Supplementary Fig. 20). Compared with surface O3
observations, S_2010 overestimates O3by 5.75 ppbv averaged over all stations in
the US (Supplementary Fig. 27), and 0.65 ppbv over Europe (Supplementary
Fig. 28), but captures well the seasonal cycles.
Code availability. The CAM-chem model code used to perform all the simulations
is available at: https://www2.cesm.ucar.edu/models.
The diagnostic package used to perform the model evaluation is developed and
maintained by the NCAR AMWG, and code can be found at:
https://www2.cesm.ucar.edu/working-groups/amwg/amwg-diagnostics-
package/find-code.
Data availability. Hourly O3observations for the remote sites of Barrow,
Mauna Loa, Samoa and South Pole are maintained by the NOAA Global
Monitoring Division (GMD) and can be found at: ftp://aftp.cmdl.noaa.gov/data/
ozwv/SurfaceOzone.
Hourly ozone data from Hohenpeissenberg for the years 1971–2010 were
downloaded from the Global Atmosphere Watch (GAW) World Data Centre for
Greenhouse Gases: http://ds.data.jma.go.jp/gmd/wdcgg. The Meteorological
Observatory Hohenpeissenberg is operated and financed by the German
Meteorological Service (DWD).
The Whiteface Mountain Summit ozone data were collected by the University
at Albany-SUNY with instrumentation provided by the New York State
Department of Environmental Conservation. The data set was provided by
J. J. Schwab, Atmospheric Sciences Research Center, University at Albany-SUNY,
and is archived by the United States Environmental Protection Agency:
http://www.epa.gov/ttn/airs/aqsdatamart.
Ozone profiles from commercial aircraft were collected and made freely
available by the Measurement of Ozone and water vapour on Airbus In-service
airCraft – In-service Aircraft for a Global Observing System (MOZAIC-IAGOS)
programme (www.iagos.org). The data were made possible by: the European
Commission’s support for the MOZAIC project (1994–2003) and the preparatory
phase of IAGOS (2005–2012); the partner institutions of the IAGOS Research
Infrastructure (FZJ, DLR, MPI, KIT in Germany, CNRS, CNES, Météo-France in
France and the University of Manchester in the UK); ETHER (CNES-CNRS/INSU)
for hosting the database; and the participating airlines (Lufthansa, Air France,
Austrian, China Airlines, Iberia, Cathay Pacific) for the transport free of charge of
the instrumentation.
The Hanoi, Vietnam ozonesondes were made freely available by the
NASA Southern Hemisphere ADditional OZonesondes (SHADOZ) project23.
The processed ozonesonde climatology of ref. 25 was shared by P. Young of
Lancaster University.
Monthly OMI/MLS tropospheric column ozone data27 were provided by
J. Ziemke, Morgan State University, Baltimore, and downloaded from:
http://acd-ext.gsfc.nasa.gov/Data_services/cloud_slice. The OMI/MLS
tropospheric column ozone product is derived from the Ozone Monitoring
Instrument (OMI) and Microwave Limb Sounder (MLS) remote sensors onboard
NASA’s polar orbiting Aura satellite. MLS retrievals are the latest available,
version 4.2.
The data that support the findings of this study are available from the
corresponding author on request.
References
31. Moss, R. H. et al. The next generation of scenarios for climate change research
and assessment. Nature 463, 747–756 (2010).
32. Schopp, W., Klimont, Z., Suutari, R. & Cofala, J. Uncertainty analysis of
emissions estimates in the RAINS integrated assessment model. Environ. Sci.
Policy 8, 601–613 (2005).
© ƐƎƏƖɥMacmillan Publishers LimitedƦɥ/13ɥ.$ɥ/1(-%#1ɥ341#. All rights reservedƥ
NATURE GEOSCIENCE |www.nature.com/naturegeoscience
NATURE GEOSCIENCE DOI: 10.1038/NGEO2827 LETTERS
33. Bond, T. C. et al. Historical emissions of black and organic carbon aerosol from
energy-related combustion, 1850–2000. Glob. Biogeochem. Cycles 21,
GB2018 (2007).
34. Smith, S. J. et al. Anthropogenic sulfur dioxide emissions: 1850–2005.
Atmos. Chem. Phys. 11, 1101–1116 (2011).
35. Ohara, T. et al. An Asian emission inventory of anthropogenic emission
sources for the period 1980–2020. Atmos. Chem. Phys. 7, 4419–4444 (2007).
36. Liousse, C., Essamoi, E., Criqui, P., Granier, C. & Rosset, R. Explosive growth in
African combustion emissions from 2005 to 2030. Environ. Res. Lett. 9,
035003 (2014).
37. van der A, R. J. et al. Trends, seasonal variability and dominant NOxsource
derived from a ten year record of NO2measured from space. J. Geophys. Res.
113, D04302 (2008).
38. Hilboll, A., Richter, A. & Burrows, J. P. Long-term changes of tropospheric NO2
over megacities derived from multiple satellite instruments. Atmos. Chem.
Phys. 13, 4145–4169 (2013).
39. Thompson, A. M. et al. Tropospheric ozone increases over the southern Africa
region: bellwether for rapid growth in Southern Hemisphere pollution?
Atmos. Chem. Phys. 14, 9855–9869 (2014).
40. Tilmes, S. et al. Description and evaluation of tropospheric chemistry and
aerosols in the Community Earth System Model (CESM1.2). Geosci. Model
Dev. 8, 1395–1426 (2015).
41. Garcia, R. R., Marsh, D. R., Kinnison, D. E., Boville, B. A. & Sassi, F. Simulation
of secular trends in the middle atmosphere, 1950–2003. J. Geophys. Res. 112,
D09301 (2007).
42. Fry, M. M., Schwarzkopf, M. D., Adelman, Z. & West, J. J. Air quality
and radiative forcing impacts of anthropogenic volatile organic
compound emissions from ten world regions. Atmos. Chem. Phys. 14,
523–535 (2014).
43. West, J. J. et al. Co-benefits of global greenhouse gas mitigation for
future air quality and human health. Nat. Clim. Change 3,
885–889 (2013).
44. Schultz, M. G. et al.REanalysis of the TROpospheric Chemical Composition over
the Past 40 Years (RETRO)—A Long-term Global Modeling Study of Tropospheric
Chemistry Final Report (Jülich/Hamburg, Germany,
August 2007).
45. Guenther, A. B. et al. The model of emissions of gases and aerosols from nature
version 2.1 (MEGAN2.1): an extended and updated framework for modeling
biogenic emissions. Geosci. Model Dev. 5, 1471–1492 (2012).
46. Emmons, L. K. et al. Description and evaluation of the model for ozone and
related chemical tracers, version 4 (MOZART-4). Geosci. Model Dev. 3,
43–67 (2010).
47. Prather, M. et al. in Climate Change 2013: The Physical Science Basis
(eds Stocker, T. F. et al.) Annex II (IPCC, Cambridge Univ. Press, 2013).
NATURE GEOSCIENCE |www.nature.com/naturegeoscience
© ƐƎƏƖɥMacmillan Publishers LimitedƦɥ/13ɥ.$ɥ/1(-%#1ɥ341#. All rights reservedƥ
... Since 1980s, anthropogenic emissions of tropospheric ozone precursors from North America and Western Europe have decreased in response to the implementation of control measures while the emissions from Asia, Central America, and Eastern Europe have increased due to economic and population growth (Granier et al., 2011, Cooper et al., 2014 leading to an equatorward shift in global emission pattern. Zhang et al., (2016) found that this equatorward shift in precursor emissions is the dominant factor, compared to the change in magnitude of emissions and methane concentration, that has led to an increase 55 in the tropospheric ozone burden between 1980 and 2010. This is due to larger convection of polluted air masses from the boundary layer into the free troposphere over the tropical regions compared to extra-tropical regions. ...
... They demonstrated that the ozone production efficiency can directly be calculated using this method, as the ratio of tropospheric ozone attributed to a tagged emission source to the amount of precursor emission from 100 that source. NOx emissions from tropical regions such as South Asia, Southeast Asia and Central America were found to be the most efficient at producing tropospheric ozone compared to the emissions from other regions, consistent with the earlier work of Zhang et al, (2016). They further showed that using a methane perturbation simulation introduced to NOx-tagged simulation, that the contribution to tropospheric ozone burden by NOx emitted from international shipping increases especially strongly in response to changes in methane concentration. ...
... The correlation coefficient between the observed and simulated monthly regional O3 average is usually more than 0.8, and the fractional mean difference is usually within 25 % at most regions in the troposphere. Our free troposphere evaluation results 175 are consistent with previously evaluated versions of CAM4-Chem (Tilmes et al., 2012, Zhang et al., 2016, Emmons et al., 2020. ...
Preprint
Full-text available
Over the past few decades, the tropospheric ozone precursor anthropogenic emissions: nitrogen oxides (NOx) and reactive carbon (RC) from mid-latitude regions have been decreasing, and those from Asia and tropical regions have been increasing, leading to an equatorward emission redistribution. In this study, we quantify the contributions of various sources of NOx and RC emissions to tropospheric ozone using a source attribution technique during the 2000–2018 period in a global chemistry transport model: CAM4-Chem. We tag the ozone molecules with the source of their NOx or RC precursor emission in two separate simulations, one for each of NOx and RC. These tags include various natural (biogenic, biomass burning, lightning and methane), and regional anthropogenic (North American, European, East Asian, South Asian etc.) precursor emission sources. We simulate ~336 Tg O3 with an increasing trend of 0.91 Tg O3/yr (0.28 %/yr), largely contributed (and trend driven) by anthropogenic NOx emissions and methane. The ozone production efficiency of regional anthropogenic NOx emissions increases significantly when emissions decrease (Europe, North American and Russia-Belarus-Ukraine region’s emissions) and decreases significantly when emissions increase (South Asian, Middle Eastern, International Shipping etc.). Tropical regions, despite smaller emissions, contribute more to tropospheric ozone burden compared to emissions from higher latitudes, consistent with previous work, due to large convection at the tropics thereby lifting O3 and its precursor NOx molecules into the free troposphere where ozone’s lifetime is longer. We contrast the contribution to tropospheric ozone burden with that of the contribution to the global surface ozone. We simulate a smaller relative contribution from tropical regions to the global mean surface ozone compared to their contribution to the tropospheric ozone burden. The global population-weighted mean ozone (related to ozone exposure) is much larger compared to surface mean, mainly due to large anthropogenic emissions from densely populated regions: East Asia, South Asia, and other tropical regions, and a substantial contribution from international ship NOx emissions. The increasing trends in anthropogenic emissions from these regions are the main drivers of increasing global population-weighted mean ozone.
... from 1990 to 2019, respectively (Pandey et al., 2021). Meanwhile, the faster chemical reaction rates in India due to the strong convection, sunlight, and warm 45 temperatures, making it a hot spot for accumulating major air pollutants compared with other regions and easily affecting the air quality in downwind regions (Zhang et al., 2016(Zhang et al., , 2021a. ...
Preprint
Full-text available
Anthropogenic and biomass burning emissions are the major sources of ambient air pollution. India has experienced a dramatic deterioration in air quality over the past few decades, but no systematic assessment has been made to investigate the individual contributions of anthropogenic and biomass burning emissions. In this study, we conducted a pioneering comprehensive analysis of the long-term trends of particulate matter with aerodynamic diameters < 2.5 μm (PM2.5) and ozone (O3) in India and their mortality burden changes from 1995 to 2014, using a state-of-the-art high-resolution global chemical transport model (CAM-chem). Our simulations revealed a substantial nationwide increase in annual mean PM2.5 (6.71 μg m-3 decade-1) and O3 (7.08 ppbv decade-1), with the Indo-Gangetic Plain (IGP) and eastern central India as hotspots for PM2.5 and O3 trend changes individually. Noteworthy substantial O3 decreases were observed in the northern IGP which were potentially linked to NO titration due to a surge in NOx emissions. Sensitivity analyses highlighted anthropogenic emissions as primary contributors to rising PM2.5 and O3, while biomass burning played a prominent role in winter and spring. In years with high biomass burning activity, the contributions from BB on both PM2.5 and O3 changes were comparable with or even exceeding anthropogenic emissions in specific areas. The elevated air pollutants were associated with increased premature mortality attributable to PM2.5 and O3, leading to 97.83 K and 73.91 K per decade. Despite a per capita decrease in the IGP region, the increased population offset its effectiveness.
... Tropospheric column ozone trends derived from satellite products such as OMI/MLS also reported +2 to +3 DU (Dobson unit) per decade increases in the same region from 2005-2016 . Typical explanations for these rapid changes are the increase in global methane (Staniaszek et al., 2022) and a shift of ozone precursor emissions toward tropical latitudes (Zhang et al., 2016) that cause growth in background tropospheric ozone amounts. The interaction of climate oscillations like the El Niño-Southern Oscillation (ENSO; Ziemke et al., 1997Ziemke et al., , 2015Lee et al., 2010;Thompson et al., 2011), the Madden-Julian Oscillation (MJO; Stauffer et al., 2018), and the Indian Ocean Dipole with tropical tropospheric ozone variability is also well established. ...
Article
Full-text available
Positive trends in tropical free-tropospheric (FT) ozone are frequently ascribed to emissions growth, but less is known about the effects of changing dynamics. Extending a prior study (Thompson et al., 2021; https://doi.org/10.1029/2021JD034691; “T21”), we re-examine Southern Hemisphere Additional Ozonesondes (SHADOZ) ozone trends over equatorial Southeast Asia (ESEA), one of Earth's most convectively active regions, using 25 years (1998–2022) of ozone soundings. T21 posited that early-year positive FT ozone trends at equatorial SHADOZ stations are related to decreasing convection. The 25-year analysis of Kuala Lumpur and Watukosek SHADOZ records finds that FT ozone trends of +5 % to +15 % (+2 to +6 nmol mol−1) per decade from ∼ February–April coincide with large increases in satellite infrared brightness temperatures and outgoing longwave radiation, indicators of declining convective activity. MERRA-2 reanalyses exhibit increases in upper-tropospheric velocity potential and decreases in precipitable water, also indicating diminished convection. In contrast, trends in ozone and convective indicators are weak the rest of the year. These results suggest that decreases in convective intensity and frequency are primary drivers of FT ozone build-up over ESEA early in the year; i.e., waning convection suppresses lofting and dilution of ozone. Decreasing convection promotes accumulation of biomass burning emissions typical of boreal spring even though satellite FT carbon monoxide trends (2002–2022) over ESEA follow a global decrease pattern. Finally, our results demonstrate the advantages of monthly or seasonally resolved analyses over annual means for robust attribution of observed ozone trends, challenging models to reproduce these detailed features in simulations of the past 25 years.
... From 1990 onwards, a significant amount of the anthropogenic emissions responsible for O3 formation have shifted 74 from North America and Europe to Asia (Granier et al., 2011; Cooper et al., 2014;Zhang et al., 2016). In East 75 Asia, the overall long-term trend of the daytime average near-surface O3 is 0.45 ppb a -1 , contrasting with a trend 76 of − 0.28 ppb a -1 in North America in the summertime (April-September) during 2000-2014 (Chang et al., 2017). ...
Preprint
Full-text available
Long-term ozone (O3) changes in the middle to upper troposphere are critical to climate radiative forcing and tropospheric O3 pollution. Yet, these changes remain poorly quantified through observations in East Asia. Concerns also persist regarding the data quality of the ozonesondes available at the World Ozone and Ultraviolet Data Center (WOUDC) for this region. This study aims to address these gaps by analyzing O3 soundings at four sites along the northwestern Pacific coastal region over the past three decades, and assessing their consistency with an atmospheric chemistry-climate model simulation. Utilizing the European Centre for Medium-Range Weather Forecasts (ECMWF) – Hamburg (ECHAM)/Modular Earth Submodel System (MESSy) Atmospheric Chemistry (EMAC) nudged simulations, it is demonstrated that trends between model and ozonesonde measurements are overall consistent, thereby gaining confidence in the model’s ability to simulate ozone trends and confirming the utility of potentially imperfect observational data. A notable increase in O3 mixing ratio around 0.29–0.82 ppb a-1 extending from the middle to upper troposphere is observed in both observations and model simulations between 1990 and 2020, primarily during spring and summer. The timing of these O3 hotspots is delayed when moving from south to north along the measurement sites, transitioning from late spring to summer. Investigation into the drivers of these trends using tagged model tracers reveals that ozone of stratospheric origin (O3S) dominates the absolute O3 mixing ratios over the middle-to-upper troposphere in the subtropics, contributing to the observed O3 increases by up to 96 % (40 %) during winter (summer), whereas ozone of tropospheric origin (O3T) governs the absolute value throughout the tropical troposphere and contributes generally much more than 60 % to the positive O3 changes, especially during summer and autumn. During winter and spring, a decrease of O3S is partly counterbalanced by an increase of O3T in the tropical troposphere. This study highlights that the enhanced downward transport of stratospheric O3 into the troposphere in the subtropics and a surge of tropospheric source O3 in the tropics are the two key factors driving the enhancement of O3 in the middle-upper troposphere along the Northwest Pacific region.
Article
Full-text available
High-quality long-term observational records are essential to ensure appropriate and reliable trend detection of tropospheric ozone. However, the necessity of maintaining high sampling frequency, in addition to continuity, is often under-appreciated. A common assumption is that, so long as long-term records (e.g., a span of a few decades) are available, (1) the estimated trends are accurate and precise, and (2) the impact of small-scale variability (e.g., weather) can be eliminated. In this study, we show that the undercoverage bias (e.g., a type of sampling error resulting from statistical inference based on sparse or insufficient samples, such as once-per-week sampling frequency) can persistently reduce the trend accuracy of free tropospheric ozone, even if multi-decadal time series are considered. We use over 40 years of nighttime ozone observations measured at Mauna Loa, Hawaii (representative of the lower free troposphere), to make this demonstration and quantify the bias in monthly means and trends under different sampling strategies. We also show that short-term meteorological variability remains a cause of an inflated long-term trend uncertainty. To improve the trend precision and accuracy due to sampling bias, two remedies are proposed: (1) a data variability attribution of colocated meteorological influence can efficiently reduce estimation uncertainty and moderately reduce the impact of sparse sampling, and (2) an adaptive sampling strategy based on anomaly detection enables us to greatly reduce the sampling bias and produce more accurate trends using fewer samples compared to an intense regular sampling strategy.
Preprint
Full-text available
In the framework of the International Network to study Deposition and Atmospheric chemistry in Africa (INDAAF) program, we present the seasonal variability of atmospheric ozone concentrations at the regional scale. The contributions of local atmospheric chemistry and meteorological parameters to ozone photochemistry are investigated, as are long-term trends in ozone concentrations. Fourteen measurement sites were identified for this study, representative of the main African ecosystems: dry savannas (Banizoumbou, Niger, Katibougou and Agoufou, Mali, Bambey and Dahra, Senegal), wet savannas (Lamto, Côte d'Ivoire, Djougou, Benin), forests (Zoétélé, Cameroon, Bomassa, Republic of Congo) and semi-agricultural/arid savanna (Mbita, Kenya, Louis Trichardt, Amersfoort, Skukuza and Cape Point, South Africa). Over the period 1995–2020, monthly ozone concentrations were measured at these sites using passive samplers. Monthly averages of surface ozone range from 4.7±1.4 ppb (Bomassa) to 31.0±10.5 ppb (Louis Trichardt). Ozone levels in the wet season (in dry savanna) are higher and comparable to concentrations in the dry season (in wet savanna and forest). In East Africa, ozone levels show no marked seasonality, with higher levels in Southern Africa. In the Sahel, under the influence of temperature, ozone formation is closely linked to biogenic Volatile Organic Carbon (VOC) emissions. It is also sensitive to nitrogen monoxide (NO) emissions in the presence of high precipitation and humidity. Biogenic VOC (BVOC) emissions, anthropogenic NOx, temperature and radiation are the dominant contributors to O3 formation in wet savannas and forests. At the southern African sites, the most important parameters influencing surface ozone concentrations are humidity, temperature, VOC emissions (anthropogenic and biogenic) and NOx. At the annual scale, Katibougou and Banizoumbou sites (dry savanna) experienced a significant decrease in ozone concentrations respectively around -0.24 ppb yr-1 (-2.3 % yr-1) and -0.15 ppb yr-1 (-1.9 % yr-1). Seasonal Kendall statistical tests revealed decreasing trends of -0.07 ppb yr-1 in Banizoumbou and -0.24 ppb yr-1 in Katibougou. These decreasing trends are consistent with those observed for nitrogen dioxide (NO2) and biogenic VOCs. A significant increasing trend is observed in Zoétélé, with the Sen slope estimated at 0.1 ppb yr-1 and at Skukuza (Sen slope = 0.3 ppb yr-1). The increasing trends are consistent with the increase in biogenic emissions at Zoétélé and NO2 levels at Skukuza. Very few surface O3 measurements exist in Africa, and long-term results presented in this study are the most extensive for the studied ecosystems.
Article
Full-text available
The MIXv2 Asian emission inventory is developed under the framework of the Model Inter-Comparison Study for Asia (MICS-Asia) Phase IV and produced from a mosaic of up-to-date regional emission inventories. We estimated the emissions for anthropogenic and biomass burning sources covering 23 countries and regions in East, Southeast and South Asia and aggregated emissions to a uniform spatial and temporal resolution for seven sectors: power, industry, residential, transportation, agriculture, open biomass burning and shipping. Compared to MIXv1, we extended the dataset to 2010–2017, included emissions of open biomass burning and shipping, and provided model-ready emissions of SAPRC99, SAPRC07, and CB05. A series of unit-based point source information was incorporated covering power plants in China and India. A consistent speciation framework for non-methane volatile organic compounds (NMVOCs) was applied to develop emissions by three chemical mechanisms. The total Asian emissions for anthropogenic/open biomass sectors in 2017 are estimated as follows: 41.6/1.1 Tg NOx, 33.2/0.1 Tg SO2, 258.2/20.6 Tg CO, 61.8/8.2 Tg NMVOC, 28.3/0.3 Tg NH3, 24.0/2.6 Tg PM10, 16.7/2.0 Tg PM2.5, 2.7/0.1 Tg BC (black carbon), 5.3/0.9 Tg OC (organic carbon), and 18.0/0.4 Pg CO2. The contributions of India and Southeast Asia were emerging in Asia during 2010–2017, especially for SO2, NH3 and particulate matter. Gridded emissions at a spatial resolution of 0.1° with monthly variations are now publicly available. This updated long-term emission mosaic inventory is ready to facilitate air quality and climate model simulations, as well as policymaking and associated analyses.
Preprint
Full-text available
We analyze tropical ozone (O3) and carbon monoxide (CO) distributions in the upper troposphere (UT) and their temporal changes for 2005–2020 using Aura Microwave Limb Sounder (MLS) observations and chemistry climate models. The models are the Whole Atmosphere Community Climate Model (WACCM6) and two variants of the Community Atmosphere Model with Chemistry (CAM-chem), each variant using different anthropogenic emissions. Upper tropospheric trends and variability diagnostics are obtained from multiple linear regression analyses. We compare the model and MLS annual climatologies, focusing on 147 and 215 hPa pressure levels; climatological values generally fall within 10–20 % of each other, with both positive and negative differences for O3, and with models generally underestimating observed CO. In the northern hemisphere tropics, we find significantly poorer model fits to the observed phasing of CO seasonal changes at 215 hPa than at 147 hPa. This discrepancy is much smaller for the comparison of modeled and Measurements of Pollution in the Troposphere (MOPITT) V9J CO columns. We also find that the sensitivity of UT CO to El Niño / Southern Oscillation (ENSO) is positive at all tropical longitudes, in contrast to the dipolar longitudinal structure that exists for UT O3 ENSO sensitivity. MLS O3 has a zonal mean trend at 20° S–20° N of +0.39 ± 0.28 % yr-1; CAM-chem and WACCM have similar trends, though the WACCM trend is somewhat smaller. Our analyses for specific latitude/longitude bins yield positive trends up to 1.4 % yr-1 over Indonesia and East of that region, as well as over tropical Africa and the tropical Atlantic. We find broad similarities between the mapped MLS-derived UT O3 trends and corresponding mapped trends of tropospheric column ozone. Positive tropical UT mapped O3 trends are generally captured by the models, although in a more muted way. There is room for improvements in modeled tropical UT CO trends. Indeed, the MLS zonal mean CO trend is -0.25 ± 0.30 % yr-1, whereas the corresponding modeled CO trends are near zero (0.0 ± 0.14 % yr-1) when the anthropogenic emissions used in CAM-chem and WACCM are taken from Community Emissions Data System (CEDS) version 2. The non-CEDS version of CAM-chem yields CO UT trends of 0.22 ± 0.19 % yr-1, in contrast to the negative MLS CO trends throughout the tropics. The negative MLS tropical UT CO trends for 2005–2020 agree with (but tend to be smaller in magnitude than) previously published total column CO trends. Decreasing CO emissions from anthropogenic and biomass burning sources have previously been suggested as the main causes of tropospheric CO decreases, although significant regional emission trend variations exist.
Article
Full-text available
Observations and models demonstrate that ozone and its precursors can be transported between continents and across oceans. We model the influences of 10% reductions in anthropogenic nitrogen oxide (NOx) emissions from each of nine world regions on surface ozone air quality in that region and all other regions. In doing so, we quantify the relative importance of long-range transport between all source-receptor pairs, for direct short-term ozone changes. We find that for population-weighted concentrations during the three-month "ozone-season", the strongest inter-regional influences are from Europe to the Former Soviet Union, East Asia to Southeast Asia, and Europe to Africa. The largest influences per unit of NOx reduced, however, are seen for source regions in the tropics and Southern Hemisphere, which we attribute mainly to greater sensitivity to changes in NOx in the lower troposphere, and secondarily to increased vertical convection to the free troposphere in tropical regions, allowing pollutants to be transported further. Results show, for example, that NOx reductions in North America are ~20% as effective per unit NOx in reducing ozone in Europe during summer, as NOx reductions from Europe itself. Reducing anthropogenic emissions of non-methane volatile organic compounds (NMVOCs) and carbon monoxide (CO) by 10% in selected regions, can have as large an impact on long-range ozone transport as NOx reductions, depending on the source region. We find that for many source-receptor pairs, the season of greatest long-range influence does not coincide with the season when ozone is highest in the receptor region. Reducing NOx emissions in most source regions causes a larger decrease in export of ozone from the source region than in ozone production outside of the source region.
Article
Full-text available
Sulfur aerosols impact human health, ecosystems, agriculture, and global and regional climate. A new annual estimate of anthropogenic global and regional sulfur dioxide emissions has been constructed spanning the period 1850–2005 using a bottom-up mass balance method, calibrated to country-level inventory data. Global emissions peaked in the early 1970s and decreased until 2000, with an increase in recent years due to increased emissions in China, international shipping, and developing countries in general. An uncertainty analysis was conducted including both random and systemic uncertainties. The overall global uncertainty in sulfur dioxide emissions is relatively small, but regional uncertainties ranged up to 30%. The largest contributors to uncertainty at present are emissions from China and international shipping. Emissions were distributed on a 0.5° grid by sector for use in coordinated climate model experiments.
Article
Full-text available
The Model of Emissions of Gases and Aerosols from Nature version 2.1 (MEGAN2.1) is a modeling framework for estimating fluxes of 147 biogenic compounds between terrestrial ecosystems and the atmosphere using simple mechanistic algorithms to account for the major known processes controlling biogenic emissions. It is available as an offline code and has also been coupled into land surface models and atmospheric chemistry models. MEGAN2.1 is an update from the previous versions including MEGAN2.0 for isoprene emissions and MEGAN2.04, which estimates emissions of 138 compounds. Isoprene comprises about half of the estimated total global biogenic volatile organic compound (BVOC) emission of 1 Pg (1000 Tg or 10<sup>15</sup> g). Another 10 compounds including methanol, ethanol, acetaldehyde, acetone, α-pinene, β-pinene, t −β-ocimene, limonene, ethene, and propene together contribute another 30% of the estimated emission. An additional 20 compounds (mostly terpenoids) are associated with another 17% of the total emission with the remaining 3% distributed among 125 compounds. Emissions of 41 monoterpenes and 32 sesquiterpenes together comprise about 15% and 3%, respectively, of the total global BVOC emission. Tropical trees cover about 18% of the global land surface and are estimated to be responsible for 60% of terpenoid emissions and 48% of other VOC emissions. Other trees cover about the same area but are estimated to contribute only about 10% of total emissions. The magnitude of the emissions estimated with MEGAN2.1 are within the range of estimates reported using other approaches and much of the differences between reported values can be attributed to landcover and meteorological driving variables. The offline version of MEGAN2.1 source code and driving variables is available from http://acd.ucar.edu/~guenther/MEGAN/MEGAN.htm and the version integrated into the Community Land Model version 4 (CLM4) can be downloaded from http://www.cesm.ucar.edu/ .
Article
Full-text available
Non-methane volatile organic compounds (NMVOCs) influence air quality and global climate change through their effects on secondary air pollutants and climate forcers. Here we simulate the air quality and radiative forcing (RF) impacts of changes in ozone, methane, and sulfate from halving anthropogenic NMVOC emissions globally and from 10 regions individually, using a global chemical transport model and a standalone radiative transfer model. Halving global NMVOC emissions decreases global annual average tropospheric methane and ozone by 36.6 ppbv and 3.3 Tg, respectively, and surface ozone by 0.67 ppbv. All regional reductions slow the production of PAN, resulting in regional to intercontinental PAN decreases and regional NOx increases. These NOx increases drive tropospheric ozone increases nearby or downwind of source regions in the Southern Hemisphere (South America, Southeast Asia, Africa, and Australia). Some regions' NMVOC emissions contribute importantly to air pollution in other regions, such as East Asia, Middle East, and Europe, whose impact on US surface ozone is 43%, 34%, and 34% of North America's impact. Global and regional NMVOC reductions produce widespread negative net RFs (cooling) across both hemispheres from tropospheric ozone and methane decreases, and regional warming and cooling from changes in tropospheric ozone and sulfate (via several oxidation pathways). The total global net RF for NMVOCs is estimated as 0.0277 W m−2 (~1.8% of CO2 RF since the preindustrial). The 100 yr and 20 yr global warming potentials (GWP100, GWP20) are 2.36 and 5.83 for the global reduction, and 0.079 to 6.05 and −1.13 to 18.9 among the 10 regions. The NMVOC RF and GWP estimates are generally lower than previously modeled estimates, due to differences among models in ozone, methane, and sulfate sensitivities, and the climate forcings included in each estimate. Accounting for a~fuller set of RF contributions may change the relative magnitude of each region's impacts. The large variability in the RF and GWP of NMVOCs among regions suggest that regionally-specific metrics may be necessary to include NMVOCs in multi-gas climate trading schemes.
Article
Full-text available
Carbon monoxide (CO) emissions influence global and regional air quality and global climate change by affecting atmospheric oxidants and secondary species. We simulate the influence of halving anthropogenic CO emissions globally and individually from 10 regions on surface and tropospheric ozone, methane, and aerosol concentrations using a global chemical transport model (MOZART-4 for the year 2005). Net radiative forcing (RF) is then estimated using the GFDL (Geophysical Fluid Dynamics Laboratory) standalone radiative transfer model. We estimate that halving global CO emissions decreases global annual average concentrations of surface ozone by 0.45 ppbv, tropospheric methane by 73 ppbv, and global annual net RF by 36.1 mW m−2, nearly equal to the sum of changes from the 10 regional reductions. Global annual net RF per unit change in emissions and the 100 yr global warming potential (GWP100) are estimated as −0.124 mW m−2 (Tg CO)−1 and 1.34, respectively, for the global CO reduction, and ranging from −0.115 to −0.131 mW m−2 (Tg CO)−1 and 1.26 to 1.44 across 10 regions, with the greatest sensitivities for regions in the tropics. The net RF distributions show widespread cooling corresponding to the O3 and CH4 decreases, and localized positive and negative net RFs due to changes in aerosols. The strongest annual net RF impacts occur within the tropics (28° S–28° N) followed by the northern midlatitudes (28° N–60° N), independent of reduction region, while the greatest changes in surface CO and ozone concentrations occur within the reduction region. Some regional reductions strongly influence the air quality in other regions, such as East Asia, which has an impact on US surface ozone that is 93% of that from North America. Changes in the transport of CO and downwind ozone production clearly exceed the direct export of ozone from each reduction region. The small variation in CO GWPs among world regions suggests that future international climate agreements could adopt a globally uniform metric for CO with little error, or could use different GWPs for each continent. Doing so may increase the incentive to reduce CO through coordinated policies addressing climate and air quality.
Article
Full-text available
Present day tropospheric ozone and its changes between 1850 and 2100 are considered, analysing 15 global models that participated in the Atmospheric Chemistry and Climate Model Intercomparison Project (ACCMIP). The multi-model mean compares well against present day observations. The seasonal cycle correlates well, except for some locations in the tropical upper troposphere. Most (75%) of the models are encompassed with a range of global mean tropospheric ozone column estimates from satellite data, although there is a suggestion of a high bias in the Northern Hemisphere and a low bias in the Southern Hemisphere. Compared to the present day multi-model mean tropospheric ozone burden of 337 Tg, the multi-model mean burden for 1850 time slice is ~ 30% lower. Future changes were modelled using emissions and climate projections from four Representative Concentration Pathways (RCPs). Compared to 2000, the relative changes for the tropospheric ozone burden in 2030 (2100) for the different RCPs are: −5% (−22%) for RCP2.6, 3% (−8%) for RCP4.5, 0% (−9%) for RCP6.0, and 5% (15%) for RCP8.5. Model agreement on the magnitude of the change is greatest for larger changes. Reductions in precursor emissions are common across the RCPs and drive ozone decreases in all but RCP8.5, where doubled methane and a larger stratospheric influx increase ozone. Models with high ozone abundances for the present day also have high ozone levels for the other time slices, but there are no models consistently predicting large or small changes. Spatial patterns of ozone changes are well correlated across most models, but are notably different for models without time evolving stratospheric ozone concentrations. A unified approach to ozone budget specifications is recommended to help future studies attribute ozone changes and inter-model differences more clearly.
Article
Full-text available
A global climatology of tropospheric and stratospheric column ozone is derived by combining six years of Aura Ozone Monitoring Instrument (OMI) and Microwave Limb Sounder (MLS) ozone measurements for the period October 2004 through December 2010. The OMI/MLS tropospheric ozone climatology exhibits large temporal and spatial variability which includes ozone accumulation zones in the tropical south Atlantic year-round and in the subtropical Mediterranean/Asia region in summer months. High levels of tropospheric ozone in the Northern Hemisphere also persist in mid-latitudes over the eastern part of the North American continent extending across the Atlantic Ocean and the eastern part of the Asian continent extending across the Pacific Ocean. For stratospheric ozone climatology from MLS, largest column abundance is in the Northern Hemisphere in the latitude range 70° N–80° N in February–April and in the Southern Hemisphere around 40° S–50° S during August–October. Largest stratospheric ozone lies in the Northern Hemisphere and extends from the eastern Asian continent eastward across the Pacific Ocean and North America. With the advent of many newly developing 3-D chemistry and transport models it is advantageous to have such a dataset for evaluating the performance of the models in relation to dynamical and photochemical processes controlling the ozone distributions in the troposphere and stratosphere. The OMI/MLS gridded ozone climatology data are made available to the science community via the NASA Goddard Space Flight Center ozone and air quality website http://ozoneaq.gsfc.nasa.gov/ .
Article
Full-text available
Nitrogen oxides (NOx=NO+NO2) are produced during combustion processes and, thus may serve as a proxy for fossil fuel-based energy usage and coemitted greenhouse gases and other pollutants. We use high-resolution nitrogen dioxide (NO2) data from the Ozone Monitoring Instrument (OMI) to analyze changes in urban NO2 levels around the world from 2005 to 2014, finding complex heterogeneity in the changes. We discuss several potential factors that seem to determine these NOx changes. First, environmental regulations resulted in large decreases. The only large increases in the United States may be associated with three areas of intensive energy activity. Second, elevated NO2 levels were observed over many Asian, tropical, and subtropical cities that experienced rapid economic growth. Two of the largest increases occurred over recently expanded petrochemical complexes in Jamnagar (India) and Daesan (Korea). Third, pollution transport from China possibly influenced the Republic of Korea and Japan, diminishing the impact of local pollution controls. However, in China, there were large decreases over Beijing, Shanghai, and the Pearl River Delta, which were likely associated with local emission control efforts. Fourth, civil unrest and its effect on energy usage may have resulted in lower NO2 levels in Libya, Iraq, and Syria. Fifth, spatial heterogeneity within several megacities may reflect mixed efforts to cope with air quality degradation. We also show the potential of high-resolution data for identifying NOx emission sources in regions with a complex mix of sources. Intensive monitoring of the world's tropical/subtropical megacities will remain a priority, as their populations and emissions of pollutants and greenhouse gases are expected to increase significantly.
Article
Full-text available
The European Research Infrastructure IAGOS (In-service Aircraft for a Global Observing System) operates a global-scale monitoring system for atmospheric trace gases, aerosols and clouds utilising the existing global civil aircraft. This new monitoring infrastructure builds on the heritage of the former research projects MOZAIC (Measurement of Ozone and Water Vapour on Airbus In-service Aircraft) and CARIBIC (Civil Aircraft for the Regular Investigation of the Atmosphere Based on an Instrument Container). CARIBIC continues within IAGOS and acts as an important airborne measurement reference standard within the wider IAGOS fleet. IAGOS is a major contributor to the in-situ component of the Copernicus Atmosphere Monitoring Service (CAMS), the successor to the Global Monitoring for the Environment and Security – Atmospheric Service, and is providing data for users in science, weather services and atmospherically relevant policy. IAGOS is unique in collecting regular in-situ observations of reactive gases, greenhouse gases and aerosol concentrations in the upper troposphere and lowermost stratosphere (UTLS) at high spatial resolution. It also provides routine vertical profiles of these species in the troposphere over continental sites or regions, many of which are undersampled by other networks or sampling studies, particularly in Africa, Southeast Asia and South America. In combination with MOZAIC and CARIBIC, IAGOS has provided long-term observations of atmospheric chemical composition in the UTLS since 1994. The longest time series are 20 yr of temperature, H2O and O3, and 9–15 yr of aerosols, CO, NO y , CO2, CH4, N2O, SF6, Hg, acetone, ~30 HFCs and ~20 non-methane hydrocarbons. Among the scientific highlights which have emerged from these data sets are observations of extreme concentrations of O3 and CO over the Pacific basin that have never or rarely been recorded over the Atlantic region for the past 12 yr; detailed information on the temporal and regional distributions of O3, CO, H2O, NO y and aerosol particles in the UTLS, including the impacts of cross-tropopause transport, deep convection and lightning on the distribution of these species; characterisation of ice-supersaturated regions in the UTLS; and finally, improved understanding of the spatial distribution of upper tropospheric humidity including the finding that the UTLS is much more humid than previously assumed.
Article
A new method for evaluating the effectiveness of reactive organic gas (ROG) against nitrogen oxide (NOx) controls for reducing ambient ozone (O3) concentrations is presented. The method is based on ambient measurements of `photochemical indicator' species associated with O3-NOy-ROG chemistry. The indicator measurements provide a direct estimate for the effectiveness of ROG and NOx controls to evaluate the accuracy of photochemical models used to establish ozone control policy. Urban airshed model simulations for Atlanta, Georgia; New York, New York; and Los Angeles, California demonstrated that high values of investigated indicators species ratios are correlated with NOx-sensitive chemistry and low values are associated with ROG-sensitive chemistry.