You are on page 1of 4

PERSPECTIVE

PUBLISHED ONLINE: 8 AUGUST 2016 | DOI: 10.1038/NGEO2771

Role of atmospheric chemistry in the climate


impacts of stratospheric volcanic injections
Allegra N. LeGrande1*, Kostas Tsigaridis1,2* and Susanne E. Bauer1,2
The climate impact of a volcanic eruption is known to be dependent on the size, location and timing of the eruption. However, the
chemistry and composition of the volcanic plume also control its impact on climate. It is not just sulfur dioxide gas, but also the
coincident emissions of water, halogens and ash that influence the radiative and climate forcing of an eruption. Improvements
in the capability of models to capture aerosol microphysics, and the inclusion of chemistry and aerosol microphysics modules
in Earth system models, allow us to evaluate the interaction of composition and chemistry within volcanic plumes in a new
way. These modelling efforts also illustrate the role of water vapour in controlling the chemical evolution — and hence climate
impacts — of the plume. A growing realization of the importance of the chemical composition of volcanic plumes is leading to
a more sophisticated and realistic representation of volcanic forcing in climate simulations, which in turn aids in reconciling
simulations and proxy reconstructions of the climate impacts of past volcanic eruptions. More sophisticated simulations are
expected to help, eventually, with predictions of the impact on the Earth system of any future large volcanic eruptions.

V
olcanic eruptions fascinate people, not only because of the Volcanic aerosols and atmospheric chemistry
tremors, glowing cascades of ash and rivers of lava, but also Studies on the climate impacts of sulfate aerosols2–4 occurred in
because of the regional disruptions they cause, including ash parallel with research into associated changes in atmospheric
deposition and ‘dry fogs’ — airborne masses of sulfate aerosols that chemistry 16,17. Atmospheric water has long been known to be vital
do not feel wet like normal fogs. Sulfuric acid aerosol particles have in contributing to the formation rate of sulfate aerosols from SO2
long been suspected of creating climate disturbances. As far back emissions18, and it has been hypothesized to be a limiting factor
as the eighteenth century, Benjamin Franklin conjectured that the in volcanic aerosol formation19, specifically through the creation
1783–1784 eruption of the Icelandic Laki volcano had been the cause of OH radicals from water. Modelling experiments based on the
of several cold and harsh winters1. 1991 Mount Pinatubo eruption and on the Toba eruption about
Mount Agung in Indonesia erupted in 1963, reinvigorating inter- 74,000 years ago20, which was several orders of magnitude larger,
est in the principal mechanism through which volcanic eruptions indicate a potential for the creation of OH radicals to accelerate
affect climate. Initially, the most explosive volcanoes with the great- the conversion of SO2 gas into sulfate aerosols19. Sulfur dioxide is
est amount of ash emission were suspected of having the greatest a greenhouse gas that may temper the negative forcing of sulfate
influence on climate2,3. Later analysis using radiative transfer calcu- aerosols21, which makes the balance between SO2 and sulfate a
lations revealed that the volume of sulfur dioxide (SO2) injected into significant controller of the climate response following an erup-
the atmosphere, and its subsequent conversion into sulfate aerosols tion. But it is not certain that a state in which water was depleted
in the stratosphere, is the principal cause of global-scale cooling fol- through the production of OH radicals could exist — not only
lowing an eruption4,5. This insight has been confirmed by numerous because the stratospheric warming from such an aerosol layer 22
studies using measurements from the TOMS and SAGE II satellite would warm the tropopause (the extremely cold boundary which
instruments and their descendants6–9, and was further borne out by removes moisture) and allow in more moisture, but also because
the eruption of Mount Pinatubo in 1991, which released huge quan- volcanoes do not exclusively inject SO2 (ref. 16).
tities of SO2. Consequently, global surface temperatures dropped by Water, in addition to converting SO2 into sulfate, affects the
half a degree Celsius10, and tropical and monsoonal rainfall decreased oxidizing capacity of the stratosphere and influences ozone (O3)
markedly 11. This event’s forcing at its peak exceeded 6 W m–2 (simi- concentration23. Modulating ozone concentrations can cause either
lar in magnitude but opposite in sign to anthropogenic greenhouse positive or negative forcing, depending on the chemical state of the
gas forcing)6. stratosphere, including reactive nitrogen and halogen levels. The
Volcanic eruptions also influence the variability of climate modes, stratosphere is practically dry when compared with the troposphere,
for instance in causing the onset of positive North Atlantic Oscillation and this demands a closer look at the composition of the volcanic
conditions12. For the twentieth century, eruptions were coincident injection itself, since an eruption can co-inject water 24 and halogens
with establishment of positive phases of the El Niño/Southern directly into the stratosphere16. The role of injected water from an
Oscillation (El Niño events)13; in the past, they were instead, para- eruption whose plume stays in the troposphere is not expected to be
doxically, associated with negative phases14. Thus the link between significant, owing to the abundance of water vapour below the trop-
the occurrence of El Niño events and volcanic eruptions remains opause, but it could be important for eruptions reaching the strato-
uncertain. Eruptions have been found to be probable disruptors sphere. Note that here we use the term ‘plume’ to cover the various
of deep water circulation in the Atlantic15. The exact mechanisms phases of the volcanic ejecta, and the evolution and interaction of
through which these changes occur are not yet fully understood. these emissions in the climate system.

NASA Goddard Institute for Space Studies, 2880 Broadway New York, New York, USA. 2Center for Climate Systems Research, Columbia University, 2880
1

Broadway New York, New York, USA. *e-mail: allegra.n.legrande@nasa.gov; kostas.tsigaridis@columbia.edu

NATURE GEOSCIENCE | ADVANCE ONLINE PUBLICATION | www.nature.com/naturegeoscience 1


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PERSPECTIVE NATURE GEOSCIENCE DOI: 10.1038/NGEO2771

SW

Halogens

Sulfate

Sulfate

SO SO2
2 +
Ash

H2
O
Tropopause LW
H2O (g)

Time

Figure 1 | Volcanic water in the stratosphere alters aerosol evolution. The injection of stratospheric water along with volcanic SO2 (blue band) yields a
much enhanced rate of sulfate formation relative to SO2 injection alone (red band). Water increases the availability of OH radicals, converting SO2 more
quickly into sulfate aerosols (green spheres) and increasing the rate of aerosol growth. While these aerosols reflect shortwave radiation (SW) from
space, leading to cooling, sulfate aerosols may also scatter longwave radiation (LW) from the Earth, promoting warming. The nature, formation rate and
abundance of the aerosols formed will control the regional and global climate response following a volcanic eruption. The different altitudes of the two
bands is for illustration purposes only.

The water-limited stratospheric chemistry complicates the sim- again externally specified) aerosol layers with prescribed optical
ple climate-impact story of a thermal warming in the stratosphere thickness of the aerosol layer. The earlier efforts to understand the
and a shortwave cooling in the troposphere as a direct response potential emissions from volcanic events (and their significance to
to sulfate aerosol formation. For instance, stratospheric warming atmospheric chemistry and radiative forcing) were, unfortunately,
increases the temperature of the tropopause, allowing more mois- left out of these parameterizations.
ture to reach the stratosphere from below 22, which has the potential As a consequence, it is not surprising that notable mismatches
to dampen the negative solar forcing from sulfate aerosols25. This existed between model simulations of climate and either direct or
dynamical water source in the lower stratosphere can also reduce proxy measures of climate30,31. The most obvious was the models’
the water limitation of the SO2 oxidation and enhance aerosol for- response to the prescribed aerosol forcing, which was too large
mation rates. relative to the cooling implied by data compilations31. These mis-
These intricacies within the basic mechanics of volcanoes’ matches have reinvigorated the interest in past climate records of
capacity to influence climate were reasonably well understood by responses to volcanic events, and encouraged the modelling com-
the mid-1990s. The potential roles of ash, water and even halo- munity to take a closer look at the evolution of the chemistry and
gens as confounding factors to SO2 were also recognized decades physics of a volcanic plume32. From the data side, more rigorous
ago4. However, three-dimensional general circulation models were investigation of the model forcing 33 and climate response34 has been
not yet capable of handling the chemistry, aerosols and radiative initiated as well. These studies have served to reduce the uncertainty
impacts of volcanic injections simultaneously. For this, parameter- in the forcing and climate response, and hence enhance our ability
izations were developed to provide external forcing to the models to test the skill of chemistry and aerosol modules when including
for the historical period22, and the chemistry impacts and feedbacks the effects of volcanoes in models.
were not implemented. Probably owing to the inability to model the For atmospheric chemistry and climate modellers, it is an
chemistry, the interest and development in improving the way that exciting time to be involved in the research on volcanic impacts
we represent volcanoes in climate models scaled back significantly on climate. Collaborative, comparative initiatives such as SSiRC
for the next decade. (Stratospheric Sulfur and its Role in Climate) and VolMIP (Volcano
Model Intercomparison Project, CMIP6) are seeking out smarter,
Pushing models to the brink more physically based ways to implement volcanic eruptions.
As part of the Coupled Model Intercomparison Project Phase 5 Similar to the advances in climate modelling resulting from the
(CMIP5), the palaeoclimate community coordinated a suite of coupling of dynamic ocean modules with sophisticated atmospheric
experiments to emulate the past millennium26. Throughout this schemes, the evolution of climate models to include a prognostic
period, the largest abrupt climate perturbations were related to vol- calculation of volcanic aerosols yields the potential to predict the
canic eruptions, and large numbers of volcanic eruptions over an impact of future volcanic eruptions on climate, once the magnitude
extended period were linked to intervals of centennial-scale cool- and location of the gaseous injection are constrained. Determining
ing 27. This study of the past millennium brought about renewed the impact of each constituent of volcanic eruptions will aid this
interest in volcanic modelling 28,29, but it lacked any improvements capacity for predicting the climate impact of future volcanic events.
to the way in which the effects of volcanoes were implemented. The We have capitalized on improvements to the aerosol microphysics35
model representation of volcanoes ranged from externally specified of the NASA Goddard Institute for Space Studies GISS ModelE36
top-of-the-atmosphere negative forcing to more sophisticated (but to make a preliminary assessment of how the joint injection of

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


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
NATURE GEOSCIENCE DOI: 10.1038/NGEO2771 PERSPECTIVE
volcanic SO2 and water vapour modulate stratospheric chemistry,
particularly the oxidizing capacity of the stratosphere and the con- 0.10

Global stratospheric mean


aerosol optical depth
version rate of SO2 to sulfate aerosol. 0.08

It’s all in the chemistry 0.06


Chemistry plays an important role in a volcanic plume: both directly, 0.04
by converting SO2 into sulfate aerosols; and indirectly, by influenc-
ing the OH abundance (oxidizing capacity) in the plume. Following 0.02 SO2-only injection
a volcanic injection of material into the stratosphere, the oxida- SO2 + H2O injection
0.00
tion of SO2 by OH radicals begins, briefly forming HOSO2 before Jul Aug Sep Oct
reacting with O2 to form SO3 and HO2 radicals37. In the presence
of water vapour, SO3 then forms sulfuric acid (H2SO4). Owing to Figure 2 | Evolution of stratospheric global-mean aerosol optical depth
its extremely low vapour pressure, the H2SO4 formed either nucle- following a volcanic eruption. Sulfate aerosols form faster in the presence
ates to form new particles, or condenses on pre-existing aerosols, of water vapour as a result of the higher abundance of OH radicals, which
increasing the sulfate aerosol mass. accelerates SO2 oxidation. The initial new particle formation rapidly creates
The concentration of water vapour controls nucleation, since a large abundance of very small sulfate aerosols which later coagulate.
the main mechanism of new particle formation is believed to begin The resulting reduced number of larger, but still small, particles explains
with binary H2SO4–H2O nucleation38, which is especially true under the brief drop in aerosol optical depth. Further coagulation, together with
stratospheric conditions. For simplicity, models form sulfuric acid in condensation, ultimately forms the accumulation-mode volcanic aerosol
a single step, directly from the SO2 oxidation by OH, which is inde- layer. By comparison, the SO2-only injection has much slower chemistry,
pendent of water concentration. Models with aerosol microphysi- which reduces the initial rate (first 3 months) of aerosol formation.
cal parameterizations then form sulfate aerosols either through new
particle formation (which is dependent on sulfuric acid concentra- and increases the OH concentration in the plume by a factor of ten
tion and relative humidity) or through condensation39. In contrast, (ref. 47) as a direct consequence of the O(1D) reaction with water
bulk aerosol models simply assume that the sulfuric acid formed is vapour. The faster stratospheric AOD formation during the first
in the aerosol phase40. Nucleation rates decrease exponentially with few months in the wet eruption agrees better with stratospheric
increasing temperature and decreasing relative humidity 38, and are AOD reconstructions8,49.
extremely low at typical stratospheric conditions, where relative
humidity is below 1%. Moving forward
Water also has an indirect impact on sulfate aerosol formation Volcanic eruptions could inject massive amounts of water into the
in the stratosphere through OH radicals. The main OH formation stratosphere. For the Toba eruption, a water release of as much
pathway in the atmosphere starts with ozone photolysis that pro- as 27 Pg (Pg, 1015 grams) has been estimated, although the mag-
duces excited atomic oxygen, O(1D) (ref. 41) which in the presence nitude is uncertain50. Such water injections have the potential to
of water vapour can react with H2O and form two OH radicals. The severely influence the rates of chemical reactions and of aerosol
production of OH radicals through this mechanism has two impor- formation. To quantify the effect of such injections in the water-
tant implications: first, it decreases the O3 formation potential of limited stratosphere, it is essential to couple gas-phase chemistry
the system by taking O(1D) out of the Chapman cycle42; second, it with aerosol microphysics in climate models, so that the climate
increases the SO2 oxidation by OH, accelerating the sulfate aerosol impact of such massive events, and their evolution in space and
formation rate. Thus, a stratospheric volcanic injection that included time, can be quantified.
both SO2 and H2O would result in a much faster production of sul- If such an eruption were to occur today, we have much improved
fate aerosol in the otherwise extremely dry stratospheric air because models for estimating its climate impact, and a better apprecia-
of significantly higher OH levels, resulting in both a faster decay rate tion of the importance of determining the volume not only of sul-
of SO2 and a faster formation of a sulfate aerosol layer, with impor- fur injection, but also of water and other ingredients including ash
tant climate implications (Fig. 1). and halogens. A growing awareness of the importance of including
The chemistry of the stratosphere will be affected as a whole atmospheric chemistry when assessing the climate impacts of vol-
by the OH formation following a large water injection, with the canic eruptions will help with better estimates in future events.
reduced methane lifetime due to the increased OH being one of
the most climate-relevant changes. The concentration of O3 will Received 29 March 2016; accepted 21 June 2016;
also be affected, and can either decrease or increase43–45, depend- published online 8 August 2016
ing on the halogen levels in the stratosphere and changes in
stratospheric dynamics46. References
The changes in gas-phase chemistry following a ‘wet’ erup- 1. Franklin, B. Meteorological imaginations and conjectures. Mem. Lit. Phil. Soc.
tion have global-scale implications. We tested this using the GISS Manchester 2, 357–361 (1785).
ModelE36 coupled with gas-phase chemistry and the aerosol 2. Humphreys, W. J. Physics of the Air (McGraw-Hill, 1940).
microphysical module MATRIX35. We simulated a Pinatubo-sized 3. Budyko, M. I. The effect of solar radiation variations on the climate of the
wet eruption with injections of ~18  Tg SO2 (Tg, 1012 grams; refs Earth. Tellus 21, 611–619 (1969).
4. Pollack, J. B. et al. Volcanic explosions and climatic change: a theoretical
7,47) and ~150  Tg H2O (refs 25,48). We determined that during assessment. J. Geophys. Res. 81, 1071–1083 (1976).
the month of the eruption, the stratospheric global-mean aerosol 5. Rampino, M. R. & Self, S. Sulphur-rich volcanic eruptions and stratospheric
optical depth (AOD) produced by the wet eruption is more than aerosols. Nature 310, 677–679 (1984).
double the mean for a ‘dry’ eruption that only injects SO2 (Fig. 2). 6. Stenchikov, G. L. et al. Radiative forcing from the 1991 Mount Pinatubo
A month later, the global-mean AOD is still ~40% higher in the wet volcanic eruption. J. Geophys. Res. 103, 13837 (1998).
eruption case. Two months later and beyond, the slower chemis- 7. Krueger, A. J. et al. Volcanic sulfur dioxide measurements from the total ozone
mapping spectrometer instruments. J. Geophys. Res. 100, 14057–14076 (1995).
try in the dry eruption catches up, and the simulated global-mean 8. Sato, M., Hansen, J. E., McCormick, M. P. & Pollack, J. B. Stratospheric aerosol
AODs between the two cases are very similar. A single wet erup- optical depths, 1850–1990. J. Geophys. Res. 98, 22987–22994 (1993).
tion injects an amount of water into the stratosphere comparable to 9. Carn, S. A., Clarisse, L. & Prata, A. J. Multi-decadal satellite measurements of
that formed by methane oxidation over the course of a full year 24, global volcanic degassing. J. Volcanol. Geotherm. Res. 311, 99–134 (2016).

NATURE GEOSCIENCE | ADVANCE ONLINE PUBLICATION | www.nature.com/naturegeoscience 3


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PERSPECTIVE NATURE GEOSCIENCE DOI: 10.1038/NGEO2771
10. Hansen, J. et al. In The Mount Pinatubo Eruption: Effects on the Atmosphere and 35. Bauer, S. E. et al. MATRIX (Multiconfiguration Aerosol TRacker of mIXing
Climate NATO ASI Series, Vol. I 42 (eds Fiocco, G., Fua, D. & Visconti, G.) state): an aerosol microphysical module for global atmospheric models.
233–272 (Springer, 1996). Atmos. Chem. Phys. 8, 6003–6035 (2008).
11. Winter, A. et al. Persistent drying in the tropics linked to natural forcing. 36. Schmidt, G. A. et al. Configuration and assessment of the GISS
Nature Commun. 6, 7627 (2015). ModelE2 contributions to the CMIP5 archive. J. Adv. Model. Earth Syst.
12. Shindell, D. T., Schmidt, G. A., Mann, M. E. & Faluvegi, G. Dynamic winter 6, 141–184 (2014).
climate response to large tropical volcanic eruptions since 1600. J. Geophys. Res. 37. Sander, S. P. et al. Chemical Kinetics and Photochemical Data for Use in
109, D05104 (2004). Atmospheric Studies Report No. 17 (NASA Jet Propulsion Laboratory, 2011).
13. Robock, A. Volcanic eruptions and climate. Rev. Geophys. 38, 191–219 (2000). 38. Vehkamäki, H. An improved parameterization for sulfuric acid–water
14. Wahl, E., Diaz, H. F., Smerdon, J. & Ammann, C. Late winter temperature nucleation rates for tropospheric and stratospheric conditions. J. Geophys. Res.
response to large tropical volcanic eruptions in temperate western 107, http://dx.doi.org/10.1029/2002jd002184 (2002).
North America: relationship to ENSO phases. Glob. Planet. Change 39. Mann, G. W. et al. Intercomparison and evaluation of global aerosol
122, 238–250 (2014). microphysical properties among AeroCom models of a range of complexity.
15. Pausata, F. S. R., Chafik, L., Caballero, R. & Battisti, D. S. Impacts of high- Atmos. Chem. Phys. 14, 4679–4713 (2014).
latitude volcanic eruptions on ENSO and AMOC. Proc. Natl Acad. Sci. USA, 40. Textor, C. et al. Analysis and quantification of the diversities of aerosol life
http://dx.doi.org/10.1073/pnas.1509153112 (2015). cycles within AeroCom. Atmos. Chem. Phys. 6, 1777–1813 (2006).
16. Coffey, M. T. Observations of the impact of volcanic activity on stratospheric 41. Seinfield, J. H. & Pandis, S. N. Atmospheric Chemistry and Physics: From Air
chemistry. J. Geophys. Res. 101, 6767–6780 (1996). Pollution to Climate Change 2nd edn (Wiley, 2006).
17. Turco, R. P., Whitten, R. C. & Toon, O. B. Stratospheric aerosols: observation 42. Chapman, S. A Theory of Upper-atmospheric Ozone (Edward Stanford, 1930).
and theory. Rev. Geophys. Space Phys. 20, 233–279 (1982). 43. Solomon, S. et al. The role of aerosol variations in anthropogenic ozone
18. Rampino, M. R. & Self, S. Volcanic winter and accelerated glaciation following depletion at northern midlatitudes. J. Geophys. Res. 101, 6713–6727 (1996).
the Toba super-eruption. Nature 359, 50–52 (1992). 44. Tie, X. & Brasseur, G. The response of stratospheric ozone to volcanic
19. Bekki, S. Oxidation of volcanic SO2: a sink for stratospheric OH and H2O. eruptions: Sensitivity to atmospheric chlorine loading. Geophys. Res. Lett.
Geophys. Res. Lett. 22, 913–916 (1995). 22, 3035–3038 (1995).
20. Oppenheimer, C. Limited global change due to the largest known Quaternary 45. Rozanov, E. V. Climate/chemistry effects of the Pinatubo volcanic eruption
eruption, Toba ~74 kyr BP? Quat. Sci. Rev. 21, 1593–1609 (2002). simulated by the UIUC stratosphere/troposphere GCM with interactive
21. Schmidt, A. et al. Selective environmental stress from sulphur emitted by photochemistry. J. Geophys. Res. 107, http://dx.doi.org/10.1029/
continental flood basalt eruptions. Nature Geosci. 9, 77–82 (2015). 2001jd000974 (2002).
22. Robock, A. et al. Did the Toba volcanic eruption of ~74 ka B.P. 46. Aquila, V., Oman, L. D., Stolarski, R., Douglass, A. R. & Newman, P. A. The
produce widespread glaciation? J. Geophys. Res. 114, http://dx.doi. response of ozone and nitrogen dioxide to the eruption of Mt. Pinatubo at
org/10.1029/2008jd011652 (2009). southern and northern midlatitudes. J. Atmos. Sci. 70, 894–900 (2013).
23. Stenke, A. & Grewe, V. Simulation of stratospheric water vapor trends: impact 47. Read, W. G., Froidevaux, L. & Waters, J. W. Microwave limb sounder
on stratospheric ozone chemistry. Atmos. Chem. Phys. 5, 1257–1272 (2005). measurement of stratospheric SO2 from the Mt. Pinatubo volcano. Geophys.
24. Glaze, L. S., Baloga, S. M. & Wilson, L. Transport of atmospheric water vapor Res. Lett. 20, 1299–1302 (1993).
by volcanic eruption columns. J. Geophys. Res. 102, 6099 (1997). 48. Nedoluha, G. E. et al. Increases in middle atmospheric water vapor as observed
25. Joshi, M. M. & Jones, G. S. The climatic effects of the direct injection of water by the Halogen Occultation Experiment and the ground-based Water Vapor
vapour into the stratosphere by large volcanic eruptions. Atmos. Chem. Phys. Millimeter-Wave Spectrometer from 1991 to 1997. J. Geophys. Res. 103,
9, 6109–6118 (2009). 3531–3543 (1998).
26. Schmidt, G. A. et al. Climate forcing reconstructions for use in PMIP 49. Vernier, J. P. et al. Major influence of tropical volcanic eruptions on the
simulations of the last millennium (v1.0). Geosci. Model Dev. 4, 33–45 (2011). stratospheric aerosol layer during the last decade. Geophys. Res. Lett.
27. McGregor, H. V. et al. Robust global ocean cooling trend for the pre-industrial 38, http://dx.doi.org/10.1029/2011gl047563 (2011).
Common Era. Nature Geosci. 8, 671–677 (2015). 50. Bekki, S. et al. The role of microphysical and chemical processes in
28. Timmreck, C. et al. Limited temperature response to the very large prolonging the climate forcing of the Toba eruption. Geophys. Res. Lett.
AD 1258 volcanic eruption. Geophys. Res. Lett. 36, http://dx.doi. 23, 2669–2672 (1996).
org/10.1029/2009GL040083 (2009).
29. Toohey, M., Krüger, K., Niemeier, U. & Timmreck, C. The influence of eruption Acknowledgements
season on the global aerosol evolution and radiative impact of tropical volcanic We thank NASA GISS for institutional support. We also thank the NASA MAP
eruptions. Atmos. Chem. Phys. 11, 12351–12367 (2011). programme for continued support. Resources supporting this work were provided by
30. Mann, M. E., Fuentes, J. D. & Rutherford, S. Underestimation of volcanic the NASA High-End Computing (HEC) Program through the NASA Center for Climate
cooling in tree‑ring‑based reconstructions of hemispheric temperatures. Simulation (NCCS) at Goddard Space Flight Center.
Nature Geosci. 5, 202–205 (2012).
31. Frank, D. C. et al. Ensemble reconstruction constraints on the global carbon Author contributions
cycle sensitivity to climate. Nature 463, 527–530 (2010). A.N.L. was inspired to seek out better aerosol microphysics modules by the mismatches
32. LeGrande, A. N. & Anchukaitis, K. J. Volcanic eruptions and climate. PAGES of simulated and inferred climate impacts for volcanoes in the CMIP5/PMIP3 last
23 46–47 (2015). millennium experiment. A.N.L. and K.T. conceived the work, performed the model
33. Sigl, M. et al. Timing and climate forcing of volcanic eruptions for the past simulations and analysed the results. All authors contributed to the text and the design of
2,500 years. Nature 523, 543–549 (2015). figures. K.T. and S.E.B. contributed expertise with the MATRIX model.
34. Esper, J., Büntgen, U., Luterbacher, J. & Krusic, P. J. Testing the hypothesis of
post-volcanic missing rings in temperature sensitive dendrochronological data. Competing financial interests
Dendrochronologia 31, 216–222 (2013). The authors declare no competing financial interests.

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


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.

You might also like