SlideShare a Scribd company logo
1 of 9
Download to read offline
Tales of volcanoes and El-Niño southern oscillations
with the oxygen isotope anomaly of sulfate aerosol
Robina Shaheena
, Mariana Abauanzaa
, Teresa L. Jacksona
, Justin McCabea,b
, Joel Savarinoc
, and Mark H. Thiemensa,1
a
Department of Chemistry and Biochemistry, University of California San Diego, La Jolla, CA 92093; b
Pacific Ridge School, Carlsbad, CA 92009;
and c
Laboratoire de Glaciologie et Géophysique de l’Environnement, 38402 St. Martin d’Heres, Grenoble, France
Edited by Karl K. Turekian, Yale University, North Haven, CT, and approved February 1, 2013 (received for review September 19, 2012)
The ability of sulfate aerosols to reflect solar radiation and simul-
taneously act as cloud condensation nuclei renders them central
players in the global climate system. The oxidation of S(IV) com-
pounds and their transport as stable S(VI) in the Earth’s system are
intricately linked to planetary scale processes, and precise charac-
terization of the overall process requires a detailed understanding
of the linkage between climate dynamics and the chemistry lead-
ing to the product sulfate. This paper reports a high-resolution, 22-y
(1980–2002) record of the oxygen-triple isotopic composition of
sulfate (SO4) aerosols retrieved from a snow pit at the South
Pole. Observed variation in the O-isotopic anomaly of SO4 aerosol
is linked to the ozone variation in the tropical upper troposphere/
lower stratosphere via the Ozone El-Niño Southern Oscillations
(ENSO) Index (OEI). Higher Δ17
O values (3.3‰, 4.5‰, and 4.2‰)
were observed during the three largest ENSO events of the past
2 decades. Volcanic events inject significant quantities of SO4 aero-
sol into the stratosphere, which are known to affect ENSO
strength by modulating stratospheric ozone levels (OEI = 6 and
Δ17
O = 3.3‰, OEI = 11 and Δ17
O = 4.5‰) and normal oxidative
pathways. Our high-resolution data indicated that Δ17
O of sulfate
aerosols can record extreme phases of naturally occurring climate
cycles, such as ENSOs, which couple variations in the ozone levels
in the atmosphere and the hydrosphere via temperature driven
changes in relative humidity levels. A longer term, higher resolu-
tion oxygen-triple isotope analysis of sulfate aerosols from ice
cores, encompassing more ENSO periods, is required to reconstruct
paleo-ENSO events and paleotropical ozone variations.
Pinatubo | El-Chichón | Cerro Hudson | Intertropical Convergence Zone
Sulfate aerosols affect climate systems by altering radiation
balance, temperature, precipitation, and atmospheric dy-
namics (1, 2). The overall effect of sulfate aerosols on the climate
is estimated to be net cooling (−2.0 ± 0.2 Wm−2
) (3). According to
the Intergovernmental Panel on Climate Change fourth assess-
ment report, aerosols are one of the largest sources of uncer-
tainties in climate prediction models due to their temporal and
spatial variability (4). Globally, anthropogenic sulfur exceeds
natural sources by a factor of 3 to 4 (35 Tg·y−1
vs. 10 Tg·y−1
from
1990 to 2000) (5–8). Occasionally, volcanoes emit large quantities
of sulfur dioxide (SO2) directly into the atmosphere. Pinatubo, for
example, released ∼30 Tg of SO2 (5). In the troposphere, SO2 has
an atmospheric residence time of ∼2–3 d, and it is oxidized to
SO4
2−
by homogeneous and heterogeneous pathways and re-
moved via wet and dry deposition (9). Gas phase oxidation of SO2
by OH and subsequent reactions with water vapor yield sulfuric
acid vapor [H2SO4(g)] (10). The oxidation of aqueous SO2 by O3
and H2O2 far exceeds gas phase rates and is pH-dependent (11,
12). The oxygen-triple isotopic composition of sulfate aerosols has
been demonstrated to be a useful diagnostic tool to distinguish
and quantify reaction pathways and to determine the paleoox-
idant levels on centennial to millennium (glacial period) time
scales to present-day environments (13–18). Tropospheric SO2
has a mass-dependent oxygen isotopic composition (δ17
O ≈ 0.52
δ18
O) due to rapid isotopic equilibration with water vapor (19),
which erases the source-derived oxygen isotopic signature. [The
delta (δ) values denote the relative deviation of the isotope ra-
tios 17
R= (17
O/16
O) and 18
R = (18
O/16
O) in a sample (Rs) with
respect to standard material (Rst) in permill (‰) (e.g., δ17
O
[‰] = [17
Rs/17
Rst −1] * 1,000). Isotope abundance or depletion
is measured with reference to a standard material (e.g., for oxygen
isotopes, reference material is Vienna Standard Mean Ocean Water
[VSMOW]).] S(IV) species (SO2aq, HSO3
−
, and SO3
−2
) are oxi-
dized to stable sulfate [S(VI)] via OH radicals, H2O2 and O3 (19,
20). The ozone molecule is a unique quantitative tracer of oxi-
dation reactions because it possesses the highest enrichment in
the heavier isotopes of oxygen (70–120‰) and oxygen isotope
anomaly (Δ17
O = 25–30‰). The anomalous oxygen isotopic dis-
tribution of ozone has been shown to be transferred to oxygen-
carrying molecules, such as SOx-NOx-ClOx-HOx (21–25) (Δ17
O ≈
δ17
O − 0.52 δ18
O, a mass-dependent process, has Δ17
O = 0; mass-
independent processes have Δ17
O ≠ 0). The positive Δ17
O of sulfate
derives from aqueous phase oxidation of SO2 by H2O2 and O3 via
Reaction 2–Reaction 3 and involves transfer of the isotopic anomaly
from the oxidant to the product sulfate (12, 26). All other sulfate
sources, including gas-phase oxidation by OH in the troposphere via
Reaction 1a and metal-catalyzed oxidation by atmospheric O2,
possess mass-dependent signatures, as verified by laboratory and
field measurements (12, 14, 27, 28). However, OH in the strato-
sphere has been suggested to possess an O-isotopic anomaly (28,
29), which can be transferred to sulfate produced in the stratosphere
via Reaction 1b. The magnitude of the transfer of the Δ17
O depends
on the relative contribution (Reaction 1–Reaction 3):
ðSO2 + OHÞtroposphere → HSO-
3
À
Δ17
O = 0  ‰
Á
[Reaction 1a]
ðSO2 + OHÞstratosphere + M → HSO-
3 + M
À
Δ17
O ≠ 0  ‰
Á
[Reaction 1b]
HSO−
3 + H2O2 → H2SO4
À
Δ17
O = 0:5 − 1  ‰
Á
+ H2O
[Reaction 2]
SO2−
3 + O3 → SO2−
4
À
Δ17
O = 8 − 9  ‰
Á
+ O2 [Reaction 3]
Tropospheric S(IV) oxidation by O3 (Reaction 3) is the only
significant mechanism producing sulfate Δ17
O values >1‰,
therefore, Δ17
O values greater than 1‰ quantitatively reflect
the relative contribution of O3 during sulfate formation (12).
S(IV) oxidized in the stratosphere acquires an anomalous sig-
nature via Reaction 1 to Reaction 3 (30). Once sulfate is formed,
the isotopic signature is stable and permanently preserved in the
aerosol. In addition to defining reaction pathways, both S and O
isotopic anomalies of sulfate aerosols may determine paleo-
volcanic activities and reflect their upper atmospheric chemistry
Author contributions: R.S., J.M., J.S., and M.H.T. designed research; R.S., M.A., and T.L.J.
performed research; R.S., M.A., T.L.J., J.M., J.S., and M.H.T. contributed new reagents/
analytic tools; R.S. analyzed data; and R.S. wrote the paper.
The authors declare no conflict of interest.
This article is a PNAS Direct Submission.
1
To whom correspondence should be addressed. E-mail: mthiemens@ucsd.edu.
This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.
1073/pnas.1213149110/-/DCSupplemental.
17662–17667 | PNAS | October 29, 2013 | vol. 110 | no. 44 www.pnas.org/cgi/doi/10.1073/pnas.1213149110
(28, 31, 32). It is argued that large volcanic eruptions in the
Anthropocene era are depleting stratospheric ozone by providing
surfaces for heterogeneous chemical reactions (33). These varia-
tions in O3 can modulate the dynamic of the tropical stratosphere
(quasibiennial oscillations); thus, volcanoes are considered re-
sponsible for the strongest El-Niño Southern Oscillations (ENSOs)
(34). In this work, we investigate whether oxygen-triple isotopic
composition of sulfate aerosols may serve as a unique fingerprint
of ozone chemistry to parse out the effect of volcanoes and ENSO
events based on their differing oxidative pathways. The polar ice
caps are nature’s best archive of Earth’s atmospheric history and
preserve a record of paleovolcanic activities (30, 35). A high-
resolution seasonal record of sulfate aerosols is used to reconstruct
atmospheric history on annual and decadal time scales. The
oxygen-triple isotope data of sulfate aerosols from Greenland
(monthly samples from July 1999–June 2000) have demonstrated
its potential to constrain the transport and oxidation history of
sulfate aerosols to polar regions (14). Here, we present a 22-y
seasonally resolved profile of major ions and oxygen-triple isotope
data of sulfate aerosols obtained from surface snow sampled in
a snow pit (1 m × 1 m) at the South Pole to help elucidate the
processes determining the observed global variation in sulfate
aerosol from 1980 to 2002, as well as the oxidation history of
S(IV), its transport to the South Pole, and its linkage with the
dynamics of the upper atmosphere. Our data on the oxygen-triple
isotope measurements of sulfate aerosol encompass three major
volcanic events of the century [El-Chichόn (17.3° N, 93.2° W,
1,205 m), Pinatubo (15.13° N, 120.35° E, 1,745 m), and Cerro
Hudson (45° S, 72° W, 1,905 m)] and three major ENSO events
(1982–1983, 1991–1992, and 1997–1998).
Results
A high-resolution temporal record (1980–2002) of sulfate aerosols
extracted from the snow pit at the South Pole and the associated
oxygen isotopic composition (δ17
O, δ18
O, and Δ17
O) are given in
Fig. 1 and Table S1. The sulfate concentration in composite sam-
ples (details provided in Materials and Methods) ranged from 36 to
165 parts per billion (ppb). The highest sulfate concentration was
observed in ice layers deposited in 1991 and 1992 (Fig. 1A). The
single oxygen isotope ratio of sulfate aerosol showed significant
variation, ranging from 1980 to 2002 (δ18
O = −2 to +12‰).
Increases in sulfate concentrations due to volcanic activities show
a corresponding decrease in δ18
O (average δ18
O = 3.1‰ and
2.3‰ in 1983 and 1992, respectively). The oxygen isotopic anomaly
(Δ17
O) varied from 0.4 to 4.5‰ (Fig. 1B). A high-resolution con-
centration profile of sulfate aerosols indicated that volcanic sulfate
layers from Pinatubo and Cerro Hudson eruptions were deposited
from 1991–1992 (36) and El-Chichón from 1982–1983 with corre-
sponding oxygen isotope anomaly of 4.5‰ and 3.3‰. The oxygen
isotopic anomaly in sulfate aerosols observed during three major
El-Nino events (ENSO-I = 1982–1983, ENSO-II = 1991–1992,
ENSO-III 1997–1998) and a moderate event (ENSO-IV = 1986–
1987) track the Ozone ENSO Index (OEI; Fig. 1B). An unusually
high enrichment in oxygen-triple isotopic composition (δ18
O
=12‰, Δ17
O = 4.1‰), along with a higher OEI of 6 is observed
in 1990 and labeled as an unknown event in Fig. 1A.
A plot of Δ17
O vs. δ18
O indicated a very weak (r2
= 0.2) inverse
correlation (Fig. 2). In Fig. 2, the maximum Δ17
O achievable via
tropospheric ozone and peroxide aqueous phase oxidation is
represented by red and green rectangles and gas phase oxidation
via stratospheric OH/HO2 is shown in blue. A high-resolution
(∼1-cm sampling interval) concentration measurement of major
ions [sulfate, methane sulfonic acid (MSA), nitrate, and chloride]
obtained from the snow pit indicated no significant relation to the
sulfate concentration (Fig. S1). A higher resolution sulfate con-
centration profile revealed two distinct peaks 1 and 2 (Fig. 3) with
a fivefold and 3.6-fold increase, respectively, in sulfate (SO4) con-
centration in ice layers deposited during 1992. Peak 3 appeared
during a volcanically quiescent period (1990) and showed an ap-
proximately fourfold increase in sulfate concentration (190 ppb)
compared with an average background value of 50 ppb observed in
this study. The persistent background during the study period is
defined as the time period when natural and anthropogenic emis-
sions of S compounds are maintained at a relatively quasi-steady
state (1999–2002). Sulfate from the El-Chichόn volcanic activity
indicated as peak 4 (1982–1983) did not appear as a sharp peak;
rather, it is more spread out in time. In the high-resolution profile,
an increase in non-sea salt (nss) SO4 (peak 3 in 1990) is potentially
associated with a significant increase in MSA, although they are not
temporally identical. MSA concentrations varied from 2 to 50 ppb,
with a maximum increase in1985–1986 and 1990 (Fig. 3).
Discussion
The dataset presented here is the longest (1980–2002) and maxi-
mum time-resolved record of chemical and isotopic composition
of sulfate aerosols retrieved at the South Pole, and it encom-
passes three major volcanic events (El-Chichόn, Pinatubo, and
Cerro Hudson eruptions in April 1982, June 1991, and October
1991, respectively). These volcanic activities introduced significant
quantities of SO2 (Pinatubo = 30 Tg SO2, Cerro Hudson = 10 Tg
SO2, and El-Chichόn = 7 Tg SO2) (5, 37, 38) into the stratosphere,
which can be observed as distinct sulfate peaks as labeled in Figs.
1A and 3. The volcanic sulfate aerosol showed less enrichment in
Fig. 1. (A) Oxygen isotopic composition (brown squares) and sulfate con-
centration profile (green diamonds) of composite aerosol sulfate samples
extracted from the snow pit (6 m high) at the South Pole, Antarctica. The
increase in sulfate concentration due to El-Chichón and Pinatubo + Cerro
Hudson is also shown. UE, unknown event. (B) Comparison of oxygen iso-
tope anomaly (red lines) and OEI (blue lines) obtained by Ziemke et al. (65)
from the deseasonalized trend in total O3 column measured at the equa-
torial Eastern and Western Pacific, an El-Niño region. Violet bars indicate
three major ENSO events: ENSO-I (1982–1983), ENSO-II (1991–1992), and
ENSO-III (1997–1998). A moderate event, ENSO-IV (1986–1987), is also shown.
(The scale of ENSO events is defined by the National Oceanic and Atmo-
spheric Administration and is available at www.cpc.noaa.gov/products/analysis_
monitoring/ensostuff/ensoyears.shtml).
Shaheen et al. PNAS | October 29, 2013 | vol. 110 | no. 44 | 17663
EARTH,ATMOSPHERIC,
ANDPLANETARYSCIENCES
SPECIALFEATURE
the heavy isotopes of oxygen (δ18
Oaverage = 2.6 ± 1‰) but pos-
sesses a mass-independent anomaly (El-Chichόn: Δ17
O =3.3 ‰
and Pinatubo + Cerro Hudson: Δ17
O = 4.5‰). The oxygen isotope
anomaly reported here for the composite Pinatubo and Cerro
Hudson sulfate sample in 1992 is similar to the previously reported
higher resolution signal for Pinatubo (δ18
O = 5.1–9.5‰, Δ17
O =
3.8–4.7‰) (28). Higher Δ17
O values (3.3–4.5‰) of volcanic sul-
fate and nonvolcanic ENSO events (3.8–4.2 ‰) compared with
the lower tropospheric Δ17
O values (0.4–1.6‰) (13, 18) indicate
the predominant role of stratospheric OH and HO2 radicals (30,
39). The concentration of these radicals in the stratosphere
depends on ozone, water vapor, CH4, and NOx concentrations
(40), and it is suggested to be 1.5 ± 0.3 × 106
molecules per cubic
centimeter in the tropics using the global chemistry transport
model (41). Numeric simulations have also indicated that the OH
in the stratosphere acquires an oxygen isotope anomaly (Δ17
O = 2–
40‰) by means of exchange with NOx (39, 42). The anomalous
signal of OH and HO2 in the stratosphere is preserved (39, 42) due
to the extremely low water content in the stratosphere (∼5–10 ppm
by volume) (43, 44), which is normally erased in the troposphere
due to rapid isotope exchange with water vapor (∼3% in the
tropics to 0.1% in the cold polar regions) (20).
The most striking feature of the present data (Fig. 1B) is that
variables OEI and Δ17
O of sulfate aerosol track each other during
the strong El- Niño events of 1982–1983 (ENSO-I), 1991–1992
(ENSO-II), and 1997–1998 (ENSO-III) as well as during the
moderate event of 1986–1987 (ENSO-IV). These ENSO time
periods are defined according to the National Oceanic and Atmo-
spheric Administration as sea surface temperature anomalies (El-
Niño = warm and La-Niña = cool) in the tropical Pacific (5° N–5° S,
120°–170° W) (www.cpc.noaa.gov/products/analysis_monitoring/
ensostuff/ensoyears.shtml). The OEI plotted in Fig. 1B, as dis-
cussed, is obtained from variation in ozone concentrations at
tropical latitudes. The OEI and Δ17
O data in Fig. 1B track each
other but are slightly shifted, which may derive from two factors.
First, the necessity to combine samples for the nitrate (45) and
sulfate measurements introduces a modest time uncertainty. The
average of the combined depth is used to specify sample time and
assumes a uniform sulfate distribution throughout that time period.
Consequently, there is an uncertainty in time of the sulfate peak,
which, at maximum, is a few months. Second, there are different
indices used to capture El-Niño events [e.g., OEI, Oceanic Niño
Index (ONI)], which use a variety of differing geophysical obser-
vations to document El-Niño events, and they are not necessarily
exactly temporally equivalent to one another. The comparison of
the oxygen isotopic anomaly of sulfate aerosols with the OEI reveals
that higher Δ17
O values are associated with elevated ozone column
densities measured by different satellites. The ENSO signal during
two earlier events, ENSO-I [1982–1983 (El Chichόn: OEI = 6.5,
Δ17
O = 3.3‰)] and ENSO-II [1991–1992 (Pinatubo and Cerro
Hudson: OEI = 6, Δ17
O = 4.5‰)], may have been confounded due
to the intense volcanic activities, which introduced, in addition to
the SO2, significant amounts of sulfate oxidized in the troposphere
with less Δ17
O, thus diluting the higher Δ17
O signal of S(IV)
oxidation in the stratosphere via OH radicals. The Δ17
O of sulfate
and OEI track each other fairly well despite higher concentrations
of volcanic sulfate. The ENSO events in volcanically quiescent
periods manifested a higher O-isotopic anomaly and OEI [stron-
gest ENSO-III (1997–1998): OEI = 10.8, Δ17
O= 4.2‰; moderate
ENSO-IV (1986–1987): OEI = 5, Δ17
O= 3.8‰]. A significant in-
crease of total ozone during ENSO-III (1997–1998) was accom-
panied by decreased precipitation, producing extensive forest fires
in Indonesia, Australia, and South America (46, 47). The influence
of ENSOs on total columnar ozone has been attributed to the
variation in the tropopause height driven by changes of tropical
deep convection and alteration of Brewer–Dobson circulation (14,
48). These observations suggest that the Δ17
O of sulfate aerosols
can be used to track moderate to strong ENSO events and the
variation in ozone concentration. To develop this possibility, the
following points must be addressed: Where does ozone-driven ox-
idation of S(IV) occur, and how is the observed variation in the
tropical upper tropospheric O3 (OEI) and ENSOs chemically as-
sociated with Δ17
O of sulfate aerosols? It is known that tropospheric
air enters the stratosphere principally in the tropics within the In-
tertropical Convergence Zone (ITCZ, a thin and dynamic region
along the equator separating trade winds between the Southern and
Northern Hemispheres) and transports poleward in the strato-
sphere as shown in Fig. 4 (11). The ITCZ thus links the troposphere
to the stratosphere and is an important corridor for the transport of
aerosol and trace gases to the stratosphere (49). The maximum
transport of water vapor also occurs in the ITCZ (44) in the vicinity
of the ENSO region, which also corresponds to the region of the
OEI measurements used in Fig. 1B. A comparison between
stratospheric water vapor and tropical sea surface temperatures
has demonstrated a strong correlation and an impact on the
strength of El-Niño (49). The influence of ENSOs on the total
column of ozone is also linked to the variation in the tropopause
height. Tropical deep convection and changes in Brewer–Dobson
circulation may account for the observed ozone enhancement
Fig. 2. Four-isotope plot shows Δ17
O and δ18
O of sulfate aerosols extracted
from the snow pit at the South Pole. A weak observed correlation indicates
mixing of various sulfates from different sources. Red and green rectangles
display variation in δ18
O and Δ17
O sources of sulfate. The blue rectangle is an
oxidation source with stratospheric (Strat.) OH/HO2 radicals (28, 39). The green
rectangle shows the range of pure hydrogen peroxide oxidation. The red square
denotes the value of atmospheric (Atm.) oxygen. Max., maximum. Primary
sulfate produced during fossil fuel combustion at high temperature has been
shown to possess δ18
O values close to the atmospheric oxygen (51); however,
sulfate produced during biomass burning showed a range of δ18
O values (50)
depending on biomass type. The observed dataset reflects the range of pro-
cesses contributing to the observed oxygen isotopic composition of sulfate.
-600 -500 -400 -300 -200 -100 0
0
10
20
30
40
50
60
0
50
100
150
200
250
300
1977 1980 1983 1986 1989 1992 1995 1998 2001
Depth (cm)
MSA(ppb)
SO4(ppb)
Year
SO4 MSA1
23
4
Fig. 3. Concentration profile (1977–2003) of nss sulfate aerosol and MSA
extracted from the snow pit samples at the South Pole. Peaks 1, 2, and 4
represent Pinatubo, Cerro Hudson, and El-Chichón volcanic sulfates, re-
spectively. Peak 3 represents an unknown event.
17664 | www.pnas.org/cgi/doi/10.1073/pnas.1213149110 Shaheen et al.
during ENSO periods (14, 48). Cumulatively, all components
needed to provide ENSO-related Δ17
O enrichments are present.
There is coexistence of increased ozone and water content and
a dynamic mechanism of transport of aerosol into the strato-
sphere with poleward migration. The ozone-induced oxidation of
S(IV) via OH radicals is likely stratospheric and occurs during
transport (Fig. 4). In this model, large sulfate concentrations are
not required, simply a preferential ozone oxidation channel that
provides the heavy isotope-enriched sulfate. A detailed model
and discussion are far beyond the scope of this paper, but the
isotope measurements suggest that such experiments are war-
ranted. Our data indicate that the O-isotope anomaly of sulfate is
sensitive to changes in atmospheric dynamics and is faithfully
preserved in this record. A quasibiennial oscillation signal in the
Δ17
O of nitrate extracted from the same set of aerosol samples
from the South Pole has shown a similar linkage (45).
In the coordinate system of Fig. 2, a source of sulfate with higher
δ18
O and low Δ17
O is required. An ozone-rich source would have
high δ18
O values and also high Δ17
O values. Various oxidants of
different oxygen isotopic composition (OH, HO2, H2O2, and O3)
can oxidize SO2 to SO4 (Fig. 2, red and green rectangles are
guidelines to demonstrate variations in δ18
O with the maximal
Δ17
O obtainable via aqueous phase O3 and H2O2 oxidation, and
the blue rectangle indicates stratospheric OH and HO2 radical
reactions, as well as the associated maximum isotopic anomaly).
Fig. 2 requires the presence of a specific source, but it must be
of higher δ18
O and low Δ17
O. Laboratory experiments, field
observations, and numeric simulation of sulfur oxidation (14, 28,
50, 51) have shown that the δ18
O of sulfate coupled with Δ17
O can be
used to distinguish between primary and secondary sulfates. Pri-
mary sulfate [i.e., S(VI) produced at the emission source] exhibits
a higher enrichment in δ18
O = 20–45‰ and Δ17
O = 0 (50–52).
Secondary sulfate is formed when SO2 is oxidized in the atmo-
sphere via homogeneous and heterogeneous pathways (51).
The δ18
O of secondary sulfate thus represents a juxtaposition
of the highly variable water isotopic signature derived during SO2
oxidation processes at varying latitudes and altitudes (10, 53, 54).
A moderate correlation between δ18
O of rainwater and δ18
O(SO4)
(55) also indicated the complexity of using δ18
O to predict sources
of sulfate in rainwater. Most primary sulfate is removed via wet
and dry deposition in the free troposphere; however, a fraction of
S(IV), carbonyl sulfide (OCS), and traces of S(VI) are transported to
the stratosphere from the ITCZ and at midlatitudes via deep
convection (7, 56), thus permitting gas phase oxidation via OH/
HO2 radicals in the stratosphere (30).
An unusual increase in SO4 concentration (∼100 ppb) and
oxygen isotope enrichment (δ18
O ∼12‰ and Δ17
O ∼3.7‰)
in 1990 are immediately followed by an increase in MSA (Fig. 3).
To explain this peak, we consider two possible scenarios. The
first is stratospheric volcanic emissions due to the presence of
a higher oxygen isotope anomaly in the sulfate. Total ozone map-
ping spectrometer (TOMS) and stratospheric aerosol and gas
experiment (SAGE) satellite data indicated no significant in-
crease in the global inventory of stratospheric aerosols loading
during 1989–1990 (57), ruling out stratospheric volcanic emis-
sions, and this leaves us with one option, a local sulfate source
such as Mount Erebus (167° 25′ E, 77° 30′ S). The Smithsonian
database reports a significant increase in SO2 emissions, up to
100 t/d during this period, which could be a potential source of
SO4 peak (www.volcano.si.edu/reports/bulletin/contents.cfm?
issue=3609). Volcanic emissions, despite their complexity, change
mostly SO4 concentration (58). The second possible scenario
is that biogenic sources, such as phytoplankton, emit DMS, which,
on oxidation with O3, H2O2, NOx, or ClOx, can produce MSA,
and ultimately an increase in nss-SO4 (59, 60). This observed
spike in MSA temporally followed by an increase in nss-SO4 is
consistent with a biological source. Higher resolution measure-
ments of MSA on the high Antarctic Plateau (61) (both inland
and coastal sites) indicated postdepositional losses of MSA and
nss-SO4 production via MSA oxidation. Future sulfur isotope
measurements of sulfate aerosols in this time period may help to
elucidate and quantify this very specific source and oxidation
process further.
Conclusion
The most significant observation of the sulfate multioxygen iso-
topic record reported here is that observed trends in the mass-
independent O-isotopic anomaly are apparently linked to the
Fig. 4. Schematic depicts transport and transformation of sulfur species into the stratosphere and deposition of aged sulfate aerosol in the ice at the South
Pole. The red-shaded area indicates a significant contribution of SO2 and SO4 aerosols to the SSA in the lower stratosphere, whereas the gray-shaded region
represents carbonyl sulfide (OCS) photolysis and contribution to the SSA. The blue area sandwiched between these layers represents the ozone layer. Al-
though O3 production is maximum in the tropics, it is transported to the poles, as shown by the dynamics of the stratosphere with magenta lines. SSA,
stratospheric sulfate aerosols; UT-LS exchange, air mass exchange at midlatitude between the upper troposphere and lower stratosphere.
Shaheen et al. PNAS | October 29, 2013 | vol. 110 | no. 44 | 17665
EARTH,ATMOSPHERIC,
ANDPLANETARYSCIENCES
SPECIALFEATURE
ozone variation in the tropical upper troposphere and lower
stratosphere via the OEI and sea surface temperature anomalies
in the El-Niño 3.4 region. The combined enhanced ozone levels
observed by satellites and elevated upward flow of air masses from
the Intertropical Convergence Zone may provide a source of
anomalous sulfate not present in non-ENSO years. The presence
of stratospheric volcanic emissions in two ENSO time slots also
shows that they exert a significant effect on the upper atmospheric
odd oxygen cycle, which our study suggests is captured in the sul-
fate O-isotopic record. Future measurements of different ENSO
periods in volcanically quiescent periods will help to quantify
clearly the unique fingerprints of ENSOs on the sulfate O-isotopic
anomaly. These results provide preliminary insight into ENSO-
driven climate fluctuations on the concentration of ozone and
S(IV) oxidation, and how this transported aerosol captures new
facets of chemical and transport history during moderate and se-
vere ENSO events. Understanding of the ENSO signal and its
frequency is important to understand the perturbation in the
tropical climate and its relevance to the global climate system.
The ENSO-driven climate patterns had a significant impact on the
proliferation and collapse of the Mayan civilization, as inferred
from the rainfall patterns preserved in the δ18
O of stalagmites (62).
A higher resolution, multidecadal record of oxygen-triple isotopic
composition of sulfate aerosol is needed to investigate ocean-
atmosphere-biosphere interaction using a global comprehensive
Earth system model. The information can be used to assess so-
cioeconomic costs of climate vulnerabilities and to develop sus-
tainable solutions.
Materials and Methods
The surface snow samples were acquired to analyze oxygen-triple isotope
composition of both nitrate and sulfate [National Science Foundation polar
program, project South Pole Atmospheric Nitrate Isotopic Analysis (SPANIA)]
with the highest resolution (∼1-cm depth interval) from a snow pit (1 m ×
1 m) at the South Pole (45). Organic impurities from each aliquot were re-
moved from the composite samples (∼6-cm depth) by adding 2.0 mL of
peroxide (30% by volume) and further passing through polyvinyl pyrrolidine
C18 (Alltech) resins. Purified SO4 solution was converted to silver sulfate and
pyrolysed at 1,050 °C (14, 63) using a quartz tube. These samples were
combined in a prior study to obtain sufficient sample for the oxygen-triple
isotope measurements of nitrates, and the remaining solution was used for
oxygen-triple isotope analysis of sulfates. Oxygen-triple isotopic composition
was measured using a Thermo Finnigan Mat-253 Isotope Ratio mass spec-
trometer and corrected for high-temperature oxygen-isotope exchange with
quartz (64). The reported sample dates are calculated from an average an-
nual snow accumulation rate; therefore, actual dates in composite samples
may be shifted by ±4 mo, which defines the maximal uncertainty in time.
The oxygen isotopic anomaly is based on the nss-SO4 concentration. The sea
sulfate carries no O-isotopic anomaly (Δ17
O = 0), and this component was
removed using sodium concentration as a tracer of sea salt (28). The sea salt
contribution is ∼3–7% at maximum at the South Pole, and the correction
factor is small.
The OEI is obtained by Ziemke et al. (65) from the variation in ozone
concentrations at tropical latitudes (15° S–15° N). The following ozone data
were acquired from four different satellites: TOMS, Earth probe TOMS, solar
backscatter UV, and Aura ozone monitoring instrument. The measured
zonal variability in ozone was verified with the O3 data obtained from the
Goddard Earth Observing System chemistry climate model and from the
microwave limb sounder vertical profile of O3, cloud ice, temperature, and
pressure. This comparison also confirmed that zonal variability in total column
ozone in the tropics is mostly caused by ENSO events and provides a direct
measure of the changes in tropospheric ozone levels (65–67).
ACKNOWLEDGMENTS. We thank A. Hill and S. Chakraborty for useful
discussions about the OEI. We also thank the reviewers for valuable
suggestions that helped to improve the manuscript. M.H.T. and R.S. thank
the National Science Foundation Atmospheric Chemistry Division for the
support through Award ATM0960594, which allowed the present mea-
surements to be made. The National Science Foundation Office of Polar Programs
is also acknowledged for its generous support, which permitted the snow
pit work at the South Pole Atmospheric Nitrate Isotope Analysis through
SPANIA Award OPP0125761. J.S. thanks the Agence Nationale de la
Recherche [ANR-NT09-431976-volcanic and solar radiative forcing (VOL-
SOL)] and the Centre National de la Recherche Scientifique/Projet In-
ternational de Coopération Scientifique (PICS) exchange program for their
financial support for maintaining the collaboration with the University of
California, San Diego.
1. Jones A, Haywood J, Boucher O (2011) A comparison of the climate impacts of geo-
engineering by stratospheric SO2 injection and by brightening of marine stratocu-
mulus cloud. Atmos Sci Lett 12(2):176–183.
2. Jones A, Haywood J, Boucher O, Kravitz B, Robock A (2010) Geoengineering by
stratospheric SO2 injection: Results from the Met Office HadGEM(2) climate model
and comparison with the Goddard Institute for Space Studies ModelE. Atmos Chem
Phys 10(13):5999–6006.
3. Leibensperger EM, et al. (2012) Climatic effects of 1950-2050 changes in US anthro-
pogenic aerosols—Part 1: Aerosol trends and radiative forcing. Atmos Chem Phys
12(7):3333–3348.
4. Soloman SQD, Manning M (2007) Intergovernmental Panel on Climate Change Fourth
Assessment Report, Climate Change 2007: The Physical Science Basis (Cambridge Univ
Press, New York).
5. Deshler T (2008) A review of global stratospheric aerosol: Measurements, importance,
life cycle, and local stratospheric aerosol. Atmos Res 90(2-4):223–232.
6. Halmer MM, Schmincke HU, Graf HF (2002) The annual volcanic gas input into the
atmosphere, in particular into the stratosphere: A global data set for the past 100
years. J Volcanol Geotherm Res 115(3-4):511–528.
7. Solomon S, et al. (2011) The persistently variable “background” stratospheric aerosol
layer and global climate change. Science 333(6044):866–870.
8. Bauer E, Petoukhov V, Ganopolski A, Eliseev AV (2008) Climatic response to anthro-
pogenic sulphate aerosols versus well-mixed greenhouse gases from 1850 to 2000 AD
in CLIMBER-2. Tellus B Chem Phys Meterol 60(1):82–97.
9. Miyakawa T, Takegawa N, Kondo Y (2007) Removal of sulfur dioxide and formation
of sulfate aerosol in Tokyo. Journal of Geophysical Research. Atmospheres, 10.1029/
2006JD007896.
10. Alexander B, et al. (2012) Isotopic constraints on the formation pathways of sulfate
aerosol in the marine boundary layer of the subtropical northeast Atlantic Ocean.
Journal of Geophysical Research. Atmospheres, 10.1002/rcm.6332.
11. Seinfeld HJ, Pandis NS (2006) Atmospheric Chemistry and Physics: From Air Pollution
to Climate Change (Wiley Interscience, New York).
12. Savarino J, Lee CCW, Thiemens MH (2000) Laboratory oxygen isotopic study of sulfur
(IV) oxidation: Origin of the mass-independent oxygen isotopic anomaly in atmo-
spheric sulfates and sulfate mineral deposits on Earth. Journal of Geophysical Re-
search. Atmospheres 105(D23):29079–29088.
13. Patris N, Cliff SS, Quinn PK, Kasem M, Thiemens MH (2007) Isotopic analysis of aerosol
sulfate and nitrate during ITCT-2k2: Determination of different formation pathways as
a function of particle size. Journal of Geophysical Research. Atmospheres, 10.1029/
2005jd006214.
14. McCabe JR, Savarino J, Alexander B, Gong S, Thiemens MH (2006) Isotopic con-
straints on non-photochemical sulfate production in the Arctic winter. Geophys Res
Lett, 10.1029/2005GL025164.
15. Alexander B, Savarino J, Barkov NI, Delmas RJ, Thiemens MH (2002) Climate driven
changes in the oxidation pathways of atmospheric sulfur. Geophys Res Lett 29(14):
30–34.
16. Alexander B, Savarino J, Kreutz KJ, Thiemens MH (2004) Impact of preindustrial
biomass-burning emissions on the oxidation pathways of tropospheric sulfur and
nitrogen. Journal of Geophysical Research. Atmospheres, 10.1029/2003jd004218.
17. Alexander B, et al. (2003) East Antarctic ice core sulfur isotope measurements over
a complete glacial-interglacial cycle. Journal of Geophysical Research. Atmospheres,
10.1029/2003jd003513.
18. Jenkins KA, Bao H (2006) Multiple oxygen and sulfur isotope compositions of atmo-
spheric sulfate in Baton Rouge, LA, USA. Atmos Environ 40(24):4528–4537.
19. Holt BD, Kumar R, Cunningham PT (1981) O-18 study of the aqueous-phase oxidation
of sulfur-dioxide. Atmos Environ 15(4):557–566.
20. Holt BD, Cunningham PT, Engelkemeir AG, Graczyk DG, Kumar R (1983) O-18 study of
nonaqueous-phase oxidation of sulfur-dioxide. Atmos Environ 17(3):625–632.
21. Heidenreich JE, Thiemens MH (1983) A non-mass-dependent isotope effect in the
production of ozone from molecular-oxygen. J Chem Phys 78(2):892–895.
22. Heidenreich JE, Thiemens MH (1986) A non-mass-dependent oxygen isotope effect in
the production of ozone from molecular-oxygen—The role of molecular symmetry in
isotope chemistry. J Chem Phys 84(4):2129–2136.
23. Thiemens MH (2006) History and applications of mass-independent isotope effects.
Annu Rev Earth Planet Sci 34:217–262.
24. Thiemens MH, Chakraborty S, Dominguez G (2012) The physical chemistry of mass-
independent isotope effects and their observation in nature. Ann Rev Phys Chem
63:155–177.
25. Krankowsky D, et al. (2007) Stratospheric ozone isotope fractionations derived from col-
lected samples. Journal of Geophysical Research. Atmospheres, 10.1029/2006JD007855.
26. Savarino J, Thiemens MH (1999) Mass-independent oxygen isotope (O-16, O-17, O-18)
fractionation found in H-x, O-x reactions. J Phys Chem A 103(46):9221–9229.
27. Alexander B, Park RJ, Jacob DJ, Sunling G (2009) Transition metal-catalyzed oxidation
of atmospheric sulfur: Global implications for the sulfur budget. Journal of Geo-
physical Research. Atmospheres, 10.1029/2008JD010486.
28. Savarino J, Slimane B, Cole-Dai J, Thiemens MH (2003) Evidence from sulfate mass
independent oxygen isotopic compositions of dramatic changes in atmospheric oxi-
dation following massive volcanic eruptions. J Geophys Res 108(D21):ACH7-1–6.
17666 | www.pnas.org/cgi/doi/10.1073/pnas.1213149110 Shaheen et al.
29. Lyons JR (2001) Mass-independent fractionation of oxygen isotopes in Earth’s at-
mosphere. Astrobiology 1(3):333–334.
30. Savarino J, Bekki S, Cole-Dai JH, Thiemens MH (2003) Evidence from sulfate mass
independent oxygen isotopic compositions of dramatic changes in atmospheric oxi-
dation following massive volcanic eruptions. Journal of Geophysical Research. At-
mospheres 108(D21):ACH7-1-6.
31. Baroni M, Savarino J, Cole-Dai J, Rai VK, Thiemens MH (2008) Anomalous sulfur iso-
tope compositions of volcanic sulfate over the last millennium in Antarctic ice cores.
Journal of Geophysical Research. Atmospheres, 10.1029/2008jd010185.
32. Cole-Dai J, et al. (2009) Cold decade (AD 1810-1819) caused by Tambora (1815)
and another (1809) stratospheric volcanic eruption. Geophys Res Lett, 10.1029/
2009gl040882.
33. Halmer MM (2005) Have volcanoes already passed their zenith influencing the ozone
layer? Terra Nova 17(6):500–502.
34. Robock A, Adams T, Moore M, Oman L, Stenchikov G (2007) Southern Hemisphere
atmospheric circulation effects of the 1991 Mount Pinatubo eruption. Geophys Res
Lett, 10.1029/2007gl031403.
35. Cole-Dai J (2010) Volcanoes and climate. Wiley Interdiscip Rev Clim Change 1(6):
824–839.
36. Cole-Dai J, Mosley-Thompson E, Qin DH (1999) Evidence of the 1991 Pinatubo volcanic
eruption in South Polar snow. Chin Sci Bull 44(8):756–760.
37. Sato M, Hansen JE, McCormick MP, Pollack JB (1993) Stratospheric aerosol of optical
depths, 1850-1990. Journal of Geophysical Research. Atmospheres 98(D12):22987–22994.
38. Chaochao G, Robock A, Ammann C (2008) Volcanic forcing of climate over the past
1500 years: An improved ice core-based index for climate models. Journal of Geo-
physical Research. Atmospheres 113(D23):D23111, 10.1029/2008jd010239.
39. Zahn A, Franz P, Bechtel C, Grooss JU, Roeckmann T (2006) Modelling the budget of
middle atmospheric water vapour isotopes. Atmos Chem Phys 6:2073–2090.
40. Jones RL, et al. (1986) The water-vapor budget of the stratosphere studied using LIMS
and SAMS satellite data. Q J R Meteorol Soc 112(474):1127–1143.
41. Crutzen PJ, Lawrence MG, Poschl U (1999) On the background photochemistry of
tropospheric ozone. Tellus Series A: Dynamic Meteorology and Oceanography 51(1):
123–146.
42. Lyons JR (2001) Transfer of mass-independent fractionation in ozone to other oxy-
gen-containing radicals in the atmosphere. Geophys Res Lett 28(17):3231–3234.
43. Ravishankara AR (2012) Atmospheric science. Water vapor in the lower stratosphere.
Science 337(6096):809–810.
44. Anderson JG, Wilmouth DM, Smith JB, Sayres DS (2012) UV dosage levels in summer:
Increased risk of ozone loss from convectively injected water vapor. Science
337(6096):835–839.
45. McCabe JR, Thiemens MH, Savarino J (2007) A record of ozone variability in South
Pole Antarctic snow: Role of nitrate oxygen isotopes. Journal of Geophysical Re-
search-Atmospheres, 10.1029/2006JD007822.
46. Takegawa N, et al. (2003) Photochemical production of O-3 in biomass burning
plumes in the boundary layer over northern Australia. Geophys Res Lett, 10.1029/
2003gl017017.
47. Kita K, et al. (2002) Photochemical production of ozone in the upper troposphere in
association with cumulus convection over Indonesia. Journal of Geophysical Research.
Atmospheres, 10.1029/2001jd000844.
48. Hood LL, Soukharev BE (2005) Interannual variations of total ozone at northern
midlatitudes correlated with stratospheric EP flux and potential vorticity. J Atmos Sci
62(10):3724–3740.
49. Solomon S, et al. (2010) Contributions of stratospheric water vapor to decadal
changes in the rate of global warming. Science 327(5970):1219–1223.
50. Lee CCW, Savarino J, Cachier H, Thiemens MH (2002) Sulfur (S-32, S-33, S-34, S-36) and
oxygen (O-16, O-17, O-18) isotopic ratios of primary sulfate produced from combus-
tion processes. Tellus B Chem Phys Meterol 54(3):193–200.
51. Dominguez G, et al. (2008) Discovery and measurement of an isotopically distinct
source of sulfate in Earth’s atmosphere. Proc Natl Acad Sci USA 105(35):12769–12773.
52. Holt BD, Cunningham PT, Kumar R (1981) Oxygen isotopy of atmospheric sulfates.
Environ Sci Technol 15(7):804–808.
53. Kunasek SA, et al. (2010) Sulfate sources and oxidation chemistry over the past 230
years from sulfur and oxygen isotopes of sulfate in a West Antarctic ice core. Journal
of Geophysical Research. Atmospheres, 10.1029/2010JD013846.
54. Sofen ED, Alexander B, Kunasek SA (2011) The impact of anthropogenic emissions on
atmospheric sulfate production pathways, oxidants, and ice core Delta O-17(SO4
2−
).
Atmos Chem Phys 11(7):3565–3578.
55. Tichomirowa M, Heidel C (2012) Regional and temporal variability of the isotope
composition (O, S) of atmospheric sulphate in the region of Freiberg, Germany, and
consequences for dissolved sulphate in groundwater and river water. Isotopes Envi-
ron Health Stud 48(1):118–143.
56. Bruehl C, Lelieveld J, Crutzen PJ, Tost H (2012) The role of carbonyl sulphide as
a source of stratospheric sulphate aerosol and its impact on climate. Atmos Chem Phys
12(3):1239–1253.
57. Sunilkumar SV, Parameswaran K, Thampi BV, Ramkumar G (2011) Variability in back-
ground stratospheric aerosols over the tropics and its association with atmospheric dy-
namics. Journal of Geophysical Research. Atmospheres, 10.1029/2010jd015213.
58. Legrand M, Wagenbach D (1999) Impact of the Cerro Hudson and Pinatubo volcanic
eruptions on the Antarctic air and snow chemistry. Journal of Geophysical Research.
Atmospheres 104(D1):1581–1596.
59. Legrand M, Wolff E, Wagenbach D (1999) Antarctic aerosol and snowfall chemistry:
Implications for deep Antarctic ice-core chemistry. Annals of Glaciology 29(1):66–72.
60. Jourdain B, Legrand M (2001) Seasonal variations of atmospheric dimethylsulfide,
dimethylsulfoxide, sulfur dioxide, methanesulfonate, and non-sea-salt sulfate aero-
sols at Dumont d’Urville (coastal Antarctica) (December 1998 to July 1999). Journal of
Geophysical Research. Atmospheres 106(D13):14391–14408.
61. Preunkert S, et al. (2008) Seasonality of sulfur species (dimethyl sulfide, sulfate, and
methanesulfonate) in Antarctica: Inland versus coastal regions. Journal of Geo-
physical Research. Atmospheres 113(D15).
62. Kennett DJ, et al. (2012) Development and disintegration of Maya political systems in
response to climate change. Science 338(6108):788–791.
63. Savarino J, Alexander B, Darmohusodo V, Thiemens MH (2001) Sulfur and oxygen
isotope analysis of sulfate at micromole levels using a pyrolysis technique in a con-
tinuous flow system. Anal Chem 73(18):4457–4462.
64. Schauer AJ, et al. (2012) Oxygen isotope exchange with quartz during pyrolysis of
silver sulfate and silver nitrate. Rapid Commun Mass Spectrom 26(18):2151–2157.
65. Ziemke JR, Chandra S, Oman LD, Bhartia PK (2010) A new ENSO index derived from
satellite measurements of column ozone. Atmos Chem Phys 10(8):3711–3721.
66. Chandra S, Ziemke JR, Bhartia PK, Martin RV (2002) Tropical tropospheric ozone:
implications for dynamics and biomass burning. J Geophys Res 107(D14):ACH3-1–17.
67. Oman LD, et al. (2011) The response of tropical tropospheric ozone to ENSO. Geophys
Res Lett, 10.1029/2011gl047865.
Shaheen et al. PNAS | October 29, 2013 | vol. 110 | no. 44 | 17667
EARTH,ATMOSPHERIC,
ANDPLANETARYSCIENCES
SPECIALFEATURE
Supporting Information
Shaheen et al. 10.1073/pnas.1213149110
Fig. S1. Concentration of SO4, NO3, Cl, and methane sulfonic acid (MSA) of aerosol samples extracted from snow pit samples at the South Pole station,
Antarctica, from 1982 to 2002.
Shaheen et al. www.pnas.org/cgi/content/short/1213149110 1 of 2
Table S1. Oxygen-triple isotope composition and concentration profile of sulfate aerosol retrieved from a 1-m × 1-m snow pit at the
South Pole
Sample depth (cm) Year SO4 (ppb) δ17
O (‰) δ18
O (‰) Δ17
O (‰) Δ17
O* (‰)
18–28 2002–2001.55 43.71 3.89 5.13 1.35 1.51
28–42 2001.55–2000.92 57.50 1.29 −0.08 1.43 1.61
42–70 2000.92–2000.07 50.55 5.10 7.54 1.26 1.41
70–81 2000.07–1999.6 59.50 2.42 1.73 1.63 1.85
87–94 1999.5–1999.2 52.80 2.67 2.51 1.64 1.85
99–103 1999.08–1999.0 59.03 5.80 9.67 0.82 0.90
103–113 1999–1998.33 49.33 4.69 5.05 2.52 2.89
113–117 1998.33–1998.0 80.96 6.10 8.86 1.54 1.73
117–122 1998.0–1997.7 99.72 3.49 2.10 2.52 2.89
122–133 1997.7–1997.25 99.72 5.98 7.40 2.24 2.55
140–148 1996.9–1996.6 57.48 4.81 3.85 3.00 3.45
148–154 1996.6–1996.3 57.48 4.87 2.82 3.63 4.18
154–162 1996.3–1996 65.93 3.44 0.76 3.22 3.71
168–177 1995.84–1995.52 119.09 3.92 3.60 2.14 2.44
177–188 1995.52–1995.20 56.45 4.65 4.51 2.58 2.95
194–200 1995.0–1994.71 80.02 0.55 −1.64 1.46 1.64
232–238 1993.36–1993.18 66.08 2.25 1.26 1.70 1.92
250–256 1992.70–1992.49 76.87 3.78 1.39 3.34 3.84
256–263 1992.49–1992.31 80.15 3.18 1.73 2.50 2.86
263–268 1992.31–1992.21 127.16 4.07 2.86 2.71 3.10
268–274 1992.21–1991.69 164.64 4.82 3.57 3.12 3.58
274–279 1991.69–1991.30 160.96 3.71 1.48 3.06 3.52
279–284 1991.30–1991.0 104.14 4.42 3.09 2.94 3.37
284–289 1991.0–1990.82 75.92 4.20 2.50 2.98 3.42
289–295 1990.82–1990.52 71.87 5.40 4.27 3.29 3.78
295–305 1990.52–1990.13 70.45 5.19 3.12 3.87 4.47
305–316 1990.13–1989.71 114.14 7.39 11.87 1.28 1.43
316–320 1989.71–1989.53 110.97 3.05 2.10 2.02 2.30
320–328 1989.53–1989.28 66.42 1.39 0.79 1.04 1.15
328–338 1989.21–1988.68 40.18 2.37 −0.65 2.98 3.42
338–347 1988.68–1988.45 48.61 3.13 0.49 3.14 3.61
347–356 1988.45–1988 49.08 2.32 0.74 2.04 2.32
356–365 1988–1987.71 54.67 2.24 −0.42 2.69 3.08
365–378 1987.71–1987 64.42 2.88 0.61 2.75 3.16
378–386 1987–1986.74 36.52 3.55 1.68 2.88 3.31
386–397 1986.7–1986.4 41.11 2.77 2.08 1.79 2.03
397–402 1986.4–1986.2 74.18 1.19 3.95 1.27 1.42
402–412 1986.2–1985.8 46.22 3.84 1.73 3.32 3.82
420–428 1985.3–1984.9 80.09 1.75 −1.77 2.99 3.44
428–441 1984.9–1984.2 83.01 1.46 −1.20 2.26 2.59
441–445 1984.2–1984 60.70 3.3.7 2.97 1.93 2.20
445–456 1984–1983.5 47.64 2.71 1.46 2.22 2.54
456–465 1983.5–1983 63.01 3.67 5.87 0.70 0.76
478–490 1982.5–1982.0 99.50 3.07 2.27 2.00 2.28
490–502 1982–1981.5 44.82 4.45 4.00 2.87 3.30
525–537 1980.5–1980 45.53 5.73 10.40 0.40 0.41
Depth profile indicates initial and final ice layers of the composite samples used to measure oxygen-triple isotopic composition. Δ17
O*, corrected for high-
temperature pyrolysis (1,050 °C) using a quartz cup as Δ17
O* = 1.17 × Δ17
O(quartz) − 0.06 (1).
1. Schauer AJ, et al. (2012) Oxygen isotope exchange with quartz during pyrolysis of silver sulfate and silver nitrate. Rapid Commun Mass Spectrom 26(18):2151–2157.
Shaheen et al. www.pnas.org/cgi/content/short/1213149110 2 of 2

More Related Content

What's hot

Effects of climate change on the emmission of n2o
Effects of climate change on the emmission of n2oEffects of climate change on the emmission of n2o
Effects of climate change on the emmission of n2oaparmar85
 
Global Sea Level Linked To Global Temperature
Global Sea Level Linked To Global TemperatureGlobal Sea Level Linked To Global Temperature
Global Sea Level Linked To Global Temperatureireportergr
 
List of publications Jesper Platz
List of publications Jesper PlatzList of publications Jesper Platz
List of publications Jesper PlatzJesper Platz
 
ResearchPaper_PastClimates_DrRoark_AaronMunsart_050915
ResearchPaper_PastClimates_DrRoark_AaronMunsart_050915ResearchPaper_PastClimates_DrRoark_AaronMunsart_050915
ResearchPaper_PastClimates_DrRoark_AaronMunsart_050915Aaron Munsart
 
radio isotope in ground water study
radio isotope in ground water studyradio isotope in ground water study
radio isotope in ground water studypreethi durairaj
 
Human and natural_influences_on_the_changing_thermal_structure_of_the_atmosphere
Human and natural_influences_on_the_changing_thermal_structure_of_the_atmosphereHuman and natural_influences_on_the_changing_thermal_structure_of_the_atmosphere
Human and natural_influences_on_the_changing_thermal_structure_of_the_atmosphereSérgio Sacani
 
SOME STUDIES IN REGIONAL ENVIRONMENTAL AIR QUALITY MODELING, CLIMATE CHANGE A...
SOME STUDIES IN REGIONAL ENVIRONMENTAL AIR QUALITY MODELING, CLIMATE CHANGE A...SOME STUDIES IN REGIONAL ENVIRONMENTAL AIR QUALITY MODELING, CLIMATE CHANGE A...
SOME STUDIES IN REGIONAL ENVIRONMENTAL AIR QUALITY MODELING, CLIMATE CHANGE A...grssieee
 
Oil&Gas Thought-Leader Webinar - New Plays for Old Ideas - Dr. Sander Houben
Oil&Gas Thought-Leader Webinar - New Plays for Old Ideas - Dr. Sander HoubenOil&Gas Thought-Leader Webinar - New Plays for Old Ideas - Dr. Sander Houben
Oil&Gas Thought-Leader Webinar - New Plays for Old Ideas - Dr. Sander HoubenAnn-Marie Roche
 
Distribution and habitability of (meta)stable brines on present-day Mars
Distribution and habitability of (meta)stable brines on present-day MarsDistribution and habitability of (meta)stable brines on present-day Mars
Distribution and habitability of (meta)stable brines on present-day MarsSérgio Sacani
 
Brito gabriel-rvq-no prelo
Brito gabriel-rvq-no preloBrito gabriel-rvq-no prelo
Brito gabriel-rvq-no prelowfkam
 
Geochemical_conditions_and_environmental
Geochemical_conditions_and_environmentalGeochemical_conditions_and_environmental
Geochemical_conditions_and_environmentalMiltiadis Nimfopoulos
 
Fedunik hofman 2019. kinetics of solid-gas reactions and their carbonatos
Fedunik hofman 2019. kinetics of solid-gas reactions and their carbonatosFedunik hofman 2019. kinetics of solid-gas reactions and their carbonatos
Fedunik hofman 2019. kinetics of solid-gas reactions and their carbonatosMayliSanchezAlcocer
 
Jiang 2015 GCA water isotopes and salinity
Jiang 2015 GCA water isotopes and salinityJiang 2015 GCA water isotopes and salinity
Jiang 2015 GCA water isotopes and salinityRichard Worden
 
Corbett mortality ship death / a factor 5.4 worse due to BIMCO & Co on actual...
Corbett mortality ship death / a factor 5.4 worse due to BIMCO & Co on actual...Corbett mortality ship death / a factor 5.4 worse due to BIMCO & Co on actual...
Corbett mortality ship death / a factor 5.4 worse due to BIMCO & Co on actual...www.thiiink.com
 
Assessment of impact of climatic change on groundwater quality around igbokod...
Assessment of impact of climatic change on groundwater quality around igbokod...Assessment of impact of climatic change on groundwater quality around igbokod...
Assessment of impact of climatic change on groundwater quality around igbokod...Alexander Decker
 

What's hot (20)

Effects of climate change on the emmission of n2o
Effects of climate change on the emmission of n2oEffects of climate change on the emmission of n2o
Effects of climate change on the emmission of n2o
 
Global Sea Level Linked To Global Temperature
Global Sea Level Linked To Global TemperatureGlobal Sea Level Linked To Global Temperature
Global Sea Level Linked To Global Temperature
 
List of publications Jesper Platz
List of publications Jesper PlatzList of publications Jesper Platz
List of publications Jesper Platz
 
Isotope hydrology
Isotope hydrologyIsotope hydrology
Isotope hydrology
 
Isotope Hydrology
Isotope HydrologyIsotope Hydrology
Isotope Hydrology
 
ResearchPaper_PastClimates_DrRoark_AaronMunsart_050915
ResearchPaper_PastClimates_DrRoark_AaronMunsart_050915ResearchPaper_PastClimates_DrRoark_AaronMunsart_050915
ResearchPaper_PastClimates_DrRoark_AaronMunsart_050915
 
radio isotope in ground water study
radio isotope in ground water studyradio isotope in ground water study
radio isotope in ground water study
 
1268 LPSC XXX abstract
1268 LPSC XXX abstract1268 LPSC XXX abstract
1268 LPSC XXX abstract
 
Human and natural_influences_on_the_changing_thermal_structure_of_the_atmosphere
Human and natural_influences_on_the_changing_thermal_structure_of_the_atmosphereHuman and natural_influences_on_the_changing_thermal_structure_of_the_atmosphere
Human and natural_influences_on_the_changing_thermal_structure_of_the_atmosphere
 
SOME STUDIES IN REGIONAL ENVIRONMENTAL AIR QUALITY MODELING, CLIMATE CHANGE A...
SOME STUDIES IN REGIONAL ENVIRONMENTAL AIR QUALITY MODELING, CLIMATE CHANGE A...SOME STUDIES IN REGIONAL ENVIRONMENTAL AIR QUALITY MODELING, CLIMATE CHANGE A...
SOME STUDIES IN REGIONAL ENVIRONMENTAL AIR QUALITY MODELING, CLIMATE CHANGE A...
 
Oil&Gas Thought-Leader Webinar - New Plays for Old Ideas - Dr. Sander Houben
Oil&Gas Thought-Leader Webinar - New Plays for Old Ideas - Dr. Sander HoubenOil&Gas Thought-Leader Webinar - New Plays for Old Ideas - Dr. Sander Houben
Oil&Gas Thought-Leader Webinar - New Plays for Old Ideas - Dr. Sander Houben
 
Distribution and habitability of (meta)stable brines on present-day Mars
Distribution and habitability of (meta)stable brines on present-day MarsDistribution and habitability of (meta)stable brines on present-day Mars
Distribution and habitability of (meta)stable brines on present-day Mars
 
Brito gabriel-rvq-no prelo
Brito gabriel-rvq-no preloBrito gabriel-rvq-no prelo
Brito gabriel-rvq-no prelo
 
Geochemical_conditions_and_environmental
Geochemical_conditions_and_environmentalGeochemical_conditions_and_environmental
Geochemical_conditions_and_environmental
 
Studies of dissolved carbohydrates (or carbohydrate - like substances) in an ...
Studies of dissolved carbohydrates (or carbohydrate - like substances) in an ...Studies of dissolved carbohydrates (or carbohydrate - like substances) in an ...
Studies of dissolved carbohydrates (or carbohydrate - like substances) in an ...
 
Fedunik hofman 2019. kinetics of solid-gas reactions and their carbonatos
Fedunik hofman 2019. kinetics of solid-gas reactions and their carbonatosFedunik hofman 2019. kinetics of solid-gas reactions and their carbonatos
Fedunik hofman 2019. kinetics of solid-gas reactions and their carbonatos
 
Jiang 2015 GCA water isotopes and salinity
Jiang 2015 GCA water isotopes and salinityJiang 2015 GCA water isotopes and salinity
Jiang 2015 GCA water isotopes and salinity
 
Corbett mortality ship death / a factor 5.4 worse due to BIMCO & Co on actual...
Corbett mortality ship death / a factor 5.4 worse due to BIMCO & Co on actual...Corbett mortality ship death / a factor 5.4 worse due to BIMCO & Co on actual...
Corbett mortality ship death / a factor 5.4 worse due to BIMCO & Co on actual...
 
Poster_2
Poster_2Poster_2
Poster_2
 
Assessment of impact of climatic change on groundwater quality around igbokod...
Assessment of impact of climatic change on groundwater quality around igbokod...Assessment of impact of climatic change on groundwater quality around igbokod...
Assessment of impact of climatic change on groundwater quality around igbokod...
 

Similar to Tales of volcanoes-ENSO-PNAS-2013-Shaheen-1213149110-with cover

S-isotope-PNAS-2014-Shaheen-1406315111 with cover
S-isotope-PNAS-2014-Shaheen-1406315111 with coverS-isotope-PNAS-2014-Shaheen-1406315111 with cover
S-isotope-PNAS-2014-Shaheen-1406315111 with coverRobina Shaheen
 
Climate Sensitivity, Forcings, And Feedbacks_2.pptx
Climate Sensitivity, Forcings, And Feedbacks_2.pptxClimate Sensitivity, Forcings, And Feedbacks_2.pptx
Climate Sensitivity, Forcings, And Feedbacks_2.pptxObulReddy61
 
Shaheen_PNAS2014_ALH84001-oldest rock from Mars
Shaheen_PNAS2014_ALH84001-oldest rock from MarsShaheen_PNAS2014_ALH84001-oldest rock from Mars
Shaheen_PNAS2014_ALH84001-oldest rock from MarsRobina Shaheen
 
Cassini finds molecular hydrogen in the Enceladus plume: Evidence for hydroth...
Cassini finds molecular hydrogen in the Enceladus plume: Evidence for hydroth...Cassini finds molecular hydrogen in the Enceladus plume: Evidence for hydroth...
Cassini finds molecular hydrogen in the Enceladus plume: Evidence for hydroth...Sérgio Sacani
 
The future life span of Earth’s oxygenated atmosphere
The future life span of Earth’s oxygenated atmosphereThe future life span of Earth’s oxygenated atmosphere
The future life span of Earth’s oxygenated atmosphereSérgio Sacani
 
A measurement of water vapour amid a largely
A measurement of water vapour amid a largelyA measurement of water vapour amid a largely
A measurement of water vapour amid a largelyFelipe Hime
 
Airflow_Wels_Lefebvre_Robertson_Sponcomb.pdf
Airflow_Wels_Lefebvre_Robertson_Sponcomb.pdfAirflow_Wels_Lefebvre_Robertson_Sponcomb.pdf
Airflow_Wels_Lefebvre_Robertson_Sponcomb.pdfYADIIRWANTO
 
Abundance and isotopic_composition_of_gases_in_the_martian_atmosphere_from_th...
Abundance and isotopic_composition_of_gases_in_the_martian_atmosphere_from_th...Abundance and isotopic_composition_of_gases_in_the_martian_atmosphere_from_th...
Abundance and isotopic_composition_of_gases_in_the_martian_atmosphere_from_th...Sérgio Sacani
 
acp-8-4117-2008
acp-8-4117-2008acp-8-4117-2008
acp-8-4117-2008Alan Kwan
 
Trapping mechanism of O2 in water ice as first measured by Rosetta spacecraft
Trapping mechanism of O2 in water ice as first measured by Rosetta spacecraftTrapping mechanism of O2 in water ice as first measured by Rosetta spacecraft
Trapping mechanism of O2 in water ice as first measured by Rosetta spacecraftssuserf71e46
 
Lab 3 Sources of CO2 Emissions Part 1IntroductionThe natural.docx
Lab 3 Sources of CO2 Emissions Part 1IntroductionThe natural.docxLab 3 Sources of CO2 Emissions Part 1IntroductionThe natural.docx
Lab 3 Sources of CO2 Emissions Part 1IntroductionThe natural.docxsmile790243
 
Introduction to the principles of 18O.pdf
Introduction to the principles of 18O.pdfIntroduction to the principles of 18O.pdf
Introduction to the principles of 18O.pdfFeyzaAkgnlDiner
 
Poster Final Draft
Poster Final DraftPoster Final Draft
Poster Final DraftKyle Gosnell
 
Poster Final Draft
Poster Final DraftPoster Final Draft
Poster Final DraftKyle Gosnell
 
1 Unit Atmosphere .pptx
1 Unit Atmosphere .pptx1 Unit Atmosphere .pptx
1 Unit Atmosphere .pptxADARSHDINKAR1
 
Lecture4 sep16-bb
Lecture4 sep16-bbLecture4 sep16-bb
Lecture4 sep16-bbPeter Shiv
 
Geo chemistry ; Geo chemistry of Atomosphere
Geo chemistry ; Geo chemistry of AtomosphereGeo chemistry ; Geo chemistry of Atomosphere
Geo chemistry ; Geo chemistry of AtomosphereRuhr Universität Bochum
 
Chemical Fractionation Modeling of Plumes Indicates a Gas-rich, Moderately Al...
Chemical Fractionation Modeling of Plumes Indicates a Gas-rich, Moderately Al...Chemical Fractionation Modeling of Plumes Indicates a Gas-rich, Moderately Al...
Chemical Fractionation Modeling of Plumes Indicates a Gas-rich, Moderately Al...Sérgio Sacani
 

Similar to Tales of volcanoes-ENSO-PNAS-2013-Shaheen-1213149110-with cover (20)

S-isotope-PNAS-2014-Shaheen-1406315111 with cover
S-isotope-PNAS-2014-Shaheen-1406315111 with coverS-isotope-PNAS-2014-Shaheen-1406315111 with cover
S-isotope-PNAS-2014-Shaheen-1406315111 with cover
 
Climate Sensitivity, Forcings, And Feedbacks_2.pptx
Climate Sensitivity, Forcings, And Feedbacks_2.pptxClimate Sensitivity, Forcings, And Feedbacks_2.pptx
Climate Sensitivity, Forcings, And Feedbacks_2.pptx
 
Shaheen_PNAS2014_ALH84001-oldest rock from Mars
Shaheen_PNAS2014_ALH84001-oldest rock from MarsShaheen_PNAS2014_ALH84001-oldest rock from Mars
Shaheen_PNAS2014_ALH84001-oldest rock from Mars
 
Cassini finds molecular hydrogen in the Enceladus plume: Evidence for hydroth...
Cassini finds molecular hydrogen in the Enceladus plume: Evidence for hydroth...Cassini finds molecular hydrogen in the Enceladus plume: Evidence for hydroth...
Cassini finds molecular hydrogen in the Enceladus plume: Evidence for hydroth...
 
The future life span of Earth’s oxygenated atmosphere
The future life span of Earth’s oxygenated atmosphereThe future life span of Earth’s oxygenated atmosphere
The future life span of Earth’s oxygenated atmosphere
 
A measurement of water vapour amid a largely
A measurement of water vapour amid a largelyA measurement of water vapour amid a largely
A measurement of water vapour amid a largely
 
Airflow_Wels_Lefebvre_Robertson_Sponcomb.pdf
Airflow_Wels_Lefebvre_Robertson_Sponcomb.pdfAirflow_Wels_Lefebvre_Robertson_Sponcomb.pdf
Airflow_Wels_Lefebvre_Robertson_Sponcomb.pdf
 
Abundance and isotopic_composition_of_gases_in_the_martian_atmosphere_from_th...
Abundance and isotopic_composition_of_gases_in_the_martian_atmosphere_from_th...Abundance and isotopic_composition_of_gases_in_the_martian_atmosphere_from_th...
Abundance and isotopic_composition_of_gases_in_the_martian_atmosphere_from_th...
 
acp-8-4117-2008
acp-8-4117-2008acp-8-4117-2008
acp-8-4117-2008
 
Trapping mechanism of O2 in water ice as first measured by Rosetta spacecraft
Trapping mechanism of O2 in water ice as first measured by Rosetta spacecraftTrapping mechanism of O2 in water ice as first measured by Rosetta spacecraft
Trapping mechanism of O2 in water ice as first measured by Rosetta spacecraft
 
Lab 3 Sources of CO2 Emissions Part 1IntroductionThe natural.docx
Lab 3 Sources of CO2 Emissions Part 1IntroductionThe natural.docxLab 3 Sources of CO2 Emissions Part 1IntroductionThe natural.docx
Lab 3 Sources of CO2 Emissions Part 1IntroductionThe natural.docx
 
Introduction to the principles of 18O.pdf
Introduction to the principles of 18O.pdfIntroduction to the principles of 18O.pdf
Introduction to the principles of 18O.pdf
 
Poster Final Draft
Poster Final DraftPoster Final Draft
Poster Final Draft
 
Poster Final Draft
Poster Final DraftPoster Final Draft
Poster Final Draft
 
1 Unit Atmosphere .pptx
1 Unit Atmosphere .pptx1 Unit Atmosphere .pptx
1 Unit Atmosphere .pptx
 
Why does climate change?
Why does climate change?Why does climate change?
Why does climate change?
 
Lecture4 sep16-bb
Lecture4 sep16-bbLecture4 sep16-bb
Lecture4 sep16-bb
 
global warming 2023.pdf
global warming 2023.pdfglobal warming 2023.pdf
global warming 2023.pdf
 
Geo chemistry ; Geo chemistry of Atomosphere
Geo chemistry ; Geo chemistry of AtomosphereGeo chemistry ; Geo chemistry of Atomosphere
Geo chemistry ; Geo chemistry of Atomosphere
 
Chemical Fractionation Modeling of Plumes Indicates a Gas-rich, Moderately Al...
Chemical Fractionation Modeling of Plumes Indicates a Gas-rich, Moderately Al...Chemical Fractionation Modeling of Plumes Indicates a Gas-rich, Moderately Al...
Chemical Fractionation Modeling of Plumes Indicates a Gas-rich, Moderately Al...
 

Tales of volcanoes-ENSO-PNAS-2013-Shaheen-1213149110-with cover

  • 1.
  • 2. Tales of volcanoes and El-Niño southern oscillations with the oxygen isotope anomaly of sulfate aerosol Robina Shaheena , Mariana Abauanzaa , Teresa L. Jacksona , Justin McCabea,b , Joel Savarinoc , and Mark H. Thiemensa,1 a Department of Chemistry and Biochemistry, University of California San Diego, La Jolla, CA 92093; b Pacific Ridge School, Carlsbad, CA 92009; and c Laboratoire de Glaciologie et Géophysique de l’Environnement, 38402 St. Martin d’Heres, Grenoble, France Edited by Karl K. Turekian, Yale University, North Haven, CT, and approved February 1, 2013 (received for review September 19, 2012) The ability of sulfate aerosols to reflect solar radiation and simul- taneously act as cloud condensation nuclei renders them central players in the global climate system. The oxidation of S(IV) com- pounds and their transport as stable S(VI) in the Earth’s system are intricately linked to planetary scale processes, and precise charac- terization of the overall process requires a detailed understanding of the linkage between climate dynamics and the chemistry lead- ing to the product sulfate. This paper reports a high-resolution, 22-y (1980–2002) record of the oxygen-triple isotopic composition of sulfate (SO4) aerosols retrieved from a snow pit at the South Pole. Observed variation in the O-isotopic anomaly of SO4 aerosol is linked to the ozone variation in the tropical upper troposphere/ lower stratosphere via the Ozone El-Niño Southern Oscillations (ENSO) Index (OEI). Higher Δ17 O values (3.3‰, 4.5‰, and 4.2‰) were observed during the three largest ENSO events of the past 2 decades. Volcanic events inject significant quantities of SO4 aero- sol into the stratosphere, which are known to affect ENSO strength by modulating stratospheric ozone levels (OEI = 6 and Δ17 O = 3.3‰, OEI = 11 and Δ17 O = 4.5‰) and normal oxidative pathways. Our high-resolution data indicated that Δ17 O of sulfate aerosols can record extreme phases of naturally occurring climate cycles, such as ENSOs, which couple variations in the ozone levels in the atmosphere and the hydrosphere via temperature driven changes in relative humidity levels. A longer term, higher resolu- tion oxygen-triple isotope analysis of sulfate aerosols from ice cores, encompassing more ENSO periods, is required to reconstruct paleo-ENSO events and paleotropical ozone variations. Pinatubo | El-Chichón | Cerro Hudson | Intertropical Convergence Zone Sulfate aerosols affect climate systems by altering radiation balance, temperature, precipitation, and atmospheric dy- namics (1, 2). The overall effect of sulfate aerosols on the climate is estimated to be net cooling (−2.0 ± 0.2 Wm−2 ) (3). According to the Intergovernmental Panel on Climate Change fourth assess- ment report, aerosols are one of the largest sources of uncer- tainties in climate prediction models due to their temporal and spatial variability (4). Globally, anthropogenic sulfur exceeds natural sources by a factor of 3 to 4 (35 Tg·y−1 vs. 10 Tg·y−1 from 1990 to 2000) (5–8). Occasionally, volcanoes emit large quantities of sulfur dioxide (SO2) directly into the atmosphere. Pinatubo, for example, released ∼30 Tg of SO2 (5). In the troposphere, SO2 has an atmospheric residence time of ∼2–3 d, and it is oxidized to SO4 2− by homogeneous and heterogeneous pathways and re- moved via wet and dry deposition (9). Gas phase oxidation of SO2 by OH and subsequent reactions with water vapor yield sulfuric acid vapor [H2SO4(g)] (10). The oxidation of aqueous SO2 by O3 and H2O2 far exceeds gas phase rates and is pH-dependent (11, 12). The oxygen-triple isotopic composition of sulfate aerosols has been demonstrated to be a useful diagnostic tool to distinguish and quantify reaction pathways and to determine the paleoox- idant levels on centennial to millennium (glacial period) time scales to present-day environments (13–18). Tropospheric SO2 has a mass-dependent oxygen isotopic composition (δ17 O ≈ 0.52 δ18 O) due to rapid isotopic equilibration with water vapor (19), which erases the source-derived oxygen isotopic signature. [The delta (δ) values denote the relative deviation of the isotope ra- tios 17 R= (17 O/16 O) and 18 R = (18 O/16 O) in a sample (Rs) with respect to standard material (Rst) in permill (‰) (e.g., δ17 O [‰] = [17 Rs/17 Rst −1] * 1,000). Isotope abundance or depletion is measured with reference to a standard material (e.g., for oxygen isotopes, reference material is Vienna Standard Mean Ocean Water [VSMOW]).] S(IV) species (SO2aq, HSO3 − , and SO3 −2 ) are oxi- dized to stable sulfate [S(VI)] via OH radicals, H2O2 and O3 (19, 20). The ozone molecule is a unique quantitative tracer of oxi- dation reactions because it possesses the highest enrichment in the heavier isotopes of oxygen (70–120‰) and oxygen isotope anomaly (Δ17 O = 25–30‰). The anomalous oxygen isotopic dis- tribution of ozone has been shown to be transferred to oxygen- carrying molecules, such as SOx-NOx-ClOx-HOx (21–25) (Δ17 O ≈ δ17 O − 0.52 δ18 O, a mass-dependent process, has Δ17 O = 0; mass- independent processes have Δ17 O ≠ 0). The positive Δ17 O of sulfate derives from aqueous phase oxidation of SO2 by H2O2 and O3 via Reaction 2–Reaction 3 and involves transfer of the isotopic anomaly from the oxidant to the product sulfate (12, 26). All other sulfate sources, including gas-phase oxidation by OH in the troposphere via Reaction 1a and metal-catalyzed oxidation by atmospheric O2, possess mass-dependent signatures, as verified by laboratory and field measurements (12, 14, 27, 28). However, OH in the strato- sphere has been suggested to possess an O-isotopic anomaly (28, 29), which can be transferred to sulfate produced in the stratosphere via Reaction 1b. The magnitude of the transfer of the Δ17 O depends on the relative contribution (Reaction 1–Reaction 3): ðSO2 + OHÞtroposphere → HSO- 3 À Δ17 O = 0  ‰ Á [Reaction 1a] ðSO2 + OHÞstratosphere + M → HSO- 3 + M À Δ17 O ≠ 0  ‰ Á [Reaction 1b] HSO− 3 + H2O2 → H2SO4 À Δ17 O = 0:5 − 1  ‰ Á + H2O [Reaction 2] SO2− 3 + O3 → SO2− 4 À Δ17 O = 8 − 9  ‰ Á + O2 [Reaction 3] Tropospheric S(IV) oxidation by O3 (Reaction 3) is the only significant mechanism producing sulfate Δ17 O values >1‰, therefore, Δ17 O values greater than 1‰ quantitatively reflect the relative contribution of O3 during sulfate formation (12). S(IV) oxidized in the stratosphere acquires an anomalous sig- nature via Reaction 1 to Reaction 3 (30). Once sulfate is formed, the isotopic signature is stable and permanently preserved in the aerosol. In addition to defining reaction pathways, both S and O isotopic anomalies of sulfate aerosols may determine paleo- volcanic activities and reflect their upper atmospheric chemistry Author contributions: R.S., J.M., J.S., and M.H.T. designed research; R.S., M.A., and T.L.J. performed research; R.S., M.A., T.L.J., J.M., J.S., and M.H.T. contributed new reagents/ analytic tools; R.S. analyzed data; and R.S. wrote the paper. The authors declare no conflict of interest. This article is a PNAS Direct Submission. 1 To whom correspondence should be addressed. E-mail: mthiemens@ucsd.edu. This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10. 1073/pnas.1213149110/-/DCSupplemental. 17662–17667 | PNAS | October 29, 2013 | vol. 110 | no. 44 www.pnas.org/cgi/doi/10.1073/pnas.1213149110
  • 3. (28, 31, 32). It is argued that large volcanic eruptions in the Anthropocene era are depleting stratospheric ozone by providing surfaces for heterogeneous chemical reactions (33). These varia- tions in O3 can modulate the dynamic of the tropical stratosphere (quasibiennial oscillations); thus, volcanoes are considered re- sponsible for the strongest El-Niño Southern Oscillations (ENSOs) (34). In this work, we investigate whether oxygen-triple isotopic composition of sulfate aerosols may serve as a unique fingerprint of ozone chemistry to parse out the effect of volcanoes and ENSO events based on their differing oxidative pathways. The polar ice caps are nature’s best archive of Earth’s atmospheric history and preserve a record of paleovolcanic activities (30, 35). A high- resolution seasonal record of sulfate aerosols is used to reconstruct atmospheric history on annual and decadal time scales. The oxygen-triple isotope data of sulfate aerosols from Greenland (monthly samples from July 1999–June 2000) have demonstrated its potential to constrain the transport and oxidation history of sulfate aerosols to polar regions (14). Here, we present a 22-y seasonally resolved profile of major ions and oxygen-triple isotope data of sulfate aerosols obtained from surface snow sampled in a snow pit (1 m × 1 m) at the South Pole to help elucidate the processes determining the observed global variation in sulfate aerosol from 1980 to 2002, as well as the oxidation history of S(IV), its transport to the South Pole, and its linkage with the dynamics of the upper atmosphere. Our data on the oxygen-triple isotope measurements of sulfate aerosol encompass three major volcanic events of the century [El-Chichόn (17.3° N, 93.2° W, 1,205 m), Pinatubo (15.13° N, 120.35° E, 1,745 m), and Cerro Hudson (45° S, 72° W, 1,905 m)] and three major ENSO events (1982–1983, 1991–1992, and 1997–1998). Results A high-resolution temporal record (1980–2002) of sulfate aerosols extracted from the snow pit at the South Pole and the associated oxygen isotopic composition (δ17 O, δ18 O, and Δ17 O) are given in Fig. 1 and Table S1. The sulfate concentration in composite sam- ples (details provided in Materials and Methods) ranged from 36 to 165 parts per billion (ppb). The highest sulfate concentration was observed in ice layers deposited in 1991 and 1992 (Fig. 1A). The single oxygen isotope ratio of sulfate aerosol showed significant variation, ranging from 1980 to 2002 (δ18 O = −2 to +12‰). Increases in sulfate concentrations due to volcanic activities show a corresponding decrease in δ18 O (average δ18 O = 3.1‰ and 2.3‰ in 1983 and 1992, respectively). The oxygen isotopic anomaly (Δ17 O) varied from 0.4 to 4.5‰ (Fig. 1B). A high-resolution con- centration profile of sulfate aerosols indicated that volcanic sulfate layers from Pinatubo and Cerro Hudson eruptions were deposited from 1991–1992 (36) and El-Chichón from 1982–1983 with corre- sponding oxygen isotope anomaly of 4.5‰ and 3.3‰. The oxygen isotopic anomaly in sulfate aerosols observed during three major El-Nino events (ENSO-I = 1982–1983, ENSO-II = 1991–1992, ENSO-III 1997–1998) and a moderate event (ENSO-IV = 1986– 1987) track the Ozone ENSO Index (OEI; Fig. 1B). An unusually high enrichment in oxygen-triple isotopic composition (δ18 O =12‰, Δ17 O = 4.1‰), along with a higher OEI of 6 is observed in 1990 and labeled as an unknown event in Fig. 1A. A plot of Δ17 O vs. δ18 O indicated a very weak (r2 = 0.2) inverse correlation (Fig. 2). In Fig. 2, the maximum Δ17 O achievable via tropospheric ozone and peroxide aqueous phase oxidation is represented by red and green rectangles and gas phase oxidation via stratospheric OH/HO2 is shown in blue. A high-resolution (∼1-cm sampling interval) concentration measurement of major ions [sulfate, methane sulfonic acid (MSA), nitrate, and chloride] obtained from the snow pit indicated no significant relation to the sulfate concentration (Fig. S1). A higher resolution sulfate con- centration profile revealed two distinct peaks 1 and 2 (Fig. 3) with a fivefold and 3.6-fold increase, respectively, in sulfate (SO4) con- centration in ice layers deposited during 1992. Peak 3 appeared during a volcanically quiescent period (1990) and showed an ap- proximately fourfold increase in sulfate concentration (190 ppb) compared with an average background value of 50 ppb observed in this study. The persistent background during the study period is defined as the time period when natural and anthropogenic emis- sions of S compounds are maintained at a relatively quasi-steady state (1999–2002). Sulfate from the El-Chichόn volcanic activity indicated as peak 4 (1982–1983) did not appear as a sharp peak; rather, it is more spread out in time. In the high-resolution profile, an increase in non-sea salt (nss) SO4 (peak 3 in 1990) is potentially associated with a significant increase in MSA, although they are not temporally identical. MSA concentrations varied from 2 to 50 ppb, with a maximum increase in1985–1986 and 1990 (Fig. 3). Discussion The dataset presented here is the longest (1980–2002) and maxi- mum time-resolved record of chemical and isotopic composition of sulfate aerosols retrieved at the South Pole, and it encom- passes three major volcanic events (El-Chichόn, Pinatubo, and Cerro Hudson eruptions in April 1982, June 1991, and October 1991, respectively). These volcanic activities introduced significant quantities of SO2 (Pinatubo = 30 Tg SO2, Cerro Hudson = 10 Tg SO2, and El-Chichόn = 7 Tg SO2) (5, 37, 38) into the stratosphere, which can be observed as distinct sulfate peaks as labeled in Figs. 1A and 3. The volcanic sulfate aerosol showed less enrichment in Fig. 1. (A) Oxygen isotopic composition (brown squares) and sulfate con- centration profile (green diamonds) of composite aerosol sulfate samples extracted from the snow pit (6 m high) at the South Pole, Antarctica. The increase in sulfate concentration due to El-Chichón and Pinatubo + Cerro Hudson is also shown. UE, unknown event. (B) Comparison of oxygen iso- tope anomaly (red lines) and OEI (blue lines) obtained by Ziemke et al. (65) from the deseasonalized trend in total O3 column measured at the equa- torial Eastern and Western Pacific, an El-Niño region. Violet bars indicate three major ENSO events: ENSO-I (1982–1983), ENSO-II (1991–1992), and ENSO-III (1997–1998). A moderate event, ENSO-IV (1986–1987), is also shown. (The scale of ENSO events is defined by the National Oceanic and Atmo- spheric Administration and is available at www.cpc.noaa.gov/products/analysis_ monitoring/ensostuff/ensoyears.shtml). Shaheen et al. PNAS | October 29, 2013 | vol. 110 | no. 44 | 17663 EARTH,ATMOSPHERIC, ANDPLANETARYSCIENCES SPECIALFEATURE
  • 4. the heavy isotopes of oxygen (δ18 Oaverage = 2.6 ± 1‰) but pos- sesses a mass-independent anomaly (El-Chichόn: Δ17 O =3.3 ‰ and Pinatubo + Cerro Hudson: Δ17 O = 4.5‰). The oxygen isotope anomaly reported here for the composite Pinatubo and Cerro Hudson sulfate sample in 1992 is similar to the previously reported higher resolution signal for Pinatubo (δ18 O = 5.1–9.5‰, Δ17 O = 3.8–4.7‰) (28). Higher Δ17 O values (3.3–4.5‰) of volcanic sul- fate and nonvolcanic ENSO events (3.8–4.2 ‰) compared with the lower tropospheric Δ17 O values (0.4–1.6‰) (13, 18) indicate the predominant role of stratospheric OH and HO2 radicals (30, 39). The concentration of these radicals in the stratosphere depends on ozone, water vapor, CH4, and NOx concentrations (40), and it is suggested to be 1.5 ± 0.3 × 106 molecules per cubic centimeter in the tropics using the global chemistry transport model (41). Numeric simulations have also indicated that the OH in the stratosphere acquires an oxygen isotope anomaly (Δ17 O = 2– 40‰) by means of exchange with NOx (39, 42). The anomalous signal of OH and HO2 in the stratosphere is preserved (39, 42) due to the extremely low water content in the stratosphere (∼5–10 ppm by volume) (43, 44), which is normally erased in the troposphere due to rapid isotope exchange with water vapor (∼3% in the tropics to 0.1% in the cold polar regions) (20). The most striking feature of the present data (Fig. 1B) is that variables OEI and Δ17 O of sulfate aerosol track each other during the strong El- Niño events of 1982–1983 (ENSO-I), 1991–1992 (ENSO-II), and 1997–1998 (ENSO-III) as well as during the moderate event of 1986–1987 (ENSO-IV). These ENSO time periods are defined according to the National Oceanic and Atmo- spheric Administration as sea surface temperature anomalies (El- Niño = warm and La-Niña = cool) in the tropical Pacific (5° N–5° S, 120°–170° W) (www.cpc.noaa.gov/products/analysis_monitoring/ ensostuff/ensoyears.shtml). The OEI plotted in Fig. 1B, as dis- cussed, is obtained from variation in ozone concentrations at tropical latitudes. The OEI and Δ17 O data in Fig. 1B track each other but are slightly shifted, which may derive from two factors. First, the necessity to combine samples for the nitrate (45) and sulfate measurements introduces a modest time uncertainty. The average of the combined depth is used to specify sample time and assumes a uniform sulfate distribution throughout that time period. Consequently, there is an uncertainty in time of the sulfate peak, which, at maximum, is a few months. Second, there are different indices used to capture El-Niño events [e.g., OEI, Oceanic Niño Index (ONI)], which use a variety of differing geophysical obser- vations to document El-Niño events, and they are not necessarily exactly temporally equivalent to one another. The comparison of the oxygen isotopic anomaly of sulfate aerosols with the OEI reveals that higher Δ17 O values are associated with elevated ozone column densities measured by different satellites. The ENSO signal during two earlier events, ENSO-I [1982–1983 (El Chichόn: OEI = 6.5, Δ17 O = 3.3‰)] and ENSO-II [1991–1992 (Pinatubo and Cerro Hudson: OEI = 6, Δ17 O = 4.5‰)], may have been confounded due to the intense volcanic activities, which introduced, in addition to the SO2, significant amounts of sulfate oxidized in the troposphere with less Δ17 O, thus diluting the higher Δ17 O signal of S(IV) oxidation in the stratosphere via OH radicals. The Δ17 O of sulfate and OEI track each other fairly well despite higher concentrations of volcanic sulfate. The ENSO events in volcanically quiescent periods manifested a higher O-isotopic anomaly and OEI [stron- gest ENSO-III (1997–1998): OEI = 10.8, Δ17 O= 4.2‰; moderate ENSO-IV (1986–1987): OEI = 5, Δ17 O= 3.8‰]. A significant in- crease of total ozone during ENSO-III (1997–1998) was accom- panied by decreased precipitation, producing extensive forest fires in Indonesia, Australia, and South America (46, 47). The influence of ENSOs on total columnar ozone has been attributed to the variation in the tropopause height driven by changes of tropical deep convection and alteration of Brewer–Dobson circulation (14, 48). These observations suggest that the Δ17 O of sulfate aerosols can be used to track moderate to strong ENSO events and the variation in ozone concentration. To develop this possibility, the following points must be addressed: Where does ozone-driven ox- idation of S(IV) occur, and how is the observed variation in the tropical upper tropospheric O3 (OEI) and ENSOs chemically as- sociated with Δ17 O of sulfate aerosols? It is known that tropospheric air enters the stratosphere principally in the tropics within the In- tertropical Convergence Zone (ITCZ, a thin and dynamic region along the equator separating trade winds between the Southern and Northern Hemispheres) and transports poleward in the strato- sphere as shown in Fig. 4 (11). The ITCZ thus links the troposphere to the stratosphere and is an important corridor for the transport of aerosol and trace gases to the stratosphere (49). The maximum transport of water vapor also occurs in the ITCZ (44) in the vicinity of the ENSO region, which also corresponds to the region of the OEI measurements used in Fig. 1B. A comparison between stratospheric water vapor and tropical sea surface temperatures has demonstrated a strong correlation and an impact on the strength of El-Niño (49). The influence of ENSOs on the total column of ozone is also linked to the variation in the tropopause height. Tropical deep convection and changes in Brewer–Dobson circulation may account for the observed ozone enhancement Fig. 2. Four-isotope plot shows Δ17 O and δ18 O of sulfate aerosols extracted from the snow pit at the South Pole. A weak observed correlation indicates mixing of various sulfates from different sources. Red and green rectangles display variation in δ18 O and Δ17 O sources of sulfate. The blue rectangle is an oxidation source with stratospheric (Strat.) OH/HO2 radicals (28, 39). The green rectangle shows the range of pure hydrogen peroxide oxidation. The red square denotes the value of atmospheric (Atm.) oxygen. Max., maximum. Primary sulfate produced during fossil fuel combustion at high temperature has been shown to possess δ18 O values close to the atmospheric oxygen (51); however, sulfate produced during biomass burning showed a range of δ18 O values (50) depending on biomass type. The observed dataset reflects the range of pro- cesses contributing to the observed oxygen isotopic composition of sulfate. -600 -500 -400 -300 -200 -100 0 0 10 20 30 40 50 60 0 50 100 150 200 250 300 1977 1980 1983 1986 1989 1992 1995 1998 2001 Depth (cm) MSA(ppb) SO4(ppb) Year SO4 MSA1 23 4 Fig. 3. Concentration profile (1977–2003) of nss sulfate aerosol and MSA extracted from the snow pit samples at the South Pole. Peaks 1, 2, and 4 represent Pinatubo, Cerro Hudson, and El-Chichón volcanic sulfates, re- spectively. Peak 3 represents an unknown event. 17664 | www.pnas.org/cgi/doi/10.1073/pnas.1213149110 Shaheen et al.
  • 5. during ENSO periods (14, 48). Cumulatively, all components needed to provide ENSO-related Δ17 O enrichments are present. There is coexistence of increased ozone and water content and a dynamic mechanism of transport of aerosol into the strato- sphere with poleward migration. The ozone-induced oxidation of S(IV) via OH radicals is likely stratospheric and occurs during transport (Fig. 4). In this model, large sulfate concentrations are not required, simply a preferential ozone oxidation channel that provides the heavy isotope-enriched sulfate. A detailed model and discussion are far beyond the scope of this paper, but the isotope measurements suggest that such experiments are war- ranted. Our data indicate that the O-isotope anomaly of sulfate is sensitive to changes in atmospheric dynamics and is faithfully preserved in this record. A quasibiennial oscillation signal in the Δ17 O of nitrate extracted from the same set of aerosol samples from the South Pole has shown a similar linkage (45). In the coordinate system of Fig. 2, a source of sulfate with higher δ18 O and low Δ17 O is required. An ozone-rich source would have high δ18 O values and also high Δ17 O values. Various oxidants of different oxygen isotopic composition (OH, HO2, H2O2, and O3) can oxidize SO2 to SO4 (Fig. 2, red and green rectangles are guidelines to demonstrate variations in δ18 O with the maximal Δ17 O obtainable via aqueous phase O3 and H2O2 oxidation, and the blue rectangle indicates stratospheric OH and HO2 radical reactions, as well as the associated maximum isotopic anomaly). Fig. 2 requires the presence of a specific source, but it must be of higher δ18 O and low Δ17 O. Laboratory experiments, field observations, and numeric simulation of sulfur oxidation (14, 28, 50, 51) have shown that the δ18 O of sulfate coupled with Δ17 O can be used to distinguish between primary and secondary sulfates. Pri- mary sulfate [i.e., S(VI) produced at the emission source] exhibits a higher enrichment in δ18 O = 20–45‰ and Δ17 O = 0 (50–52). Secondary sulfate is formed when SO2 is oxidized in the atmo- sphere via homogeneous and heterogeneous pathways (51). The δ18 O of secondary sulfate thus represents a juxtaposition of the highly variable water isotopic signature derived during SO2 oxidation processes at varying latitudes and altitudes (10, 53, 54). A moderate correlation between δ18 O of rainwater and δ18 O(SO4) (55) also indicated the complexity of using δ18 O to predict sources of sulfate in rainwater. Most primary sulfate is removed via wet and dry deposition in the free troposphere; however, a fraction of S(IV), carbonyl sulfide (OCS), and traces of S(VI) are transported to the stratosphere from the ITCZ and at midlatitudes via deep convection (7, 56), thus permitting gas phase oxidation via OH/ HO2 radicals in the stratosphere (30). An unusual increase in SO4 concentration (∼100 ppb) and oxygen isotope enrichment (δ18 O ∼12‰ and Δ17 O ∼3.7‰) in 1990 are immediately followed by an increase in MSA (Fig. 3). To explain this peak, we consider two possible scenarios. The first is stratospheric volcanic emissions due to the presence of a higher oxygen isotope anomaly in the sulfate. Total ozone map- ping spectrometer (TOMS) and stratospheric aerosol and gas experiment (SAGE) satellite data indicated no significant in- crease in the global inventory of stratospheric aerosols loading during 1989–1990 (57), ruling out stratospheric volcanic emis- sions, and this leaves us with one option, a local sulfate source such as Mount Erebus (167° 25′ E, 77° 30′ S). The Smithsonian database reports a significant increase in SO2 emissions, up to 100 t/d during this period, which could be a potential source of SO4 peak (www.volcano.si.edu/reports/bulletin/contents.cfm? issue=3609). Volcanic emissions, despite their complexity, change mostly SO4 concentration (58). The second possible scenario is that biogenic sources, such as phytoplankton, emit DMS, which, on oxidation with O3, H2O2, NOx, or ClOx, can produce MSA, and ultimately an increase in nss-SO4 (59, 60). This observed spike in MSA temporally followed by an increase in nss-SO4 is consistent with a biological source. Higher resolution measure- ments of MSA on the high Antarctic Plateau (61) (both inland and coastal sites) indicated postdepositional losses of MSA and nss-SO4 production via MSA oxidation. Future sulfur isotope measurements of sulfate aerosols in this time period may help to elucidate and quantify this very specific source and oxidation process further. Conclusion The most significant observation of the sulfate multioxygen iso- topic record reported here is that observed trends in the mass- independent O-isotopic anomaly are apparently linked to the Fig. 4. Schematic depicts transport and transformation of sulfur species into the stratosphere and deposition of aged sulfate aerosol in the ice at the South Pole. The red-shaded area indicates a significant contribution of SO2 and SO4 aerosols to the SSA in the lower stratosphere, whereas the gray-shaded region represents carbonyl sulfide (OCS) photolysis and contribution to the SSA. The blue area sandwiched between these layers represents the ozone layer. Al- though O3 production is maximum in the tropics, it is transported to the poles, as shown by the dynamics of the stratosphere with magenta lines. SSA, stratospheric sulfate aerosols; UT-LS exchange, air mass exchange at midlatitude between the upper troposphere and lower stratosphere. Shaheen et al. PNAS | October 29, 2013 | vol. 110 | no. 44 | 17665 EARTH,ATMOSPHERIC, ANDPLANETARYSCIENCES SPECIALFEATURE
  • 6. ozone variation in the tropical upper troposphere and lower stratosphere via the OEI and sea surface temperature anomalies in the El-Niño 3.4 region. The combined enhanced ozone levels observed by satellites and elevated upward flow of air masses from the Intertropical Convergence Zone may provide a source of anomalous sulfate not present in non-ENSO years. The presence of stratospheric volcanic emissions in two ENSO time slots also shows that they exert a significant effect on the upper atmospheric odd oxygen cycle, which our study suggests is captured in the sul- fate O-isotopic record. Future measurements of different ENSO periods in volcanically quiescent periods will help to quantify clearly the unique fingerprints of ENSOs on the sulfate O-isotopic anomaly. These results provide preliminary insight into ENSO- driven climate fluctuations on the concentration of ozone and S(IV) oxidation, and how this transported aerosol captures new facets of chemical and transport history during moderate and se- vere ENSO events. Understanding of the ENSO signal and its frequency is important to understand the perturbation in the tropical climate and its relevance to the global climate system. The ENSO-driven climate patterns had a significant impact on the proliferation and collapse of the Mayan civilization, as inferred from the rainfall patterns preserved in the δ18 O of stalagmites (62). A higher resolution, multidecadal record of oxygen-triple isotopic composition of sulfate aerosol is needed to investigate ocean- atmosphere-biosphere interaction using a global comprehensive Earth system model. The information can be used to assess so- cioeconomic costs of climate vulnerabilities and to develop sus- tainable solutions. Materials and Methods The surface snow samples were acquired to analyze oxygen-triple isotope composition of both nitrate and sulfate [National Science Foundation polar program, project South Pole Atmospheric Nitrate Isotopic Analysis (SPANIA)] with the highest resolution (∼1-cm depth interval) from a snow pit (1 m × 1 m) at the South Pole (45). Organic impurities from each aliquot were re- moved from the composite samples (∼6-cm depth) by adding 2.0 mL of peroxide (30% by volume) and further passing through polyvinyl pyrrolidine C18 (Alltech) resins. Purified SO4 solution was converted to silver sulfate and pyrolysed at 1,050 °C (14, 63) using a quartz tube. These samples were combined in a prior study to obtain sufficient sample for the oxygen-triple isotope measurements of nitrates, and the remaining solution was used for oxygen-triple isotope analysis of sulfates. Oxygen-triple isotopic composition was measured using a Thermo Finnigan Mat-253 Isotope Ratio mass spec- trometer and corrected for high-temperature oxygen-isotope exchange with quartz (64). The reported sample dates are calculated from an average an- nual snow accumulation rate; therefore, actual dates in composite samples may be shifted by ±4 mo, which defines the maximal uncertainty in time. The oxygen isotopic anomaly is based on the nss-SO4 concentration. The sea sulfate carries no O-isotopic anomaly (Δ17 O = 0), and this component was removed using sodium concentration as a tracer of sea salt (28). The sea salt contribution is ∼3–7% at maximum at the South Pole, and the correction factor is small. The OEI is obtained by Ziemke et al. (65) from the variation in ozone concentrations at tropical latitudes (15° S–15° N). The following ozone data were acquired from four different satellites: TOMS, Earth probe TOMS, solar backscatter UV, and Aura ozone monitoring instrument. The measured zonal variability in ozone was verified with the O3 data obtained from the Goddard Earth Observing System chemistry climate model and from the microwave limb sounder vertical profile of O3, cloud ice, temperature, and pressure. This comparison also confirmed that zonal variability in total column ozone in the tropics is mostly caused by ENSO events and provides a direct measure of the changes in tropospheric ozone levels (65–67). ACKNOWLEDGMENTS. We thank A. Hill and S. Chakraborty for useful discussions about the OEI. We also thank the reviewers for valuable suggestions that helped to improve the manuscript. M.H.T. and R.S. thank the National Science Foundation Atmospheric Chemistry Division for the support through Award ATM0960594, which allowed the present mea- surements to be made. The National Science Foundation Office of Polar Programs is also acknowledged for its generous support, which permitted the snow pit work at the South Pole Atmospheric Nitrate Isotope Analysis through SPANIA Award OPP0125761. J.S. thanks the Agence Nationale de la Recherche [ANR-NT09-431976-volcanic and solar radiative forcing (VOL- SOL)] and the Centre National de la Recherche Scientifique/Projet In- ternational de Coopération Scientifique (PICS) exchange program for their financial support for maintaining the collaboration with the University of California, San Diego. 1. Jones A, Haywood J, Boucher O (2011) A comparison of the climate impacts of geo- engineering by stratospheric SO2 injection and by brightening of marine stratocu- mulus cloud. Atmos Sci Lett 12(2):176–183. 2. Jones A, Haywood J, Boucher O, Kravitz B, Robock A (2010) Geoengineering by stratospheric SO2 injection: Results from the Met Office HadGEM(2) climate model and comparison with the Goddard Institute for Space Studies ModelE. Atmos Chem Phys 10(13):5999–6006. 3. Leibensperger EM, et al. (2012) Climatic effects of 1950-2050 changes in US anthro- pogenic aerosols—Part 1: Aerosol trends and radiative forcing. Atmos Chem Phys 12(7):3333–3348. 4. Soloman SQD, Manning M (2007) Intergovernmental Panel on Climate Change Fourth Assessment Report, Climate Change 2007: The Physical Science Basis (Cambridge Univ Press, New York). 5. Deshler T (2008) A review of global stratospheric aerosol: Measurements, importance, life cycle, and local stratospheric aerosol. Atmos Res 90(2-4):223–232. 6. Halmer MM, Schmincke HU, Graf HF (2002) The annual volcanic gas input into the atmosphere, in particular into the stratosphere: A global data set for the past 100 years. J Volcanol Geotherm Res 115(3-4):511–528. 7. Solomon S, et al. (2011) The persistently variable “background” stratospheric aerosol layer and global climate change. Science 333(6044):866–870. 8. Bauer E, Petoukhov V, Ganopolski A, Eliseev AV (2008) Climatic response to anthro- pogenic sulphate aerosols versus well-mixed greenhouse gases from 1850 to 2000 AD in CLIMBER-2. Tellus B Chem Phys Meterol 60(1):82–97. 9. Miyakawa T, Takegawa N, Kondo Y (2007) Removal of sulfur dioxide and formation of sulfate aerosol in Tokyo. Journal of Geophysical Research. Atmospheres, 10.1029/ 2006JD007896. 10. Alexander B, et al. (2012) Isotopic constraints on the formation pathways of sulfate aerosol in the marine boundary layer of the subtropical northeast Atlantic Ocean. Journal of Geophysical Research. Atmospheres, 10.1002/rcm.6332. 11. Seinfeld HJ, Pandis NS (2006) Atmospheric Chemistry and Physics: From Air Pollution to Climate Change (Wiley Interscience, New York). 12. Savarino J, Lee CCW, Thiemens MH (2000) Laboratory oxygen isotopic study of sulfur (IV) oxidation: Origin of the mass-independent oxygen isotopic anomaly in atmo- spheric sulfates and sulfate mineral deposits on Earth. Journal of Geophysical Re- search. Atmospheres 105(D23):29079–29088. 13. Patris N, Cliff SS, Quinn PK, Kasem M, Thiemens MH (2007) Isotopic analysis of aerosol sulfate and nitrate during ITCT-2k2: Determination of different formation pathways as a function of particle size. Journal of Geophysical Research. Atmospheres, 10.1029/ 2005jd006214. 14. McCabe JR, Savarino J, Alexander B, Gong S, Thiemens MH (2006) Isotopic con- straints on non-photochemical sulfate production in the Arctic winter. Geophys Res Lett, 10.1029/2005GL025164. 15. Alexander B, Savarino J, Barkov NI, Delmas RJ, Thiemens MH (2002) Climate driven changes in the oxidation pathways of atmospheric sulfur. Geophys Res Lett 29(14): 30–34. 16. Alexander B, Savarino J, Kreutz KJ, Thiemens MH (2004) Impact of preindustrial biomass-burning emissions on the oxidation pathways of tropospheric sulfur and nitrogen. Journal of Geophysical Research. Atmospheres, 10.1029/2003jd004218. 17. Alexander B, et al. (2003) East Antarctic ice core sulfur isotope measurements over a complete glacial-interglacial cycle. Journal of Geophysical Research. Atmospheres, 10.1029/2003jd003513. 18. Jenkins KA, Bao H (2006) Multiple oxygen and sulfur isotope compositions of atmo- spheric sulfate in Baton Rouge, LA, USA. Atmos Environ 40(24):4528–4537. 19. Holt BD, Kumar R, Cunningham PT (1981) O-18 study of the aqueous-phase oxidation of sulfur-dioxide. Atmos Environ 15(4):557–566. 20. Holt BD, Cunningham PT, Engelkemeir AG, Graczyk DG, Kumar R (1983) O-18 study of nonaqueous-phase oxidation of sulfur-dioxide. Atmos Environ 17(3):625–632. 21. Heidenreich JE, Thiemens MH (1983) A non-mass-dependent isotope effect in the production of ozone from molecular-oxygen. J Chem Phys 78(2):892–895. 22. Heidenreich JE, Thiemens MH (1986) A non-mass-dependent oxygen isotope effect in the production of ozone from molecular-oxygen—The role of molecular symmetry in isotope chemistry. J Chem Phys 84(4):2129–2136. 23. Thiemens MH (2006) History and applications of mass-independent isotope effects. Annu Rev Earth Planet Sci 34:217–262. 24. Thiemens MH, Chakraborty S, Dominguez G (2012) The physical chemistry of mass- independent isotope effects and their observation in nature. Ann Rev Phys Chem 63:155–177. 25. Krankowsky D, et al. (2007) Stratospheric ozone isotope fractionations derived from col- lected samples. Journal of Geophysical Research. Atmospheres, 10.1029/2006JD007855. 26. Savarino J, Thiemens MH (1999) Mass-independent oxygen isotope (O-16, O-17, O-18) fractionation found in H-x, O-x reactions. J Phys Chem A 103(46):9221–9229. 27. Alexander B, Park RJ, Jacob DJ, Sunling G (2009) Transition metal-catalyzed oxidation of atmospheric sulfur: Global implications for the sulfur budget. Journal of Geo- physical Research. Atmospheres, 10.1029/2008JD010486. 28. Savarino J, Slimane B, Cole-Dai J, Thiemens MH (2003) Evidence from sulfate mass independent oxygen isotopic compositions of dramatic changes in atmospheric oxi- dation following massive volcanic eruptions. J Geophys Res 108(D21):ACH7-1–6. 17666 | www.pnas.org/cgi/doi/10.1073/pnas.1213149110 Shaheen et al.
  • 7. 29. Lyons JR (2001) Mass-independent fractionation of oxygen isotopes in Earth’s at- mosphere. Astrobiology 1(3):333–334. 30. Savarino J, Bekki S, Cole-Dai JH, Thiemens MH (2003) Evidence from sulfate mass independent oxygen isotopic compositions of dramatic changes in atmospheric oxi- dation following massive volcanic eruptions. Journal of Geophysical Research. At- mospheres 108(D21):ACH7-1-6. 31. Baroni M, Savarino J, Cole-Dai J, Rai VK, Thiemens MH (2008) Anomalous sulfur iso- tope compositions of volcanic sulfate over the last millennium in Antarctic ice cores. Journal of Geophysical Research. Atmospheres, 10.1029/2008jd010185. 32. Cole-Dai J, et al. (2009) Cold decade (AD 1810-1819) caused by Tambora (1815) and another (1809) stratospheric volcanic eruption. Geophys Res Lett, 10.1029/ 2009gl040882. 33. Halmer MM (2005) Have volcanoes already passed their zenith influencing the ozone layer? Terra Nova 17(6):500–502. 34. Robock A, Adams T, Moore M, Oman L, Stenchikov G (2007) Southern Hemisphere atmospheric circulation effects of the 1991 Mount Pinatubo eruption. Geophys Res Lett, 10.1029/2007gl031403. 35. Cole-Dai J (2010) Volcanoes and climate. Wiley Interdiscip Rev Clim Change 1(6): 824–839. 36. Cole-Dai J, Mosley-Thompson E, Qin DH (1999) Evidence of the 1991 Pinatubo volcanic eruption in South Polar snow. Chin Sci Bull 44(8):756–760. 37. Sato M, Hansen JE, McCormick MP, Pollack JB (1993) Stratospheric aerosol of optical depths, 1850-1990. Journal of Geophysical Research. Atmospheres 98(D12):22987–22994. 38. Chaochao G, Robock A, Ammann C (2008) Volcanic forcing of climate over the past 1500 years: An improved ice core-based index for climate models. Journal of Geo- physical Research. Atmospheres 113(D23):D23111, 10.1029/2008jd010239. 39. Zahn A, Franz P, Bechtel C, Grooss JU, Roeckmann T (2006) Modelling the budget of middle atmospheric water vapour isotopes. Atmos Chem Phys 6:2073–2090. 40. Jones RL, et al. (1986) The water-vapor budget of the stratosphere studied using LIMS and SAMS satellite data. Q J R Meteorol Soc 112(474):1127–1143. 41. Crutzen PJ, Lawrence MG, Poschl U (1999) On the background photochemistry of tropospheric ozone. Tellus Series A: Dynamic Meteorology and Oceanography 51(1): 123–146. 42. Lyons JR (2001) Transfer of mass-independent fractionation in ozone to other oxy- gen-containing radicals in the atmosphere. Geophys Res Lett 28(17):3231–3234. 43. Ravishankara AR (2012) Atmospheric science. Water vapor in the lower stratosphere. Science 337(6096):809–810. 44. Anderson JG, Wilmouth DM, Smith JB, Sayres DS (2012) UV dosage levels in summer: Increased risk of ozone loss from convectively injected water vapor. Science 337(6096):835–839. 45. McCabe JR, Thiemens MH, Savarino J (2007) A record of ozone variability in South Pole Antarctic snow: Role of nitrate oxygen isotopes. Journal of Geophysical Re- search-Atmospheres, 10.1029/2006JD007822. 46. Takegawa N, et al. (2003) Photochemical production of O-3 in biomass burning plumes in the boundary layer over northern Australia. Geophys Res Lett, 10.1029/ 2003gl017017. 47. Kita K, et al. (2002) Photochemical production of ozone in the upper troposphere in association with cumulus convection over Indonesia. Journal of Geophysical Research. Atmospheres, 10.1029/2001jd000844. 48. Hood LL, Soukharev BE (2005) Interannual variations of total ozone at northern midlatitudes correlated with stratospheric EP flux and potential vorticity. J Atmos Sci 62(10):3724–3740. 49. Solomon S, et al. (2010) Contributions of stratospheric water vapor to decadal changes in the rate of global warming. Science 327(5970):1219–1223. 50. Lee CCW, Savarino J, Cachier H, Thiemens MH (2002) Sulfur (S-32, S-33, S-34, S-36) and oxygen (O-16, O-17, O-18) isotopic ratios of primary sulfate produced from combus- tion processes. Tellus B Chem Phys Meterol 54(3):193–200. 51. Dominguez G, et al. (2008) Discovery and measurement of an isotopically distinct source of sulfate in Earth’s atmosphere. Proc Natl Acad Sci USA 105(35):12769–12773. 52. Holt BD, Cunningham PT, Kumar R (1981) Oxygen isotopy of atmospheric sulfates. Environ Sci Technol 15(7):804–808. 53. Kunasek SA, et al. (2010) Sulfate sources and oxidation chemistry over the past 230 years from sulfur and oxygen isotopes of sulfate in a West Antarctic ice core. Journal of Geophysical Research. Atmospheres, 10.1029/2010JD013846. 54. Sofen ED, Alexander B, Kunasek SA (2011) The impact of anthropogenic emissions on atmospheric sulfate production pathways, oxidants, and ice core Delta O-17(SO4 2− ). Atmos Chem Phys 11(7):3565–3578. 55. Tichomirowa M, Heidel C (2012) Regional and temporal variability of the isotope composition (O, S) of atmospheric sulphate in the region of Freiberg, Germany, and consequences for dissolved sulphate in groundwater and river water. Isotopes Envi- ron Health Stud 48(1):118–143. 56. Bruehl C, Lelieveld J, Crutzen PJ, Tost H (2012) The role of carbonyl sulphide as a source of stratospheric sulphate aerosol and its impact on climate. Atmos Chem Phys 12(3):1239–1253. 57. Sunilkumar SV, Parameswaran K, Thampi BV, Ramkumar G (2011) Variability in back- ground stratospheric aerosols over the tropics and its association with atmospheric dy- namics. Journal of Geophysical Research. Atmospheres, 10.1029/2010jd015213. 58. Legrand M, Wagenbach D (1999) Impact of the Cerro Hudson and Pinatubo volcanic eruptions on the Antarctic air and snow chemistry. Journal of Geophysical Research. Atmospheres 104(D1):1581–1596. 59. Legrand M, Wolff E, Wagenbach D (1999) Antarctic aerosol and snowfall chemistry: Implications for deep Antarctic ice-core chemistry. Annals of Glaciology 29(1):66–72. 60. Jourdain B, Legrand M (2001) Seasonal variations of atmospheric dimethylsulfide, dimethylsulfoxide, sulfur dioxide, methanesulfonate, and non-sea-salt sulfate aero- sols at Dumont d’Urville (coastal Antarctica) (December 1998 to July 1999). Journal of Geophysical Research. Atmospheres 106(D13):14391–14408. 61. Preunkert S, et al. (2008) Seasonality of sulfur species (dimethyl sulfide, sulfate, and methanesulfonate) in Antarctica: Inland versus coastal regions. Journal of Geo- physical Research. Atmospheres 113(D15). 62. Kennett DJ, et al. (2012) Development and disintegration of Maya political systems in response to climate change. Science 338(6108):788–791. 63. Savarino J, Alexander B, Darmohusodo V, Thiemens MH (2001) Sulfur and oxygen isotope analysis of sulfate at micromole levels using a pyrolysis technique in a con- tinuous flow system. Anal Chem 73(18):4457–4462. 64. Schauer AJ, et al. (2012) Oxygen isotope exchange with quartz during pyrolysis of silver sulfate and silver nitrate. Rapid Commun Mass Spectrom 26(18):2151–2157. 65. Ziemke JR, Chandra S, Oman LD, Bhartia PK (2010) A new ENSO index derived from satellite measurements of column ozone. Atmos Chem Phys 10(8):3711–3721. 66. Chandra S, Ziemke JR, Bhartia PK, Martin RV (2002) Tropical tropospheric ozone: implications for dynamics and biomass burning. J Geophys Res 107(D14):ACH3-1–17. 67. Oman LD, et al. (2011) The response of tropical tropospheric ozone to ENSO. Geophys Res Lett, 10.1029/2011gl047865. Shaheen et al. PNAS | October 29, 2013 | vol. 110 | no. 44 | 17667 EARTH,ATMOSPHERIC, ANDPLANETARYSCIENCES SPECIALFEATURE
  • 8. Supporting Information Shaheen et al. 10.1073/pnas.1213149110 Fig. S1. Concentration of SO4, NO3, Cl, and methane sulfonic acid (MSA) of aerosol samples extracted from snow pit samples at the South Pole station, Antarctica, from 1982 to 2002. Shaheen et al. www.pnas.org/cgi/content/short/1213149110 1 of 2
  • 9. Table S1. Oxygen-triple isotope composition and concentration profile of sulfate aerosol retrieved from a 1-m × 1-m snow pit at the South Pole Sample depth (cm) Year SO4 (ppb) δ17 O (‰) δ18 O (‰) Δ17 O (‰) Δ17 O* (‰) 18–28 2002–2001.55 43.71 3.89 5.13 1.35 1.51 28–42 2001.55–2000.92 57.50 1.29 −0.08 1.43 1.61 42–70 2000.92–2000.07 50.55 5.10 7.54 1.26 1.41 70–81 2000.07–1999.6 59.50 2.42 1.73 1.63 1.85 87–94 1999.5–1999.2 52.80 2.67 2.51 1.64 1.85 99–103 1999.08–1999.0 59.03 5.80 9.67 0.82 0.90 103–113 1999–1998.33 49.33 4.69 5.05 2.52 2.89 113–117 1998.33–1998.0 80.96 6.10 8.86 1.54 1.73 117–122 1998.0–1997.7 99.72 3.49 2.10 2.52 2.89 122–133 1997.7–1997.25 99.72 5.98 7.40 2.24 2.55 140–148 1996.9–1996.6 57.48 4.81 3.85 3.00 3.45 148–154 1996.6–1996.3 57.48 4.87 2.82 3.63 4.18 154–162 1996.3–1996 65.93 3.44 0.76 3.22 3.71 168–177 1995.84–1995.52 119.09 3.92 3.60 2.14 2.44 177–188 1995.52–1995.20 56.45 4.65 4.51 2.58 2.95 194–200 1995.0–1994.71 80.02 0.55 −1.64 1.46 1.64 232–238 1993.36–1993.18 66.08 2.25 1.26 1.70 1.92 250–256 1992.70–1992.49 76.87 3.78 1.39 3.34 3.84 256–263 1992.49–1992.31 80.15 3.18 1.73 2.50 2.86 263–268 1992.31–1992.21 127.16 4.07 2.86 2.71 3.10 268–274 1992.21–1991.69 164.64 4.82 3.57 3.12 3.58 274–279 1991.69–1991.30 160.96 3.71 1.48 3.06 3.52 279–284 1991.30–1991.0 104.14 4.42 3.09 2.94 3.37 284–289 1991.0–1990.82 75.92 4.20 2.50 2.98 3.42 289–295 1990.82–1990.52 71.87 5.40 4.27 3.29 3.78 295–305 1990.52–1990.13 70.45 5.19 3.12 3.87 4.47 305–316 1990.13–1989.71 114.14 7.39 11.87 1.28 1.43 316–320 1989.71–1989.53 110.97 3.05 2.10 2.02 2.30 320–328 1989.53–1989.28 66.42 1.39 0.79 1.04 1.15 328–338 1989.21–1988.68 40.18 2.37 −0.65 2.98 3.42 338–347 1988.68–1988.45 48.61 3.13 0.49 3.14 3.61 347–356 1988.45–1988 49.08 2.32 0.74 2.04 2.32 356–365 1988–1987.71 54.67 2.24 −0.42 2.69 3.08 365–378 1987.71–1987 64.42 2.88 0.61 2.75 3.16 378–386 1987–1986.74 36.52 3.55 1.68 2.88 3.31 386–397 1986.7–1986.4 41.11 2.77 2.08 1.79 2.03 397–402 1986.4–1986.2 74.18 1.19 3.95 1.27 1.42 402–412 1986.2–1985.8 46.22 3.84 1.73 3.32 3.82 420–428 1985.3–1984.9 80.09 1.75 −1.77 2.99 3.44 428–441 1984.9–1984.2 83.01 1.46 −1.20 2.26 2.59 441–445 1984.2–1984 60.70 3.3.7 2.97 1.93 2.20 445–456 1984–1983.5 47.64 2.71 1.46 2.22 2.54 456–465 1983.5–1983 63.01 3.67 5.87 0.70 0.76 478–490 1982.5–1982.0 99.50 3.07 2.27 2.00 2.28 490–502 1982–1981.5 44.82 4.45 4.00 2.87 3.30 525–537 1980.5–1980 45.53 5.73 10.40 0.40 0.41 Depth profile indicates initial and final ice layers of the composite samples used to measure oxygen-triple isotopic composition. Δ17 O*, corrected for high- temperature pyrolysis (1,050 °C) using a quartz cup as Δ17 O* = 1.17 × Δ17 O(quartz) − 0.06 (1). 1. Schauer AJ, et al. (2012) Oxygen isotope exchange with quartz during pyrolysis of silver sulfate and silver nitrate. Rapid Commun Mass Spectrom 26(18):2151–2157. Shaheen et al. www.pnas.org/cgi/content/short/1213149110 2 of 2