ArticlePDF Available

Dynamic and thermodynamic influences on precipitation in Northeast Mexico on orbital to millennial timescales

Authors:

Abstract and Figures

The timing and mechanisms of past hydroclimate change in northeast Mexico are poorly constrained, limiting our ability to evaluate climate model performance. To address this, we present a multiproxy speleothem record of past hydroclimate variability spanning 62.5 to 5.1 ka from Tamaulipas, Mexico. Here we show a strong influence of Atlantic and Pacific sea surface temperatures on orbital and millennial scale precipitation changes in the region. Multiple proxies show no clear response to insolation forcing, but strong evidence for dry conditions during Heinrich Stadials. While these trends are consistent with other records from across Mesoamerica and the Caribbean, the relative importance of thermodynamic and dynamic controls in driving this response is debated. An isotope-enabled climate model shows that cool Atlantic SSTs and stronger easterlies drive a strong inter-basin sea surface temperature gradient and a southward shift in moisture convergence, causing drying in this region.
This content is subject to copyright. Terms and conditions apply.
Article https://doi.org/10.1038/s41467-023-37700-9
Dynamic and thermodynamic inuences on
precipitation in Northeast Mexico on orbital
to millennial timescales
Kevin T. Wright
1
, Kathleen R. Johnson
1
, Gabriela Serrato Marks
2
,
David McGee
2
, Tripti Bhattacharya
3
, Gregory R. Goldsmith
4
,
Clay R. Tabor
5
, Jean-Louis Lacaille-Muzquiz
6
, Gianna Lum
1
&
Laura Beramendi-Orosco
7
The timing and mechanisms of past hydroclimate change in northeast Mexico
are poorly constrained, limiting our ability to evaluate climate model perfor-
mance. To address this, we present a multiproxy speleothem record of past
hydroclimate variability spanning 62.5 to 5.1 ka from Tamaulipas, Mexico. Here
we show a strong inuence of Atlantic and Pacic sea surface temperatures on
orbital and millennial scale precipitation changes in the region. Multiple
proxies show no clear response to insolation forcing, but strong evidence for
dry conditions during Heinrich Stadials. While these trends are consistent with
other records from across Mesoamerica and the Caribbean, the relative
importance of thermodynamic and dynamic controls in driving this response
is debated. An isotope-enabled climate model shows that cool Atlantic SSTs
and stronger easterlies drive a strong inter-basin sea surface temperature
gradient and a southward shift in moisture convergence, causing drying in this
region.
A majority of climate models project that Northern Mexico will
become drier in the future, but the spatial distribution and magni-
tude of drying is poorly constrained at present due to a lack of model
agreement especially at the local scale1,2. Improving hydroclimate
projections for Northern Mexico is critical given the substantial
social, economic, and ecological impacts that shifts in mean pre-
cipitation or precipitation extremes can have in the region. For
instance, severe droughts in the past have led to agriculture
disruptions3, national food shortages, and have been linkedto surges
in international immigration4. Records of past hydroclimate can
provide critical constraints on the dynamical drivers of regional
precipitation variability, and contribute to improved climate
projections5, yet few records exist in Northern Mexico. Specically,
paleoclimate records can contribute to evaluating and improving
climate models used for projecting future hydroclimate, by helping
to: (1) Constrain the magnitude and timing of precipitation change in
response to external forcings and internal ocean-atmosphere
variability6, (2) Evaluate the spatial pattern of regional precipitation
changes in models7, and (3) Provide robust data for proxy-model
comparison studies, which may help reveal model biases8. Spe-
leothems are ideally suited archives of past hydroclimate due to their
precise U-Th based age models and the multiple hydrologically sen-
sitive proxies they contain.Despite the prevalence of limestone karst
landscapes in Northeast (NE) Mexico, though, there has only been
one published speleothem record spanning the last millennium from
the region9.
Received: 10 June 2021
Accepted: 28 March 2023
Check for updates
1
Dept. of Earth System Science, University of California, Irvine, 3200 Croul Hall, Irvine, CA, USA.
2
Department of Earth, Atmospheric and Planetary Sciences,
Massachusetts Institute of Technology, Cambridge, MA, USA.
3
Department of Earth Sciences, Syracuse University, Syracuse, NY, USA.
4
Schmid College of
Science and Technology, Chapman University, Orange, CA, USA.
5
Department of Geosciences, University of Connecticut, Storrs, CT, USA.
6
Independent
researcher, Ciudad Mante, Tamaulipas, Mexico.
7
Instituto de Geología, Universidad Nacional Autónoma de México, Ciudad Universitaria, Ciudad de,
México, xico. e-mail: ktwright@uci.edu;kathleen.johnson@uci.edu
Nature Communications | (2023) 14:2279 1
1234567890():,;
1234567890():,;
Content courtesy of Springer Nature, terms of use apply. Rights reserved
The modern climatology of NE Mexico (Fig. 1) is dominated by the
Caribbean Low-Level Jet (CLLJ), which transports moisture from the
Atlantic Ocean and Caribbean Sea to the Atlantic slope of Mexico and
Central America during boreal summer10,11. The strengthening of the
summer CLLJ, dened as an increase in low-level easterly wind velocity
off the coast of northern South America, is driven by increased solar
heating and a northward migration of the Intertropical Convergence
Zone (ITCZ)12. The CLLJ, however, also strengthens in February, driven
by an intensied meridional pressure gradient linked to heating over
South America. It is important to note that only the CLLJ maximum in
Boreal summer is associated with increased moisture ux and pre-
cipitation over Mexico. However, transient weather events like tropical
cyclones can alsobring heavy rainfall to the region in late summer and
early autumn13.
On orbital timescales, insolation variations dominated by pre-
cessionhave been proposed to impact regional hydroclimate in similar
ways as the seasonal ITCZ migration, with strengthening of the CLLJ
and precipitation increases occurring during Northern Hemisphere
summer insolation (NHSI) maxima when the ITCZ migrates north.
Available proxy records from NE Mexico show unclear evidence for a
strong insolation control on regional hydroclimate, though. For
instance, while a sediment record from the El Potosi Basin in NE Mexico
demonstrates a strong positive correlation of runoff to NHSI14,other
studies from the region have found a limited role for NHSI and sug-
gested that autumn or spring insolation may be more important dri-
vers of hydroclimate15,16. In contrast, the role of insolation in NW
Mexicoand Southern Mexico is much better understood, withmultiple
records demonstratinga strong positive correlation between NHSI and
precipitation via alteration of the North American Monsoon1721 and
shifts in the ITCZ22.
On millennial timescales, precipitation variability in NE Mexico
has also been linked to the strength of the CLLJ. For instance, Roy
et al.14 found decreased Ti concentration in lake sediments from El
Potosi Basin during Heinrich Stadial (HS) 1, which were interpreted as
reecting reduced precipitation caused by a weakening of the CLLJ as
the ITCZ shifted south. However, this interpretation may be incon-
sistent with modern dynamics of the CLLJ, which actually strengthens
during Boreal winter (February) when the ITCZ migrates south23,24.
Using the seasonal ITCZ migration as an analog, it is possible that the
CLLJ could actually strengthen during HS events, suggesting some
other factor, such as sea surface temperatures (SSTs), may play a more
important role indriving millennial-scale hydroclimate variability in NE
Mexico.
While previous records have not shown a strong SST control on
mean precipitation in NE Mexico on orbital or millennial timescales,
several previous paleoclimate reconstructions and modeling studies
have linked hydroclimate variations across Mexico, Central America
and the Caribbean to changes in SSTs. Although NE Mexico is outside
the nuclear Mesoamericanregion, due to thecultural and climatic links
of NE Mexico to Southern Mexico and Central America, we will here-
after refer to this entire region as Mesoamerica. Tree ring and climate
modeling studies have shown that both Pacic and Atlantic SSTs exert
a strong precipitation control across Mesoamerica on interannual to
multidecadal timescales2529. Over the Common Era (last 2000 years)
changes in Atlantic SSTs have been associated with a strong, out-of-
phase, dipole precipitation pattern between northern and southern
Mesoamerica (See Fig. 5from28). However, a recent speleothem study
has suggested the dipole precipitation pattern is biased towards winter
precipitation, and precipitation on annualtimescales, and longer, may
respond more in-phase throughout the region9. Unfortunately, the
impact of SST variations on Mesoamerican hydroclimate patterns on
millennial and orbital timescales is poorly constrained due to the
paucity of records. For instance, while records from southern Mesoa-
merica consistently show drying during HS events22,29, lake sediment
records from northern Mexico demonstrate wet, dry and neutral
responses3032. Although the inconsistency of the northern Mexico
records may simply be driven by uncertainties in the age models30,or
the inuence of tropical Paciccyclones
32, the sparse paleoclimate
record from NE Mexico hinders our ability to assess the spatial pattern
of precipitation response to SST changes on orbital and millennial
timescales.
In addition to the spatial response of precipitation to SST varia-
bility, there is some evidence to suggest SSTs may impact extreme
precipitation in NE Mexico. For instance, increased clay mineral con-
centration (Al+Si+K + Fe/Ca), interpreted to reect increased water-
shed erosion from high intensity rainfall, in lake sediments from the
Cieneguilla Basin and the Sandia Basin in NE Mexico have linked per-
iods of increased tropical cyclones to warm Gulf of Mexico SSTs during
Fig. 1 | Summer (JJAS) climatology and nearby paleoclimate records. Map of
regional precipitation and magnitude of low-level (850 mb) winds using PERSIANN
precipitation data106 and winds from MPI-ESM-Historical107. Nearby records include
(1) an oceansediment core fromthe Gulf of Mexico59 and the (2) Florida Straight49,
speleothem records from (3) Cuba46, (4) Southern Mexico22 and (8) Costa Rica108,
additional ocean sediment cores from the (6) Caribbean Sea58, and (7) Cariaco
Basin74, and a lake sediment core (5) from Guatemala50.
Article https://doi.org/10.1038/s41467-023-37700-9
Nature Communications | (2023) 14:2279 2
Content courtesy of Springer Nature, terms of use apply. Rights reserved
the mid-Holocene and Bølling-Allerød15,30. However, notably, these
shifts in precipitation extremes were not clearly associated with shifts
in mean precipitation and, overall, the correlation between GOM SSTs
and NE Mexico precipitation on orbital and millennial timescales has
been shown to be weak or inconsistent14. Given these discrepancies,
additional interglacial-glacial records of hydroclimate variability are
needed to further explicate the role of SSTs on regional precipitation
change.
Specically, the relative importance of CLLJ strength and SSTs in
driving regional hydroclimate across Mesoamerica remains an open
question. Robust paleoclimate records are needed to help resolve this
issue. To this end, we present a new decadal-resolution, multi-proxy
(δ18O, δ13C, Mg/Ca) speleothem record from Tamaulipas, Mexico that
spans 62.5 to 5.1 ka. Our results show strong hydrologic responses to
key millennial-scale events including the Younger Dryas and Heinrich
Stadials 1, 36, and a muted response to NHSI. Furthermore, we utilize
results of a freshwater-forcing experiment conducted with an isotope-
enabled climate model to investigate the importance of dynamic and
thermodynamic controls on NE Mexico precipitation.
Results and discussion
Multiproxy reconstruction of hydroclimate variability in NE
Mexico
We present a ~57,000 year record of hydroclimate utilizing oxygen
isotopes (δ18O), carbon isotopes (δ13C), and trace elements (Mg/Ca)
from a 78 cm-long candle-shaped stalagmite, CB2 (Fig. 2,SeeSINote1).
CB2 was collected from Cueva Bonita (23°N, 99°W; 1071 m above sea
level), located in the highlands of the Sierra Madre Oriental in the
northeastern Mexico state of Tamaulipas (Fig. 1;SeeSINote2).The
climate of this region is characterized by cool-dry winters and warm-
wet summers13 (Fig. S1), with a precipitation maximum in summer
(July) driven by warm North Atlantic SSTs and anintensication of the
Caribbean Low-Level Jet13 (Fig. 1). The stable isotope and trace element
proxy data are tied to a U-series age-depth model, constrained by 33
230Th234U ages and the mean of 2000 Monte-Carlo simulations using
the age-modeling software COPRA33. The age model indicates the
sample formed continuously from 62.5 to 5.1 ka with an average
growth rate of ~14 μm/yr, and an average temporal resolution of
~36 years (Fig. 2). This represents the highest resolution, continuous
paleoclimate proxy record in Mexico over this time-period.
Previous work in Southern Mexico has consistently interpreted
the oxygen isotope signature of precipitation (δ18O
p
)asreective of
precipitation amount22,34,35,36, with greater amounts of rainfall asso-
ciated with more negative δ18O
p
values. Risi et al.37 argues the amount
effect dominates in the tropics due to high rainfall rates, which limits
isotopic exchange with near-surface moisture. Furthermore, more
recent analysis in the nearby mid-latitudes has interpreted δ18O
p
to
reect shifting moisture source, temperature, relative proportions of
stratiform vs convective precipitation, seasonality, and shifts in thun-
derstorm size and duration3840. To improve our understanding of
modern precipitation isotope systematics at our site, we have used an
array of modeling and observational data. We analyzed moisture
bearing air trajectories over a 15-year period (See SI Note 6) which
demonstratethatmoistureisconsistentlysourcedfromtheGulfof
Mexico and Caribbean Sea. An observational record of precipitation
δ18O from approximately 1 km from the cave, established aspart of this
study, suggests the δ18O of monthly precipitation is strongly depen-
dent on precipitation amount (p<0.01; r2=0.88; slope=2/
100 mm) (Fig. S2). This correlation is further supported by isotope-
enabled GCM simulations of precipitation spanning the last 40 years,
which suggest lower δ18O
p
values primarily reect an increase in
regional precipitation amount (Fig. S2). Given the stable cave envir-
onmental conditions in CuevaBonita (Fig. S3), theminimal inuence of
cave variability (temperature, evaporation, or degassing) on spe-
leothem δ18O as demonstrated by a simple proxy system model
(Fig. S4), and the fact that δ18O values of calcite precipitated on glass
plates from Cueva Bonita areclose to oxygenisotopic equilibriumwith
drip waters, suggests that speleothems from this cave preserve varia-
tions in the δ18O values of ancient drip water and precipitation (See SI
Note 8). We therefore interpret CB2 speleothem δ18Oasreective of
regional precipitation amount in NE Mexico.
While δ18O is often a proxy for large scale atmospheric processes,
the controls of δ13C are more localized and record changes in overlying
vegetation (amount and type), soil respiration, temperature, and CO
2
degassing within caves and associated with prior calcite precipitation
(PCP) in the epikarst41,42. Despite these complex controls, speleothem
Fig. 2 | Stalagmite CB2, age-depth model and Mg/Ca, δ18Oandδ13C results.
aResults of 1578 stable isotope and 789 trace element measurements. δ18Oistop
and blue, δ13C is central and in orange, Mg/Ca is in bottom and green. Dates with
associated uncertaintiesare below Mg/Ca. Heinrich Stadialshighlightedin light red.
bCB2 Age-Depth Model constructed using 2000 Monte-Carlo simulations via the
age-depth modeling software COPRA and 33 U-Th ages. Uncertainty in age-depth
model indicated by gray shading, uncertainty in U-Th ages indicated by red error
bars. cSample CB2 after being cut and polished.
Article https://doi.org/10.1038/s41467-023-37700-9
Nature Communications | (2023) 14:2279 3
Content courtesy of Springer Nature, terms of use apply. Rights reserved
δ13Cvalueshavebeenincreasinglyshowntoreect local water balance,
with PCP and soil/vegetation changes all leading to higher δ13Cvalues
during drier periods9,4143. We suggestthat a major driver of δ13CinCB2
is PCP, which occurs when there is reduced local water balance and is
the result of enhanced CO
2
degassing and calcite precipitation in the
epikarst41. To further constrain the mechanisms of CB2 δ13C variability,
we conducted Mg/Ca analyses, which can also reect PCP. In addition
to the preferential loss of 12C from the drip water dissolved inorganic
carbon pool during CO
2
degassing, PCP leads to the preferential
uptake of Ca2+ leaving the remaining drip waters, and the speleothem,
enriched in trace elements (Mg2+
). CB2 δ13C and Mg/Ca ratios exhibit
similar variability throughout the late-Pleistocene and weakly correlate
(autocorrelation corrected r=0.39, p< 0.01) throughout this time
period, suggesting PCP inuences both Mg/Ca ratios and δ13C. While
we cannot rule out the additional inuence of other factors, such as
soil/vegetation changes, on speleothem δ13C or temperature changes
on Mg/Ca, our multi-proxy approach allows for a more robust inter-
pretation of our proxy record.
The CB2 δ18O record is remarkably smooth over the glacial period,
is punctuated by a several prominent millennial scale variations during
the glacial and deglacial periods (HS1, Bølling-Allerød, Younger Dryas),
and thereis a clear ~1decreasefrom the Last Glacial Maximum (LGM;
~21 ka) to the Holocene (after correcting for global ice volume). In
contrast to speleothem records from Asia and South America44,45,the
CB2 δ18O record shows no clear precessional signal on orbital time-
scales. While increased precipitation in the early-Holocene has been
recorded elsewhere in Mesoamerica22,46, the effects of changing
glacial-interglacial cave temperatures could easily explain the 1
glacial-interglacial shift observed in our δ18O record (See SI Note 10).
The lack of orbital variability provides a relatively stable and un-
varying background, which large amplitude millennial scale variability
is superimposed upon. The δ18O time series consistently exhibits large
positive excursions of ~2during Heinrich Stadials, indicating shifts
towards drier conditions. Speleothem δ18Ovalues(n=1578) range
from 1.29to 6.30(VPDB) with the most 18O-enriched samples
occurring during Heinrich Stadial 1 (HS 1) at ~16.8 ka (Fig. 2). However,
excursions toward higher δ18O values are also noted during 5047 ka,
4342 ka, 3128 ka, 1815 ka, and 1210 ka, corresponding to HS 5, 4, 3,
1 and the Younger Dryas (YD), respectively (Fig. 2). While the timing of
HS 4 in our record is slightly older (~4240) than the expected age
(~4038 ka)47, the offset is possibly attributed to a relatively large
uncertainty in our age model around this time (Fig. 2, Fig. S5, Table S1).
The CB2 δ13C data co-varies with δ18O(r=0.53, p<0.01) and is
similarly dominated by large amplitude millennial variations during
the glacial, and a ~3decrease across the deglaciation (Fig. 2). A lake
sediment record from the region suggests the decrease in δ13Ccould
reect a shift from C
4
to C
3
dominant vegetation15, but this shift could
also potentially reect some combination of decreased PCP, increased
soil respiration or vegetation intensity, and/or decreased water-rock
interaction during the Holocene42,48. Millennial-scale shifts in δ18Oare
reproduced in the δ13C record, consistent with decreased local water
balance during the Younger Dryas and Heinrich Stadials. Increases in
δ13C were as large as 3.94during HS 5. However, not all changes in
δ13C were as dramatic, such as during HS 3 the shift in δ13C was as subtle
as 1.04. We suggest the varying responses recorded by CB2 proxies
may reect real differences between individual Heinrich Stadials49,
though some inuence of complex proxy controls may also play a role.
The CB2 Mg/Ca values exhibit similar variations on millennial
timescales as the stable isotopes, particularly to δ13C. Mg/Ca values
increase above average glacial values (27 mmol/mol) to 37 mmol/mol
during the YD, 68 mmol/mol during HS1, 34 mmol/mol during HS3,
37 mmol/mol during HS4, 35 mmol//mol during HS5, and to 33mmol/
mol during HS6. Interestingly, the response of Mg/Ca ratios diverge
from that of the stable isotopes during the deglaciation, with an
increase from 27 to 40 mmol/mol (Fig. 2). While this could potentially
indicate that PCP increased during the Holocene-Pleistocene transi-
tion, the δ13Candδ18O evidence both point towards wetter conditions,
consistent with regional records22,46,50, which should lead to less PCP
rather than more. We therefore suggest the increasing Mg/Ca trend
may instead reect the inuence of temperature on Mg partitioning
into calcite51,52 (see SI Note 5). During the glacial period, Mg/Ca data is
consistent with a PCP control, especially during HS2-6, as evidencedby
asignicant positive correlation between δ13CandMg/Ca
(r=0.510.77, p0.01). We therefore interpret higher Mg/Ca ratios
and enriched 13Ctoreect enhanced PCP during periods of reduced
local water balance (See SI Note 4).
Potential forcings of precipitation on orbital timescales
Orbital-driven variations in insolation have been invoked to explain
widespread moisture variations in the broader region of
Mesoamerica22, as well in NE Mexico14, but most of these records only
span one precession cycle. The CB2 record, which spans ~2.5 preces-
sion cycles and extends ~25,000 years beyond the oldest lake record
from NE Mexico30, thus offers a unique opportunity to further con-
strain the impacts of insolation on precipitation and local water bal-
ance. While summer insolation has been proposed to drive
precipitation in NE Mexico via a northward shift in the ITCZ and a
strengthening of the CLLJ14,andinNWMexicoviaanintensied
NAM17,18, our record does not demonstrate a strong correlation to
summer insolation. However, our record is not alone in that a growing
number of records across Mesoamerica have also found a weak cor-
relation to NHSI, suggesting the inuence of autumn, winter, or spring
insolation as the dominant driver of hydroclimate variability on orbital
timescales. For instance, Roy et al.15,31,53, have attributed a strong cor-
relation between increased watershed erosion and/or runoff in north-
ern Mexico to autumn insolation through increased tropical cyclone
and hurricane activity. Furthermore, water scarcity in North Central
Mexico recorded by increased authigenic calcite precipitation in a
sediment core from ephemeral Lake Santiaguillo, has been linked to
peaks in spring insolation16. Even winter insolation has been linked to
an extended wet season in speleothem δ18O records from Santo Tomás
Cave in Cuba and Terciopelo Cave in Costa Rica46,54. However, com-
parison of CB2 δ18O to Northern Hemisphere insolation from different
seasons demonstrates consistently weak correlations over the last ~57
ka (Fig. S6). While CB2 δ18O seemingly changes with insolation over the
last 20,000 years (Fig. 3, Fig. S6), exhibiting an in-phase response with
autumn insolation or a lagged response to summer insolation, this
relationship does not continue throughout the late-Pleistocene. This
lack of a consistent insolation pattern suggests that other factors such
as global ice volume, radiative forcing from atmospheric pCO
2
,orSSTs
may play a more direct role in explaining glacial-interglacial hydro-
climate variability in NE Mexico compared to insolation alone.
The CB2 δ18O time series shows much stronger similarity to
regional and global temperature records, indicating that precipitation
at our study site may be more sensitive to thermodynamic controls on
orbital timescales. Comparison with the Greenland ice core δ18O
record (ref. 55; Fig. 3), shows the CB2 record is dominated by relatively
cool and/or dry glacial conditions, as evidenced by relatively positive
δ18O values from 62.5 to 20 ka, with a shift towards warmer and/or
wetter conditions during the deglacial and Holocene. Superimposed
on this orbital-scale trend are millennial scale shifts towards more
positive δ18O values, indicating even drier conditions during Heinrich
Stadials and the Younger Dryas. These trends are reproduced by the
CB2 δ13C and Mg/Ca records. The CB2 record exhibits a much closer
relationship with atmospheric pCO
2
than with insolation (r=0.61,
p<0.05, Fig. 3b).
However, the observed correlation between δ18OandpCO
2
is
primarily associated with the deglaciation, evident in the large lag
times, different topology, and an insignicant correlation over the
glacial period between 2062.5 ka (r=0.41, p= 0.11).
Article https://doi.org/10.1038/s41467-023-37700-9
Nature Communications | (2023) 14:2279 4
Content courtesy of Springer Nature, terms of use apply. Rights reserved
In contrast to insolation and pCO
2
,CB2δ18O exhibits a much
better correlation (r=0.49 to 0.73) with regional SSTs, including the
Tropical North Atlantic, the Gulf of Mexico, the Caribbean Sea, and the
Tropical Eastern Equatorial Pacic(Fig.3c, Fig. S75659). Other records
from the region have also shown a strong correlation to changes in
local SSTs on orbital timescales. For instance, increased rainfall in SE
USA indicated by increased Mississippi River discharge has been
attributed to warmer SSTs60. Also, stalagmite records from Cuba,
Guatemala, and Puerto Rico have also attributed precipitation varia-
bility to Gulf of Mexico and Caribbean Sea SSTs on glacial-interglacial
timescales46,61,62. Even lake sediment records from NE Mexico have
identied SSTs asan important driver of climate on orbital timescales,
but this record did not maintain a strong correlation to SSTs into the
Holocene, consequently, more emphasis was placed on insolation,
shifts in the ITCZ, and a stronger CLLJ14.
While all SST records co-vary with CB2 δ18O on orbital timescales,
only the Tropical Pacic and Atlantic SSTs exhibit similar variability on
millennial timescales. The lack of millennial scale variability in some
records is possibly driven by slow sedimentation rates58, seasonal
biases59, or intra-Gulf of Mexico SST variability from Mississippi River
discharge or loop current strength63. Regardless, the CB2 δ18Orecord
suggests precipitation is responsive to broad-scale, basinwide changes
in Tropical North Atlantic and Eastern PacicSSTs.
Previous work has attributed cooler Tropical North Atlantic SSTs
to reduced precipitation in southern Mexico primarily by reducing
boundary layer moisture and convective activity28,however,warmer
Atlantic SSTs were thought to decrease moisture transport to North-
ern Mexico from a weakened inter-basin pressure gradient and CLLJ23.
The CB2 record, instead, demonstrates precipitation responds simi-
larly to southern Mexico via a positive correlation to Atlantic SST
variability. This nding is consistent with a recent interannual spe-
leothem record from the Common Era, which also suggests cooler
Atlantic SSTs drive decreased precipitation in theregion9.Additionally,
the sensitivity of CB2 δ18O to PacicSSTs(Fig.3) has been shown in
other records to alter precipitation throughout Northern Mexico,
primarily through shifts in the sub-tropical jet stream and reductionsin
winter storms9,64,65. We therefore suggest glacial-interglacial variations
in sea surface temperatures, indirectly reecting orbital forcing and
associated feedbacks, are the most important driver of precipitation
variability in NE Mexico on orbital timescales.
Millennial-scale droughts linked to strengthened CLLJ and
lower SSTs
While previous records have suggested a variable response of North-
ern Mexico precipitation to millennial-scale AMOC changes30,66,CB2
proxies consistently record drying in northeast Mexico during Hein-
rich Stadials, with the exception of HS 2 (Fig. 2). Roy et al.14 proposed
that a southward shift of the ITCZ drove a weakening of the CLLJ
during Heinrich Events, thus decreasing the northward moisture
transport and subsequently leading to dry conditions in NE Mexico.
However, on seasonal timescales, the CLLJ actually strengthens when
the ITCZ shifts south during Borealwinter. Furthermore, other records
in southern Mesoamerica and the circum-Caribbean region point
towards SSTs as a more important inuence on regional
hydroclimate46,61,62. Clearly, a better understanding of the dynamical
and thermodynamical inuences on precipitation change during
Heinrich Stadials is needed in this region.
To evaluate the underlying climate dynamics associated with
drying during Heinrich Stadials, we analyzed results of an isotope-
enabled Earth System Model simulation (iCESM1;67) forced with
freshwater added to the North Atlantic on a glacial background state
(See methods). A comparison of pre-industrial iCESM1 precipitation to
merged model-observation data from the Global Precipitation Clima-
tology Project (GPCP) demonstrates iCESM1 correctly replicates the
overall pattern of precipitation in Mesoamerica, but overestimatesthe
magnitude of precipitation change in the Eastern Tropical Pacic
(Fig. S8). Previous work hasdemonstrated this is likely driven by model
sensitivity to orography and poorly resolved topography in Central
America, however, the response of precipitation in higher subtropical
latitudes has been shown to more closely match simulations68.Our
discussion will focus on the underlying dynamics driving patterns of
precipitation change rather than specic changes in precipitation
amount (i.e. mm/day).
Model results show signicant SST cooling in the North Atlantic
(Fig. 4a) during Heinrich Stadials, relative to the LGM. The magnitude
of cooling is large but is supported by realistic proxy-based estimates.
For instance, two sediment cores reconstructing summer SSTs from
monospecic planktonic foraminifera suggest upwards of 10 degrees
of cooling during HS1 compared to the LGM69, and annual SST simu-
lations of the Tropical North Atlantic which show more moderate
cooling (24 °C) are in close agreement with regional SSTs recon-
structions (Fig. 3). Cooling is simultaneously driven by a decrease in
heat transport by a weakened western Gulf Stream and the positive
wind-evaporation-SST feedback via an enhanced Bermuda High70.
Moreover, cooling in the tropical Atlantic (~6 °C) is considerably
stronger than cooling in the tropical Pacic (~1 °C), creating an East-
West inter-basin temperature and sea level pressure (SLP) gradient
(Fig. 4a). A stronger Bermuda High results in an intensication of the
tropical easterlies which funnel into the Caribbean Sea (Fig. 4c), and
the inter-basin temperature and sea level pressure gradient further
magnify the ow of winds across the Isthmus of Tehuantepec (Fig. 4c).
However, unlike the seasonal intensication of the tropical easterlies
leading to increased rainfall, resulting wind anomalies combine with
considerably cooler SSTs to signicantly reduce vertically integrated
moisture ux (Fig. 4c) and precipitation amount (Fig. 4b) over most of
Mesoamerica. While circulation changes likely modify precipitation
elsewhere in North America, we focus on modeled precipitation
changes solely in Mesoamerica, where changes are statistically sig-
nicant (Fig. 4). This is consistent with other hosing experiments,
which show divergent responses of rainfall to North Atlantic fresh-
water forcing71.
The modeled reduction in regional precipitation (Fig. 4b) is cou-
pled withan increase in precipitation δ18OinNEMexico(Fig.4d), which
Fig. 3 | Comparison of CB2 δ18O to various potential forcings. a Autumn (SON,
r=0.48, p<0.14, orange) and Summer (JJA, r=0.26 p= 0.46 black) insolation.
bAtmospheric pCO
2
(r=0.61, p<0.01, maroon109). cSSTs from Gulf of Mexico
(r=0.52, p<0.03, silver
59), Caribbean (r=0.59, p<0.05,limegreen
58), Tropical
N. Pacic(r=0.55, p<0.03,magenta
56) and Tropical N. Atlantic (r=0.73, p<0.01,
teal57). dGreenland temperatures (r=0.61, p< 0.01, indigo110).
Article https://doi.org/10.1038/s41467-023-37700-9
Nature Communications | (2023) 14:2279 5
Content courtesy of Springer Nature, terms of use apply. Rights reserved
is reasonably consistent with the 1.5increase observed in
CB2 speleothem δ18OduringHS1(Fig.2). The freshwater added to the
model experiment (0.5 Sv for 100 years) is most similar to the fresh-
water released during HS 147, upwards of 66% more than during other
Heinrich Stadials and explains why we see a stronger drying during HS1
than during other stadials. Although through different mechanisms,
the Younger Dryas is estimated to exhibit similar SST cooling as
Heinrich Stadials, explaining the similarmagnitude of drying indicated
by δ13C. Moreover, the CB2 record suggests that changes in AMOC
work in both directions, with increased precipitation leading to a 2
excursion in δ18O, a 3shift in δ13C and, a decrease of 15 mmol/mol of
Mg/Ca during the Bølling-Allerød when AMOC was strengthened and
SSTs are warmer72.
In juxtaposition to previous proxy studies from Northern Mexico
which speculated a weakening of tropical easterlies and the CLLJ in
response to Heinrich Stadials, our climate model results demonstrate
tropical easterlies in fact strengthened during reductionsin AMOC and
a southward shifted ITCZ, similar to the dynamical response of the
modern CLLJ during boreal winter. Moisture budget analysis suggests
drying in southern Mesoamerica is primarily driven by the strength-
ening of the easterlies and an intensied zonal inter-basin SST and SLP
gradient causing stronger winds across the Isthmus of Tehuantepec
and a southward migration of moisture convergence (Fig. S9). This
reduction in precipitation is part of a larger scale response in which an
inter-hemispheric SST and SLP gradient leads to anomalous cross-
equatorial winds and a southward shift in ITCZ precipitation. Although
models currently do not capture tropical cyclones, stronger low-level
winds cause an amplication of vertical wind shear and is therefore
also likely associated with a reduction in transient weather events,
including tropical cyclones and hurricanes9,13. Stronger tropical east-
erlies also signicantly reduce Tropical Atlantic SSTs through the
positive wind-evaporation-SST feedback loop. The reduction in Tro-
pical Atlantic SSTs is alsoknown to further reduce summer convection
in the region73, thereby leadingto thermodynamic driven reductions in
precipitation throughout regions of Northern Mexico (Fig. S9). We
therefore demonstrate the combination of stronger easterlies and
cooler SSTs during Heinrich Stadials are particularly important for
reducing convection in NE Mexico and highlight the possible
mechanisms in which colder SSTs could contribute to the strong
similarities observed between CB2 δ18O and regional SST reconstruc-
tions on millennial to orbital timescales.
Our work also suggests thatprecipitation in northern Mexico is in
phase with southern Mesoamerica and the Caribbean. We nd CB2
exhibits similar patterns of variability to speleothem δ18Orecordsin
Costa Rica54 and Cuba46, Cariaco Basin sediment records74,andmag-
netic susceptibility records from Guatemala50 on millennial timescales
(Fig. 5). Widespread regional drying during Heinrich Stadials has pre-
viously been attributed to a southward displaced ITCZ and/or cooler
Fig. 4 | iCESM1 simulation of Heinrich Stadials compared to LGM over North
America. a Annual temperature and Sea Level Pressure changes. bAnnual pre-
cipitation change, with statistically signicant changes contoured by dashed gray
lines. cAnnual changes in total column precipitable water and 850 mb winds.
dAnnual changes in the stable oxygen isotope ratio of precipitation.
Article https://doi.org/10.1038/s41467-023-37700-9
Nature Communications | (2023) 14:2279 6
Content courtesy of Springer Nature, terms of use apply. Rights reserved
SSTs in response to a weakening of AMOC46,61,62,75,76. However,previous
work has cautioned against extrapolating paleo precipitation records
to large-scale inter-hemispheric atmospheric changes, such as mer-
idionalshifts in the ITCZ, because data synthesisand modeling analysis
suggests a regionally variable precipitation response62,77.
Through the addition of our paired proxy and climate model
analysis, we suggest the importance of the more localized Atlantic-
Pacic inter-basin SST and SLP gradient on driving easterly wind
anomalies, which subsequently reduces local SSTs,vertical wind shear,
and convective activity in the region. The culmination of these pro-
cesses drives moisture convergence to shift southward, as supported
by our modeling results and previous paleo records78. Ultimately, the
intensied inter-basin gradient causes spatially ubiquitous drying
across Mesoamerica and the Caribbean (Figs. 4,5). Although cold
events like Heinrich Stadials are never perfect analogs for future pre-
cipitation change, these new results suggest precipitation across
Mesoamerica and the Caribbean is highly sensitive to the Atlantic-
Pacic SST gradient, and could exhibit a similar broadly coherent
response to changes in AMOC in the future.
Wet conditions in NE Mexico during the Last Glacial Maximum
and HS 2
The LGM is an important time period for highlighting mechanisms of
precipitation change and evaluating climate model performance5,79.
The CB2 record provides an additional record of high-resolution pre-
cipitation change before, during, and after the LGM, and is well suited
for future model-proxy comparison studies. In contrast to the other
Heinrich Stadials, the CB2 δ13C and Mg/Ca records suggest an exten-
ded period of increased local water balance at our study site during HS
2(~24ka;Fig.2), just before the LGM. This is in contrast to other proxy
records from Lake Péten Itzá, Guatemala50,80, Santo Tomás Cave,
Cuba46,andTerciopeloCave,CostaRica
54, which all show the expected
response of drying during HS 2 (Fig. 5). While the CB2 response could
potentially reect a highly localized signal or be impacted by non-
climatic proxy controls, the clear covariation between δ13CandMg/Ca
during this event does suggest it was characterized by increased water
balance. There are three potential mechanisms we consider that could
explain the wetter conditions in NE Mexico during HS 2, which occurs
around the time of the LGM: (1) Increased winter precipitation derived
from the Pacic winter storm track during the LGM, (2) A weaker HS 2
event which NE Mexico hydroclimate is more sensitive to than other
regions, and/or (3) increased water balance due to colder tempera-
tures during HS 2 and the LGM.
An intensied Pacic winter storm track during the LGM has been
frequently evoked to explain increased precipitation across the Great
Basin, Southwest US, and Northern Mexico, but the spatial extent of
the winter storm track has been previously disputed81,82. Precipitation
sourced from the Pacic would have an isotopic composition heavily
depleted in 18O due to the more distal moisture source83, which is not
observed in CB2 δ18O during this time period. In fact, speleothem δ18O
demonstrates no signicant change during HS 2 while both δ13Cand
Mg/Ca ratios (Fig. 2) decrease, suggesting anomalous increases in local
water balance. PMIP3 model simulations of the LGM further support
this interpretation, showing an increase in the magnitude of the CLLJ
but no increase in winter precipitation (Figs. S10, 11). Therefore, we
rule out contribution of rainfall from an enhanced winter storm track
in NE Mexico during HS 2 or the LGM.
There is still some debate about whether HS 2 resulted in a sig-
nicant AMOC reduction similar to other Heinrich Stadials. Higher Pa/
Th ratios in North Atlantic sediment cores are traditionally interpreted
to be reective of a weakened AMOC49. However, increases in opal ux
as noted during HS 2 and HS 3 could also contribute to higher Pa/Th
ratios49. Additionally, oxygen isotopes from benthic foraminifera in
cores from the Caribbean suggest changes in oceanic circulation
during HS 2 were lower in magnitude compared to those during HS 1
and the YD49. A stronger than anticipated western gulf stream, as
potentially indicated from Caribbean benthic foraminifera, would have
helped mitigate SST cooling, leading to less severe drying in NE Mex-
ico. This idea is further supported from SSTs from the Gulf of Mexico,
the North Atlantic and the Caribbean,that show elevated temperatures
compared to other Heinrich Stadials. Therefore, thelack of response of
CB2 δ18O to HS 2 in our proxy record lends further support to the
possibility that HS 2 may have been weaker than other Heinrich
Stadials.
Finally, we suggest the decrease in δ13C and Mg/Ca ratios during
HS 2 and the LGM may be driven by decreased temperature at our
study site, inhibiting evapotranspiration (ET), and leading to a higher
local water balance (P-ET). We suggest precipitation remained rela-
tively constant as suggested by stable δ18O during this time-period.
PMIP3 models also support this interpretation, where increased local
water balance is observed in elevated soil moisture content in the
LGM, compared to both the Mid-Holocene and Pre-Industrial period,
even where models consistently disagree on both the sign and
magnitude of precipitation change (Fig. S12). Increased soil moisture
is likely driven by a reduction in ET, linked to increased cloudiness84,
reduced land temperatures85, or changes in relative humidity over
land86. While evaluating the exact mechanism of increased local
water balance is beyond the scope of this paper, our record is con-
sistent with PMIP3 data and recent modeling studies87, demonstrat-
ing high lake stands in Mexico and Central America50 may have been
driven by decreased ET.
Fig. 5 | Comparison of CB2 with otherpaleoclimate records. Comparison of CB2
δ13O(blue)andδ13C (orange) to Pa/Th ratios(burgundy; Henryet al. 111) with Florida
Strait gulf stream circulation (magenta49), Cariaco Basin reectance (purple74),
Juxtlahuaca (lightgreen22), CostaRica (blue108), Cuba(dark green46) and LakePetén-
Itzá (black50).
Article https://doi.org/10.1038/s41467-023-37700-9
Nature Communications | (2023) 14:2279 7
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Implications for climate model simulations
We have presented the rst multi-proxy speleothem record from NE
Mexico that highlights millennial and orbital scale hydroclimate
variability from 62.5 to 5.1 ka. In contrast to other speleothem records
from tropical and monsoon regions, we nd no strong evidence for a
direct insolation control on regional hydroclimate. We show instead
that glacial-interglacial variations are closely linked to Atlantic and
Pacic SST variations on orbital timescales. On millennial timescales,
we nd dry conditions in NE Mexico during the Younger Dryas and HS
1, 36. We utilize an isotope-enabled climate model to show that
Heinrich Stadial drying is dominantly driven by signicant reductions
in Atlantic SSTs and a strengthening of tropical easterlies,which drives
a strong Atlantic-Pacic SST/SLP gradient and a southward shift in
moisture convergence. Comparison to records from southern
Mesoamerica and the Caribbean suggests this mechanism leads to a
spatially broad and large magnitude precipitation change in response
to weakened AMOC during Heinrich Stadials. However, we nd evi-
dence for increased local water balance during HS 2 and the LGM,
which we suggest reects a combination of a relatively weaker HS 2 and
reduced evapotranspiration at our study site.
CB2 provides a much-needed record of precipitation that can be
useful for model validation. Our results suggest an important role for
both strengthened tropical easterlies, which feed in to the CLLJ, and
cool SSTs. The combined inuence decreases moisture transport to
the region during millennial-scale cold eventsindicating that both
dynamical and thermodynamical mechanisms are important drivers of
hydroclimate in NE Mexico. Furthermore, the disagreement about the
magnitude of drying in Mesoamerica is largely due to inter-model
spread in Atlantic-Pacic SST gradients29. Records of past climates,
such as CB2, which include key intervals such as the LGM, Heinrich
Stadials, and Pleistocene-Holocene transition, provide critical con-
straints for model simulations of the interbasin gradients and their
respective impact on hydroclimate. This work, along with Wright et al.9
and Bhattacharya and Coats29, emphasizes how a better understanding
of the trend and variability in the Atlantic-Pacic inter-basin gradient
will help generate more reliable predictions of future rainfall in
Mesoamerica and the Caribbean.
Methods
Chronology
The CB2 stalagmite was cut, polished and sampled for 33 U-Th dates at
2.5 cm intervals along its vertical growth axis using a Dremel hand drill
with a diamond dental bur. The CB2 sample has uranium concentra-
tions ranging from 18 to 63 ng/g (Table S1). Calcite powder samples
weighing 250300 mg were prepared at Massachusetts Institute of
Technology following methods similar to Edwards et al., (1987)88.
Powders were dissolved in nitric acid and spiked with a 229Th 233U
236U tracer, followed by isolation of U and Th by iron co-precipitation
and elution in columns with AG1-X8 resin. The isolated U and Th
fractions were analyzed using a Nu Plasma II-ES multi-collector
inductively coupled plasma mass spectrometer (MC-ICP-MS) equip-
ped with an Aridus 2 desolvating nebulizer, following methods
described in Burns et al.89. The corrected ages were calculated using an
initial 230Th/232Th value of 10.5 ± 5.3 ppm to correct for detrital 230Th.
The initial 230Th/232Th value wasdetermined by testing dates corrected
with different initial 230Th corrections for stratigraphic order and
assuming a ± 50% uncertainty, similar to methods by Hellstrom90,
Cheng (2000)91,andLin(1998)
92. The stratigraphically determined
value of 10.5 ppm closely matches the measured initial Th value of 9.5
ppm of a modern speleothem sample from Cueva Bonita (CB4), which
was determined by matching various U-Th dates with the rise in the
radiocarbon bomb peak (Fig. S13, Table S2). Although the initial Th
value of 10.5 ± 5.3 ppm is considerably higher than the traditional
assumed correction of 4.4 ± 2.2 ppm, previous work has demonstrated
select caves can have considerably higher initial 230Th/232Th values. For
instance, speleothems contaminated from detrital limestone, rather
than detrital shale predominantly composed of aluminosilicate clays,
have been shown to have initial 230Th/232Th values as high as
56111 ppm93.
U-Th ages range from 5230 ± 3200 to 57800 ± 3900 years before
present (where present is 1950 CE), and all 33 dates fall in stratigraphic
order within 2σuncertainty (Table S1). The 95% condence interval for
the age-depth model was constructed using 2000 Monte-Carlo simu-
lations through the age-depth modeling software COPRA33.Agemod-
els constructed with various 230Th/232Th values demonstrate age
uncertainties increase with larger initial Th values.But due to the use of
COPRA and abundance of Monte-Carlo simulations, the uncertainties
in the age model areconstrained, and do not scale proportionally with
the larger uncertainties in the U-Th dates from higher initial Th values
(Fig. S14).
Although we suggest the sample grew continuously from 5117 to
62525 years before present, there are some minor shifts in growth rate,
such as near the beginning of the Holocene between ~11 and 6 ka where
we have a gap in dating. While it is possible the slightly slower growth
rate during this period (~11 μm/yr) reects a minor growth hiatus
during this time-period, the rate is only slightly lower than the mean
growth rate for the sample (~14 μm/yr). Furthermore, we nd no
anomalous fabric changes in this part of CB2 so think it is reasonable to
assume it grew continuously during this period. Nevertheless, due to
the gap in dating between 116 ka and related uncertainty of the age
model at this time, the Holocene component of our record is solely
utilized as a point of comparison for examining glacial/interglacial
differences and we do not interpret it in detail. We also note a
decreased growth rate between 30,000 and 37,000 years before pre-
sent, where the age model exhibits the largest uncertainty, which
could potentially be indicative of a second growth hiatus. However,
there is no evidence of a change in calcite fabric or color during this
interval and 3 U-Th dates suggest the speleothem could have grown
continuously, when age uncertainties are accounted for. We note the
timing of millennial scale events discussed in this manuscript are
independently constrained by other U-Th dates and arenot dependent
upon the continuous growth of the sample during this time interval.
Precipitation δ18O and cave monitoring
The closest precipitation stations of the International Atomic Energy
Agency Global Network for Isotopes in Precipita tion (Veracruz and Sa n
Salvador) are over 600 km from our eld site and likely do not
represent local patterns at the cave. Therefore, we established a local
precipitation collection station directly at our eld site (Alta Cima). To
reduce kinetic effects due to evaporation, evaporation-limiting pre-
cipitation collectors were built and deployed during eldwork
(Fig. S2). Samples were collected after rainfall events in 1.5 ml glass
vials with conical inserts, sealed with paralm and refrigerated to limit
evaporation. In total, 48 samples were collected from June 2018 to May
2019 and analyzed for δDandδ18O using a Picarro L2130i cavity ring-
down spectrometer at Chapman University (Table S2). The long-term
standard deviation of an independent quality control sample is 0.51
δ2Hand0.11δ18O (VSMOW).
Moisture sourceanalysis at Cueva Bonita was conducted using air
parcel back-trajectory simulations using the NOAA HYSPIT model94
(Fig. S2). Air parcel trajectories were launched every 6 h at an elevation
of 1500 m between 2005 and 2018 using GDAS weekly data. Each tra-
jectory evaluated the air parcels location over the previous 72 h from
launch. In total, 3600 rain-bearing trajectories were produced for the
combined summer (JJAS) and winter (DJFM) months, only rain-bearing
trajectories were included in analysis (Fig. S2).
Stable isotope and trace element analysis
CB2 was micro-sampled for both stable isotope and trace element
analyses using a Sherline micromill at 500 μmincrementstoadepthof
Article https://doi.org/10.1038/s41467-023-37700-9
Nature Communications | (2023) 14:2279 8
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1 mm of the sample face, producing 1578 samples (Table S3). The
powder for CB2 was collected, weighed out to 4080 μg and analyzed
on a Kiel IV Carbonate Preparation Device coupled to a Thermo Sci-
entic DeltaV-IRMS at the UC Irvine Center for Isotope Tracers in Earth
Sciences (CITIES) following methods similar to McCabe-Glynn et al.95
to determine δ18Oandδ13C95.Every32samplesofunknowncomposi-
tion were analyzed with 14 standards which included a mix of NBS-18,
IAEA-CO-1, and an in-house standard. The analytical precision for δ18O
and δ13C is 0.08and 0.05, respectively. Speleothem δ18Ovalues
were ice-volume corrected using mean ocean δ18Ovalues(Fig.S15).
For trace element analysis, 2060 μg calcite powder aliquots
taken from the stable isotope samples were dissolved in 500 μLofa
double distilled 2% nitric acid solution. The samples were analyzed
using a Nu Instruments Attom High Resolution Inductively Coupled
Plasma Mass Spectrometer (HR-ICP-MS) at the CITIES laboratory. Mg/
Ca ratios were calculated from the intensity ratios using a bracketing
technique with ve standards of known concentration and an internal
standard (Ge) added to all samples to correct for instrumental drift.
Trace element analysis ofCB2 serves to complement the interpretation
of speleothem δ18Oandδ13C; therefore, a lower-resolution (multi-
decadal to centennial) analysis was conducted over the complete
record by analyzing every other sample (789 total; Table S3). The full
data set reported in the supplementary materials is unsmoothed.
Earth system model simulations
We use a water isotope tracer enabled version of the Community Earth
System Model, iCESM167. Model physics are consistent with CESM1,
which simulates present-day and historical climate change quite well96.
Here, we run a fully-coupled conguration of iCESM1 with 1.9 × 2.5°
atmosphere (CAM5) and land (CLM4), and nominal ocean (POP2)
and sea ice (CICE4). The model tracks stable water isotopes of oxygen
and hydrogen through all model components. Previous studies
demonstrate that the water isotope tracer components of iCESM1 can
accurately simulate the δ18O distribution of both present97 and past
climates98.
We congured our 21 ka simulation with period-appropriate
orbital parameters and greenhouse gas concentrations, as in the
PMIP4 protocol71, and ICE-6G ice sheets99. Initial isotopic distribution
in the ocean comes from the GISS interpolated ocean δ18Odataset
100
with globally uniform enrichment of +1 for δ18O101. All other ocean
initial conditions come from a previously performed LGM
simulation102.Werunthismodelconguration for 500 years to reach
quasi-equilibrium, with our analyses coming from the nal 50 years of
simulation. We then branch this simulation to explore the effects of
melt water ux in the North Atlantic atthe LGM. Starting from year-500
of the 21 ka simulation, we add 0.50Sv of freshwater with a δ18Oof
30into the North Atlantic (50°N70°N), sufcient to rapidly and
substantially weaken AMOC103. After 100 years, we s hutoff the fresh-
water ux and extend the simulation for another 50 years. Analyses
come from the nal 50 years of the 150-year simulation.
We also perform a moisture budget analysis on the simulation
presented in Fig. 4, following the mathematical decomposition of
Seager & Henderson104. The thermodynamic term refers to changes in
P-E due to changes in specic humidity, while the dynamic term refers
to changes due to winds. Because the analysis isperformed on monthly
mean terms, a residual term, calculated by subtracting the full
response from the dynamic and thermodynamic terms, incorporates
the inuence of higher-resolution terms (e.g. transient eddies) and
non-linear terms105.
Data availability
Speleothem stable isotope and trace element data generated in this
study are provided in the Supplementary Data 1. The radiocarbon, and
U-Th data generated in this study are provided in Supporting Infor-
mation. All data have also been deposited in the NOAA paleoclimate
database, accessible at https://www.ncei.noaa.gov/access/paleo-
search/study/37679.
Code availability
iCESM1 code is available at https://github.com/NCAR/iCESM1.2.Ver-
sion 1.15 of the COPRA depth-age modeling software utilized to build
the age model in this study is available at https://doi.org/10.5194/cp-8-
1765-2012.
References
1. Knutti, R. & Sedláček, J. Robustness and uncertainties in the new
CMIP5 climate model projections. Nat. Clim. Chang. 3,
369373 (2013).
2. Lewis,S.C.,LeGrande,A.N.,Kelley,M.&Schmidt,G.A.Water
vapour source impacts on oxygen isotope variability in tropical
precipitation during Heinrich events. Clim.Past 6,325343 (2010).
3. Liverman, D. M. Drought impacts in Mexico: climate, agriculture,
technology, and land tenure in Sonora and Puebla. Ann. Assoc.
Am. Geogr. 80,4972 (1990).
4. Johnson, R. Immigration as a response variable to climate change
from Mexico into the United States. J. Alt. Perspect. Soc. Sci. 3,
758774 (2011).
5. Tierney, J. E. et al. Past climates inform our future. Science 370,
eaay3701 (2020).
6. Coats, S. et al. Paleoclimate constraints on the spatiotemporal
character of past and future droughts. J. Clim. 33,
98839903 (2020).
7. Dinezio,P.N.&Tierney,J.E.Theeffect of sea level on glacial Indo-
Pacicclimate.Nat. Geosci. 6,485491 (2013).
8. Scussolini, P. et al. Agreement between reconstructed and mod-
eled boreal precipitation of the last interglacial. Sci. Adv. 5,
eaax7047 (2019).
9. Wright, K. T. et al. Precipitation in Northeast Mexico primarily
controlled by the relative warming of Atlantic SSTs. Geophys. Res.
Lett. 49, e2022GL098186 (2022).
10. Méndez, M. & Magaña, V. Regional aspects of prolonged
meteorological droughts over Mexico and central America. J.
Clim. 23, 11751188 (2010).
11. Wang, C. Variability of the Caribbean Low-Level Jet and its rela-
tions to climate. Clim. Dyn. 29,411422 (2007).
12. Muñoz, E., Busalacchi, A. J., Nigam, S. & Ruiz-Barradas, A. Winter
and summer structure of the Caribbean low-level jet. J. Clim. 21,
12601276 (2008).
13. Wang, C. & Lee, S. K. Atlantic warm pool, Caribbean low-level jet,
and their potential impact on Atlantic hurricanes. Geophys. Res.
Lett. 34,L02703(2007).
14. Roy, P. D. et al. Atlantic Ocean modulated hydroclimate of the
subtropical northeastern Mexico since the Last Glacial Maximum
andcomparisonwiththesouthernUS.Earth Planet. Sci. Lett. 434,
141150 (2016).
15. Roy, P. D. et al. Response of arid northeast Mexico to global cli-
mate changes during the late Pleistocene to the middle Holocene.
Earth Surf. Process. Landf. 44, 22112222 (2019).
16. Quiroz-Jiménez, J. D., Roy, P. D., Beramendi-Orosco, L. E., Lozano-
García, S. & Vázquez-Selem, L. Orbital-scale droughts in central-
northern Mexico during the late Quaternary and comparison with
other subtropical and tropical records. Geol. J. 53,
230242 (2018).
17. Metcalfe, S. E., Barron, J. A. & Davies, S. J. The Holocene history of
the North American Monsoon: known knownsand known
unknownsin understanding its spatial and temporal complexity.
Quat. Sci. Rev. 120,127 (2015).
18. Barron, J. A., Metcalfe, S. E. & Addison, J. A. Response of the North
American monsoon to regional changes in ocean surface tem-
perature. Paleoceanography 27,(2012).
Article https://doi.org/10.1038/s41467-023-37700-9
Nature Communications | (2023) 14:2279 9
Content courtesy of Springer Nature, terms of use apply. Rights reserved
19. Kirby,M.E.,Lund,S.P.&Bird,B.W.Mid-Wisconsinsediment
record from Baldwin Lake reveals hemispheric climate dynamics
(Southern CA, USA). Palaeogeogr. Palaeoclimatol. Palaeoecol. 241,
267283 (2006).
20. Roy, P. D. et al. A millennial-scale late Pleistocene-Holocene
palaeoclimatic record from the western Chihuahua Desert, Mex-
ico. Boreas 41,707718 (2012).
21. Roy, P. D. et al. A record of Holocene summer-season palaeohy-
drological changes from the southern margin of Chihuahua Desert
(Mexico) and possible forcings. Holocene 23, 11051114 (2013).
22. Lachniet, M. S., Asmerom, Y., Bernal, J. P., Polyak, V. J. & Vazquez-
Selem, L. Orbital pacing and ocean circulation-induced collapses
of the Mesoamerican monsoon over the past 22,000 y. Proc. Natl
Acad.Sci.USA110,92559260 (2013).
23. Mestas-Nuñez, A. M., Eneld, D. B. & Zhang, C. Water vapor uxes
over the Intra-Americas Sea: seasonal and interannual variability
and associations with rainfall. J. Clim. 20,19101922 (2007).
24. Cook,K.H.&Vizy,E.K.HydrodynamicsoftheCaribbeanlow-level
jet and its relationship to precipitation. J. Clim. 23,
14771494 (2010).
25. Villanueva-Diaz, J. et al. Winter-spring precipitation reconstruc-
tions from tree rings for northeast Mexico. Clim. Change 83,
117131 (2007).
26. Stahle,D.W.,Fye,F.K.,Cook,E.R.&Grifn, R. D. Tree-ring
reconstructed megadroughts over North America since a.d. 1300.
Clim. Change 83,133149 (2007).
27. Stahle, D. W. et al. The Mexican Drought Atlas: tree-ring recon-
structions of the soil moisture balance during the late pre-His-
panic, colonial, and modern eras. Quat. Sci. Rev. 149,
3460 (2016).
28. Bhattacharya, T., Chiang, J. C. H. & Cheng, W. Ocean-atmosphere
dynamics linked to 8001050 CE drying in mesoamerica. Quat.
Sci. Rev. 169,263277 (2017).
29. Bhattacharya, T. & Coats, S. Atlantic-Pacicgradientsdrivelast
Millennium hydroclimate variability in Mesoamerica. Geophys.
Res. Lett. 47, e2020GL088061 (2020).
30. Medina-Elizalde, M. et al. Synchronous precipitation reduction in
the American Tropics associated with Heinrich. Sci. Rep. 7,
112 (2017).
31. Roy, P. D. et al. Depositional histories of vegetation and rainfall
intensity in Sierra Madre Oriental Mountains (northeast Mexico)
since the late Last Glacial. Glob. Planet. Change 187,
103136 (2020).
32. Roy, P. D. et al. Last glacial hydrological variations at the southern
margin of sub-tropical North America and a regional comparison.
J. Quat. Sci. 29, 495505 (2014).
33. Breitenbach, S. F. M. et al. Constructing proxy records from age
models (COPRA). Clim 8,17651779 (2012).
34. Medina-Elizalde, M., Polanco-Martínez, J. M., Lases-Hernández, F.,
Bradley, R. & Burns, S. Testing the tropical stormhypothesis of
Yucatan Peninsula climate variability during the Maya Terminal
Classic Period. Quat. Res. 86, 111119 (2016).
35. Frappier, A. B. et al. Two millennia of tropical cyclone-induced
mud layers in a northern Yucatán stalagmite: multiple overlapping
climatic hazards during the Maya Terminal Classic mega-
droughts.Geophys. Res. Lett. 41,51485157 (2014).
36. Lachniet, M. S., Asmerom, Y., Polyak, V. & Bernal, J. P. Two mil-
lennia of Mesoamerican monsoon variability driven by Pacicand
Atlantic synergistic forcing. Quat. Sci. Rev. 155,100113 (2017).
37. Risi,C.,Bony,S.&Vimeux,F.Inuence of convective processes on
the isotopic composition (δ18OandδD) of precipitation and water
vapor in the tropics: 2. Physical interpretation of the amount
effect. J. Geophys. Res. 113,D19306(2008).
38. Maupin, C. R. et al. Abrupt Southern Great Plains thunderstorm
shifts linked to glacial climate variability. Nat. Geosci. (2021).
39. Aggarwal, P. K. et al. Proportions of convective and stratiform
precipitation revealed in water isotope ratios. Nature 9,
624629 (2016).
40. Wong,C.I.,Banner,J.L.&Musgrove,M.L.Holoceneclimate
variability in Texas, USA: An integration of existing paleoclimate
data and modeling with a new, high-resolution speleothem
record. Quat.Sci.Rev.127,155173 (2015).
41. Johnson, K. R., Hu, C., Belshaw, N. S. & Henderson, G. M. Seasonal
trace-element and stable-isotope variations in a Chinese spe-
leothem: The potential for high-resolution paleomonsoon recon-
struction. Earth Planet. Sci. Lett. 244,394407 (2006).
42. Fohlmeister, J. et al. Main controls on the stable carbon isotope
composition of speleothems. Geochim. Cosmochim. Acta 279,
6787 (2020).
43. Grifths, M. L. et al. End of Green Sahara amplied mid- to late
Holocene megadroughts in mainland Southeast Asia. Nat. Com-
mun. 11, 4204 (2020).
44. Cheng, H. et al. The Asian monsoon over the past 640,000 years
and ice age terminations. Nature 534,640646 (2016).
45. Cheng,H.,Sinha,A.,Wang,X.,Cruz,F.W.&Edwards,R.L.The
Global Paleomonsoon as seen through speleothem records from
Asia and the Americas. Clim. Dyn. 39,10451062 (2012).
46. Warken, S. F. et al. Caribbean hydroclimate and vegetation history
across the last glacial period. Quat. Sci. Rev. 218,7590 (2019).
47. Hemming,S.R.Heinrichevents:Massive late Pleistocene detritus
layers of the North Atlantic and their global climate imprint. Rev.
Geophys. 42,RG1005(2004).
48. Wong, C. I. & Breecker, D. O. Advancements in the use of spe-
leothems as climate archives. Quat. Sci. Rev. 127,118 (2015).
49. Lynch-Stieglitz, J. et al. Muted change in Atlantic overturning cir-
culation over some glacial-aged Heinrich events. Nat. Geosci. 7,
144150 (2014).
50. Escobar, J. et al. A 43-ka record of paleoenvironmental change in
the Central American lowlands inferred from stable isotopes of
lacustrine ostracods. Quat. Sci. Rev. 37,92104 (2012).
51. Stoll,H.M.,Müller,W.&Prieto,M.I-STAL,amodelforinter-
pretation of Mg/Ca, Sr/Ca and Ba/Ca variations in speleothems
and its forward and inverse application on seasonal to millennial
scales. Geochem. Geophys. Geosyst. 13,(2012).
52. Day, C. C. & Henderson, G. M. Controls on trace-element parti-
tioning in cave-analogue calcite. Geochim. Cosmochim. Acta 120,
612627 (2013).
53. Roy, P. D. et al. Paleohydrology of the Santiaguillo Basin (Mexico)
since late last glacial and climate variation in southern part of
western subtropical North America. Quat. Res. (USA) 84,
335347 (2015).
54. Lachniet, M. S. et al. Late Quaternary moisture export across
Central America and to Greenland: evidence for tropical rainfall
variability from Costa Rican stalagmites. Quat. Sci. Rev. 28,
33483360 (2009).
55. Andersen, K. K. et al. High-resolution record of Northern Hemi-
sphere climate extending into the last interglacial period. Nature
431,147151 (2004).
56. Dubois, N. et al. Millennial-scale variations in hydrography and
biogeochemistry in the Eastern Equatorial Pacicoverthelast100
kyr. Quat. Sci. Rev. 30,210223 (2011).
57. Lea, D. W., Pak, D. K., Peterson, L. C. & Hughen, K. A. Synchroneity
of tropical and high-latitude Atlantic temperatures over the last
glacial termination. Science 301,13611364 (2003).
58. Schmidt,M.W.,Spero,H.J.&Lea,D.W.Linksbetweensalinity
variation in the Caribbean and North Atlantic thermohaline circu-
lation. Nature 428,160163 (2004).
59. Ziegler,M.,Nürnberg,D.,Karas,C.,Tiedemann,R.&Lourens,L.J.
Persistent summer expansion of the Atlantic Warm Pool during
glacial abrupt cold events. Nat. Geosci. 1,601605 (2008).
Article https://doi.org/10.1038/s41467-023-37700-9
Nature Communications | (2023) 14:2279 10
Content courtesy of Springer Nature, terms of use apply. Rights reserved
60. Tripsanas, E. K. et al. Paleoenvironmental and paleoclimatic
implications of enhanced holocene discharge from the mississippi
river based on the sedimentology and geochemistry of a deep
core (JPC-26) from the gulf of Mexico. Palaios 28,623636 (2013).
61. Winter, A. et al. Initiation of a stable convective hydroclimatic
regime in Central America circa 9000 years BP. Nat. Commun. 11,
18 (2020).
62. Warken, S. F. et al. Persistent link between Caribbean precipitation
and Atlantic Ocean circulation during the last glacial revealed by a
speleothem record from Puerto Rico. Paleoceanogr. Paleoclima-
tol. 35,117 (2020).
63. Thirumalai, K. et al. Pronounced centennial-scale Atlantic Ocean
climate variability correlated with Western Hemisphere hydro-
climate. Nat. Commun. 9,392(2018).
64. Seager, R. & Hoerling, M. Atmosphere and ocean origins of North
American droughts. J. Clim. 27,45814606 (2014).
65. Seager, R. The turn of the century North American drought: global
context, dynamics, and past analogs. J. Clim. 20,
55275552 (2007).
66. Quiroz-Jimenez, J. D., Roy, P. D., Lozano-Santacruz, R. & Giron-
García, P. Hydrological responses of the Chihuahua Desert of
Mexico to possible Heinrich Stadials. J.SouthAm.EarthSci.73,
19(2017).
67. Brady,E.etal.Theconnectedisotopicwatercycleinthecom-
munity earth system model version 1. J. Adv. Model. Earth Syst. 11,
25472566 (2019).
68. Baldwin,J.W.,Atwood,A.R.,Vecchi,G.A.&Battisti,D.S.Outsize
inuence of Central American orography on global climate. AGU
Adv. 2, e2020AV000343 (2021).
69. Waelbroeck, C. et al. The timing of the last deglaciation in North
Atlantic climate records. Nature 412,724727 (2001).
70. Xie, S., Ping & Philander, S. G. H. A coupled oceanatmosphere
model of relevance to the ITCZ in the eastern Pacic. Tellus A 46,
340350 (1994).
71. Kageyama, M. et al. The PMIP4 contribution to CMIP6part 4:
Scientic objectives and experimental design of the PMIP4-CMIP6
Last Glacial Maximum experiments and PMIP4 sensitivity experi-
ments. Geosci. Model Dev. 10,40354055 (2017).
72. Thiagarajan, N., Subhas, A. V., Southon, J. R., Eiler, J. M. & Adkins, J.
F. Abrupt pre-Bølling-Allerød warming and circulation changes in
the deep ocean. Nature 511,7578 (2014).
73. Glenn, E. et al. Article tropical convection in the caribbean and
surrounding region during a regional, warming sea-surface tem-
perature period. Hydrology 8,19822020 (2021).
74. Deplazes, G. et al. Links between tropical rainfall and North
Atlantic climate during the last glacial period. Nat. Geosci. 6,
213217 (2013).
75. Them, T. R., Schmidt, M. W. & Lynch-Stieglitz, J. Millennial-scale
tropical atmospheric and Atlantic Ocean circulation change from
the Last Glacial Maximum and marine isotope stage 3. Earth Pla-
net. Sci. Lett. 427,4756 (2015).
76. Arienzo,M.M.etal.Multi-proxy evidence of millennial climate
variability from multiple Bahamian speleothems. Quat.Sci.Rev.
161,1829 (2017).
77. Singarayer, J. S., Valdes, P. J.& Roberts, W. H.G. Ocean dominated
expansion and contraction of the late Quater nary tropical rainbelt.
Sci. Rep. 7,9382(2017).
78. Bahr, A. et al. Low-latitude expressions of high-latitude forcing
during Heinrich Stadial 1 and the Younger Dryas in northern South
America. Glob. Planet. Change 160,19(2018).
79. Chevalier, M., Brewer, S. & Chase, B. M. Qualitative assessment of
PMIP3 rainfall simulations across the eastern African monsoon
domains during the mid-Holocene and the Last Glacial Maximum.
Quat. Sci. Rev. 156,107120 (2017).
80. Hodell, D. A. et al. An 85-ka record of climate change in lowland
Central America. Quat. Sci. Rev. 27, 11521165 (2008).
81. Bradbury, J. P. Sources of glacial moisture in Mesoamerica. Quat.
Int. 4344,97110 (1997).
82. Metcalfe, S. E., OHara,S.L.,Caballero,M.&Davies,S.J.Records
of Late Pleistocene-Holocene climatic change in Mexicoa
review. Quat. Sci. Rev. 19,699721 (2000).
83. Dansgaard, W. Stable isotopes in precipitation. Tellus 16,
436468 (1964).
84. Bush, M. B. et al. in Past Climate Variability in South America and
Surrounding Regions Vol 14 (eds Vimeux, F. et al.) 113128
(Springer, 2009).
85. Lachniet, M. S. & Vazquez-Selem, L. Last Glacial Maximum equi-
librium line altitudes in the circum-Caribbean (Mexico, Guate-
mala, Costa Rica, Colombia, and Venezuela). Quat. Int. 138139,
129144 (2005).
86. Bush, A. B. G. & Philander, S. G. H. The climate of the Last Glacial
Maximum: results from a coupled atmosphere-ocean general
circulation model. J. Geophys. Res. Atmos. 104,
2450924525 (1999).
87. Lowry,D.P.&Morrill,C.IstheLastGlacialMaximumareverse
analog for future hydroclimate changes in the Americas? Clim.
Dyn. 52, 44074427 (2019).
88. LawrenceEdwards,R.,Chen,J.H.&Wasserburg,G.J.
238U234U230Th232Th systematics and the precise measurement
of time over the past 500,000 years. Earth Planet. Sci. Lett. 81,
175192 (1987).
89. Burns, S. J. et al. Rapid human-induced landscape transformation
in Madagascar at the end of the rst millennium of the Common
Era. Quat. Sci. Rev. 134,9299 (2016).
90. Hellstrom,J.UTh dating of speleothems with high initial 230Th
using stratigraphical constraint. Quat. Geochronol. 1,
289295 (2006).
91. Cheng, H., Adkins, J., Edwards, R. L. & Boyle, E. A. U-Th dating of
deep-sea corals. Geochim. Cosmochim. Acta 64,
24012416 (2000).
92. Lin, J. C. et al. A reassessment of U-Th and 14C ages for late-glacial
high-frequency hydrological events at Searles Lake, California.
Quat. Res. 49,1123 (1998).
93. Carolin, S. A. et al. Varied response of western Pacic hydrology to
climate forcings over the last glacial period. Science 340,
15641566 (2013).
94. Stein,A.F.etal.Noaas hysplit atmospheric transport and dis-
persion modeling system. Bull. Am. Meteorological Soc. 96,
20592077 (2015).
95. McCabe-Glynn, S. et al. Variable North Pacicinuence on
drought in southwestern North America since AD 8 54. Nat. Geosci.
6,617621 (2013).
96. Hurrell, J. W. et al. The community earth system model: a frame-
work for collaborative research. Bull.Am.Meteorol.Soc.94,
13391360 (2013).
97. Nusbaumer, J., Wong, T. E., Bardeen, C. & Noone, D. Evaluating
hydrological processes in the Community Atmosphere Model
Version 5 (CAM5) using stable isotope ratios of water. J. Adv.
Model. Earth Syst. 9,949977 (2017).
98. He, C. et al. Hydroclimate footprint of pan-Asian monsoon
water isotope during the last deglaciation. Sci. Adv. 7,
eabe2611 (2021).
99. Peltier,W.R.,Argus,D.F.&Drummond,R.Spacegeodesycon-
strains ice age terminal deglaciation: The global ICE-6G-C (VM5a)
model. J. Geophys. Res. Solid Earth 120,450487 (2015).
100. LeGrande, A. N. & Schmidt, G. A. Global gridded data set of the
oxygen isotopic composition in seawater. Geophys. Res. Lett. 33,
L12604 (2006).
Article https://doi.org/10.1038/s41467-023-37700-9
Nature Communications | (2023) 14:2279 11
Content courtesy of Springer Nature, terms of use apply. Rights reserved
101. Duplessy, J. C., Labeyrie, L. & Waelbroeck, C. Constraints on the
ocean oxygen isotopic enrichment between the Last Glacial
Maximum and the holocene: Paleoceanographic implications.
Quat. Sci. Rev. 21,315330 (2002).
102. Di Nezio, P. N. et al. The climate response of the Indo-Pacicwarm
pool to glacial sea level. Paleoceanography 31,866894 (2016).
103. Zhu, J. et al. Investigating the direct meltwater effect in terrestrial
oxygen-isotope paleoclimate records using an isotope-enabled
earth system model. Geophys. Res. Lett. 44,12,50112,510 (2017).
104. Seager, R. & Henderson, N. Diagnostic computation of moisture
budgets in the ERA-interim reanalysis with reference to analysis of
CMIP-archived atmospheric model data. J. Clim. 26,
78767901 (2013).
105. Bhattacharya, T. & Chiang, J. C. H. Spatial variability and
mechanisms underlying El Niño-induced droughts in Mexico.
Clim. Dyn. 43, 33093326 (2014).
106. Ashouri, H. et al. PERSIANN-CDR: Daily precipitation climate data
record from multisatellite observations for hydrological and cli-
mate studies. Bull. Am. Meteorol. Soc. 96,6983 (2015).
107. Giorgetta, M. A. et al. Climate and carbon cycle changes from
1850 to 2100 in MPI-ESM simulations for the Coupled Model
Intercomparison Project phase 5. J. Adv. Model. Earth Syst. 5,
572597 (2013).
108. Lachniet, M. S. Climatic and environmental controls on spe-
leothem oxygen-isotope values. Quat. Sci. Rev. 28,
412432 (2009).
109. Bereiter, B. et al. Revision of the EPICA Dome C CO
2
record from
800 to 600-kyr before present. Geophys. Res. Lett. 42,
542549 (2015).
110. Rasmussen, S. O. et al. A new Greenland ice core chronology for
the last glacial termination. J. Geophys. Res. Atmos. 111,
D06102 (2006).
111. Henry, L. G., McManus, J. F., Curry, W. B., Roberts, N. L., Piotrowski,
A. M. & Keigwin, L. D. North Atlantic ocean circulation and abrupt
climate change during the last glaciation. Science,353,
470474 (2016).
Acknowledgements
We thank Cheva Berrones-Benitez for her assistance with rainfall sam-
pling. We thank Jim Kennedy, Esteban Berrones, Corinne Wong, and
Chris Wood for their help with eldwork. We thank Dachun Zhang, Jes-
sica Wang, Chris Wood, and Elizabeth Patterson for assistance with lab
work. We thank Crystal Tulley-Cordova for sharing the precipitation
collector design. This research was supported by a UC MEXUS-
CONACYT Collaborative Grant from the University of California Institute
for Mexico and the United States (UC MEXUS CN-16-120; K.R.J. and
L.B.O.), an MIT International Science and Technology Initiatives Mexico
Program (D.M.), National Science Foundation awards AGS-1804512 and
AGS-1806090 (K.R.J. and D.M.), and UC National Laboratory In-
Residence Graduate Fellow award LGF-19-600874 (K.T.W.).
Author contributions
K.T.W., K.R.J., D.M., and L.B-O. designed the study; K.T.W., G.S.M, K.R.J.,
and J-L.L.M. conducted eldwork; K.T.W., G.S.M., G.R.G., and G.L. con-
ducted laboratory analyses; C.R.T. conducted paleoclimate model
simulations; T.B., C.R.T., and K.T.W. analyzed model data; K.T.W. and
K.R.J. wrote the manuscript with help from coauthors. All authors con-
tributed to data analysis, interpretation and manuscript review.
Competing interests
The authors declare no competing interests.
Additional information
Supplementary information The online version contains
supplementary material available at
https://doi.org/10.1038/s41467-023-37700-9.
Correspondence and requests for materials should be addressed to
Kevin T. Wright or Kathleen R. Johnson.
Peer review information Nature Communications thanks Sylvia Dee and
the other, anonymous, reviewer(s) for their contribution to the peer
review of this work. Peer reviewer reports are available.
Reprints and permissions information is available at
http://www.nature.com/reprints
Publishers note Springer Nature remains neutral with regard to jur-
isdictional claims in published maps and institutional afliations.
Open Access This article is licensed under a Creative Commons
Attribution 4.0 International License, which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as
long as you give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons license, and indicate if
changes were made. The images or other third party material in this
article are included in the articles Creative Commons license, unless
indicated otherwise in a credit line to the material. If material is not
included in the articles Creative Commons license and your intended
use is not permitted by statutory regulation or exceeds the permitted
use, you will need to obtain permission directly from the copyright
holder. To view a copy of this license, visit http://creativecommons.org/
licenses/by/4.0/.
© The Author(s) 2023
Article https://doi.org/10.1038/s41467-023-37700-9
Nature Communications | (2023) 14:2279 12
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... Variability in zonal mean tropical rainfall originating from the Intertropical Convergence Zone (ITCZ) and the associated tropical rain belts (TRB) is thought to be influenced by inter-hemispheric temperature gradient variations and cross-equatorial energy flux perturbations (Deplazes et al., 2013;Fensterer et al., 2013;Lechleitner et al., 2017;Warken et al., 2020). For example, during the last glacial maximum (LGM) and the Heinrich stadials, colder temperatures in the North Atlantic contributed to a more southern ITCZ position over the Atlantic (Chiang et al., 2003;Wright et al., 2023). In addition, the regional heterogeneous land-sea distribution, gradients between the Pacific and the Atlantic oceans, as well as other internal and external forcing mechanisms (e.g. ...
... solar or volcanic) modulate the position and extent of the ITCZ leading to complex precipitation distributions on both spatial and temporal scales (e.g. Bhattacharya and Coats, 2020;Medina et al., 2023;Ridley et al., 2015;Singarayer et al., 2017;Warken et al., 2021;Wright et al., 2023). ...
... Further evidence of a sequence of several sub-events of the 8.2 ka event comes from the Cariaco basin as well as a Venezuelan speleothem record, where rainfall-sensitive proxy records suggest a number of rapid dry/wet fluctuations over northern South America (Figure 4f and g, Hughen et al., 1996b;Deplazes et al., 2013;Medina et al., 2023). In contrast, other Caribbean records show only one single drying event such as, for example, Venado Cave (Costa Rica, Figure 4h, (Lachniet et al., 2004)) but also records from Cuba or Central America (Fensterer et al., 2013;Hillesheim et al., 2005;Hodell et al., 1995), or none at all, as in the speleothem δ 18 O records from eastern Guatemala or north-eastern Mexico (Wright et al., 2023). Again, this might be due to different regional responses or low archive resolution. ...
Article
Full-text available
A speleothem collected from Palco Cave (Puerto Rico) spans the 8.2 ka event, a time interval associated with fluctuations of Atlantic Ocean circulation and possible drying in the Caribbean region. While stalagmite δ ¹⁸ O, δ ¹³ C, and Mg/Ca data do not show a sustained change in mean state over the 8.2 ka event, the proxies provide robust evidence for three abrupt fluctuations toward drier conditions in rapid succession, each lasting less than two decades, occurring at 8.20, 8.14, and 8.02 ka BP. A cave monitoring program at Palco Cave supports the interpretation of the speleothem proxy records. Because changes in the position of the Intertropical Convergence Zone (ITCZ) are directly coupled to sea-surface temperature variations in the North Atlantic, we hypothesize that cold events in the North Atlantic temporarily limited the northward migration of the ITCZ and tropical rain belt in boreal summer during these abrupt drying periods. The speleothem record suggests that the 8.2 ka event was associated with rapid rainfall fluctuations in the northern Caribbean followed by a comparably warm and wet phase after the 8.2 ka event. This enhanced variability during the transitional period of the deglaciation appears to be linked to a fast coupling between interacting oceanic and atmospheric processes. This involves, in particular, the Atlantic Meridional Overturning Circulation in modulating interhemispheric heat transport.
... Prior work has shown the utility of using multiple speleothem proxies to infer environmental changes through the 8.2 ka event (Allan et al., 2018;de Wet et al., 2021;Y. H. Liu et al., 2013;Oster et al., 2017;Owen et al., 2016;Waltgenbach et al., 2020), particularly in light of recent research highlighting the advantage of additional proxies such as trace elements, calcite fabric, and optical fluorescence in interpreting the δ 18 O time series Griffiths et al., 2020;Johnson, 2021aJohnson, , 2021bPatterson et al., 2023;Vanghi et al., 2018Vanghi et al., , 2019Wright et al., 2023;Zhang et al., 2018). ...
Article
Full-text available
The 8.2 ka event is the most significant global climate anomaly of the Holocene epoch, but a lack of records from Mainland Southeast Asia (MSEA) currently limits our understanding of the spatial and temporal extent of the climate response. A newly developed speleothem record from Tham Doun Mai Cave, Northern Laos provides the first high‐resolution record of this event in MSEA. Our multiproxy record (δ¹⁸O, δ¹³C, Mg/Ca, Sr/Ca, and petrographic data), anchored in time by 9 U‐Th ages, reveals a significant reduction in local rainfall amount and weakening of the monsoon at the event onset at ∼8.29 ± 0.03 ka BP. This response lasts for a minimum of ∼170 years, similar to event length estimates from other speleothem δ¹⁸O monsoon records. Interestingly, however, our δ¹³C and Mg/Ca data, proxies for local hydrology, show that abrupt changes to local rainfall amounts began decades earlier (∼70 years) than registered in the δ¹⁸O. Moreover, the δ¹³C and Mg/Ca also show that reductions in rainfall continued for at least ∼200 years longer than the weakening of the monsoon inferred from the δ¹⁸O. Our interpretations suggest that drier conditions brought on by the 8.2 ka event in MSEA were felt beyond the temporal boundaries defined by δ¹⁸O‐inferred monsoon intensity, and an initial wet period (or precursor event) may have preceded the local drying. Most existing Asian Monsoon proxy records of the 8.2 ka event may lack the resolution and/or multiproxy information necessary to establish local and regional hydrological sensitivity to abrupt climate change.
Article
Full-text available
Plain Language Summary We use geochemical markers of past rainfall in a rock from a cave (speleothem) to show that warming in Atlantic sea‐surface temperatures (SSTs) increases the amount of precipitation in Northeast Mexico. These finding are surprising, since previous rainfall reconstructions using tree rings have suggested a warmer Atlantic decreases precipitation in the region. We used a climate model to show that warming in the Atlantic increases precipitation during the summer but decreases precipitation during the winter. Although winter precipitation only accounts for 8% of annual rainfall in this region, tree rings are more reflective of the winter precipitation response. This is the first speleothem record from NE Mexico over this time‐period and suggests that projections of future rainfall should emphasize relative changes in Atlantic SST variability because it has a major impact on annual precipitation.
Article
Full-text available
Global Climate Models (GCMs) exhibit substantial biases in their simulation of tropical climate. One particularly problematic bias exists in GCMs' simulation of the tropical rainband known as the Intertropical Convergence Zone (ITCZ). Much of the precipitation on Earth falls within the ITCZ, which plays a key role in setting Earth's temperature by affecting global energy transports, and partially dictates dynamics of the largest interannual mode of climate variability: The El Niño‐Southern Oscillation (ENSO). Most GCMs fail to simulate the mean state of the ITCZ correctly, often exhibiting a “double ITCZ bias,” with rainbands both north and south rather than just north of the equator. These tropical mean state biases limit confidence in climate models' simulation of projected future and paleoclimate states, and reduce the utility of these models for understanding present climate dynamics. Adjusting GCM parameterizations of cloud processes and atmospheric convection can reduce tropical biases, as can artificially correcting sea surface temperatures through modifications to air‐sea fluxes (i.e., “flux adjustment”). Here, we argue that a significant portion of these rainfall and circulation biases are rooted in orographic height being biased low due to assumptions made in fitting observed orography onto GCM grids. We demonstrate that making different, and physically defensible, assumptions that raise the orographic height significantly improves model simulation of climatological features such as the ITCZ and North American rainfall as well as the simulation of ENSO. These findings suggest a simple, physically based, and computationally inexpensive method that can improve climate models and projections of future climate.
Article
Full-text available
Thunderstorms in the Southern Great Plains of the United States are among the strongest on Earth and have been shown to be increasing in intensity and frequency during recent years. Assessing changes in storm characteristics under different climate scenarios, however, remains highly uncertain due to limitations in climate model physics. We analyse oxygen isotopes from Texas stalactites from 30–50 thousand years ago to assess past changes in thunderstorm size and duration using a modern radar-based calibration for the region. Storm regimes shift from weakly to strongly organized on millennial timescales and are coincident with well-known abrupt climate shifts during the last glacial period. Modern-day synoptic analysis suggests that thunderstorm organization in the Southern Great Plains is strongly coupled to changes in large-scale wind and moisture patterns. These changes in the large-scale circulation may be used to assess future predictions and palaeo-simulations of mid-latitude thunderstorm climatologies.
Article
Full-text available
Warming sea-surface temperatures (SSTs) have implications for the climate-sensitive Caribbean region, including potential impacts on precipitation. SSTs have been shown to influence deep convection and rainfall, thus understanding the impacts of warming SSTs is important for predicting regional hydrometeorological conditions. This study investigates the long-term annual and seasonal trends in convection using the Galvez-Davison Index (GDI) for tropical convection from 1982–2020. The GDI is used to describe the type and potential for precipitation events characterized by sub-indices that represent heat and moisture availability, cool/warm mid-levels at 500 hPa, and subsidence inversion, which drive the regional Late, Early, and Dry Rainfall Seasons, respectively. Results show that regional SSTs are warming annually and per season, while regionally averaged GDI values are decreasing annually and for the Dry Season. Spatial analyses show the GDI demonstrates higher, statistically significant correlations with precipitation across the region than with sea-surface temperatures, annually and per season. Moreover, the GDI climatology results show that regional convection exhibits a bimodal pattern resembling the characteristic bimodal precipitation pattern experienced in many parts of the Caribbean and surrounding region. However, the drivers of these conditions need further investigation as SSTs continue to rise while the region experiences a drying trend.
Article
Full-text available
Oxygen isotope speleothem records exhibit coherent variability over the pan-Asian summer monsoon (AM) region. The hydroclimatic representation of these oxygen isotope records for the AM, however, has remained poorly understood. Here, combining an isotope-enabled Earth system model in transient experiments with proxy records, we show that the widespread AM  18 O c signal during the last deglaciation (20 to 11 thousand years ago) is accompanied by a continental-scale, coherent hydroclimate footprint, with spatially opposite signs in rainfall. This footprint is generated as a dynamically coherent response of the AM system primarily to meltwater forcing and secondarily to insolation forcing and is further reinforced by atmospheric teleconnection. Hence, widespread  18 O p depletion in the AM region is accompanied by a northward migration of the westerly jet and enhanced southwesterly monsoon wind, as well as increased rainfall from South Asia (India) to northern China but decreased rainfall in southeast China.
Article
Full-text available
The sensitivity of tropical Atlantic precipitation patterns to the mean position of the Intertropical Convergence Zone (ITCZ) at different time scales is well‐known. However, recent research suggests a more complex behavior of the northern hemispheric tropical rain belt related to the ITCZ in the western tropical Atlantic. Here we present a precisely dated speleothem multi‐proxy record from a well‐monitored cave in Puerto Rico, covering the period between 46.2 and 15.3 ka. The stable isotope and trace element records document a pronounced response of regional rainfall to abrupt climatic excursions in the North Atlantic across the Last Glacial such as Heinrich stadials and Dansgaard/Oeschger events. The annual to multidecadal resolution of the proxy time series allows substructural investigations of the recorded events. Spectral analysis suggests that multidecadal to centennial variability persisted in the regional hydroclimate mainly during interstadial conditions but also during the Last Glacial Maximum. In particular, we observe a strong agreement between the speleothem proxy data and the strength of the Atlantic meridional overturning circulation, supporting a persistent link of oceanic forcing to regional precipitation. Comparison to other paleo‐precipitation records enables the reconstruction of past changes in position, strength, and extent of the ITCZ in the western tropical Atlantic in response to millennial‐ and orbital‐scale global climate change.
Article
Full-text available
Between 5 and 4 thousand years ago, crippling megadroughts led to the disruption of ancient civilizations across parts of Africa and Asia, yet the extent of these climate extremes in mainland Southeast Asia (MSEA) has never been defined. This is despite archeological evidence showing a shift in human settlement patterns across the region during this period. We report evidence from stalagmite climate records indicating a major decrease of monsoon rainfall in MSEA during the mid- to late Holocene, coincident with African monsoon failure during the end of the Green Sahara. Through a set of modeling experiments, we show that reduced vegetation and increased dust loads during the Green Sahara termination shifted the Walker circulation eastward and cooled the Indian Ocean, causing a reduction in monsoon rainfall in MSEA. Our results indicate that vegetation-dust climate feedbacks from Sahara drying may have been the catalyst for societal shifts in MSEA via ocean-atmospheric teleconnections. The mid-Holocene has seen a number of climate shifts, which have been associated with societal changes. Here, the authors investigate in a centuries long megadrought in Southeast Asia during the mid-Holocene, possibly caused by the end of the Green Sahara period.
Article
Full-text available
The paleoclimatic record from Mexico and Central America, or Mesoamerica, documents dramatic swings in hydroclimate over the past few millennia. However, the dynamics underlying these past changes remain obscure. We use proxy indicators of hydroclimate to show that last millennium hydroclimate variability was dominated by opposite‐signed moisture anomalies in northern and southern Mesoamerica. This pattern results from changes in moisture convergence driven by Atlantic‐Pacific interbasin temperature gradients. While this pattern is reproduced by several models and multiple experiments with a single model, models appear to disagree about the underlying dynamics of this interbasin gradient. Moreover, disagreement about the interbasin gradient, and associated hydroclimate pattern, dominates spread in 21st century regional hydroclimate projections. These results emphasize the role of interbasin asymmetries in governing past and future regional climate change. They also demonstrate that paleoclimate studies can elucidate mechanisms directly relevant to projecting future hydroclimate in climate change hot spots like Mesoamerica.
Article
Machine-learning-based methods that identify drought in three-dimensional space–time are applied to climate model simulations and tree-ring-based reconstructions of hydroclimate over the Northern Hemisphere extratropics for the past 1000 years, as well as twenty-first-century projections. Analyzing reconstructed and simulated drought in this context provides a paleoclimate constraint on the spatiotemporal characteristics of simulated droughts. Climate models project that there will be large increases in the persistence and severity of droughts over the coming century, but with little change in their spatial extent. Nevertheless, climate models exhibit biases in the spatiotemporal characteristics of persistent and severe droughts over parts of the Northern Hemisphere. We use the paleoclimate record and results from a linear inverse modeling-based framework to conclude that climate models underestimate the range of potential future hydroclimate states. Complicating this picture, however, are divergent changes in the characteristics of persistent and severe droughts when quantified using different hydroclimate metrics. Collectively our results imply that these divergent responses and the aforementioned biases must be better understood if we are to increase confidence in future hydroclimate projections. Importantly, the novel framework presented herein can be applied to other climate features to robustly describe their spatiotemporal characteristics and provide constraints on future changes to those characteristics.
Article
The future in the past A major cause of uncertainties in climate projections is our imprecise knowledge of how much warming should occur as a result of a given increase in the amount of carbon dioxide in the atmosphere. Paleoclimate records have the potential to help us sharpen that understanding because they record such a wide variety of environmental conditions. Tierney et al. review the recent advances in data collection, statistics, and modeling that might help us better understand how rising levels of atmospheric carbon dioxide will affect future climate. Science , this issue p. eaay3701