Skip to main content
Science Advances logoLink to Science Advances
. 2025 Nov 28;11(48):eaea5042. doi: 10.1126/sciadv.aea5042

North Atlantic Subtropical High forcing of Atlantic Warm Pool hydroclimate variability on millennial to orbital timescales

Hanying Li 1, Ashish Sinha 1,2, Amos Winter 3,*, Pengzhen Duan 4, Sophie Warken 5, Jun Hu 6, Shihao Lei 1, Xiyu Dong 1, Lijuan Sha 1, Haiwei Zhang 1, Gayatri Kathayat 7, Liang Yi 8, Youfeng Ning 1, Hai Cheng 1,7,9,*
PMCID: PMC12662202  PMID: 41313776

Abstract

Orbital-scale variations in insolation are widely considered to drive tropical and monsoonal rainfall, with higher summer insolation linked to stronger precipitation. Here, we present a precisely dated speleothem record from Cuba that reconstructs Atlantic Warm Pool (AWP) hydroclimate over the past 129,000 years and challenges this paradigm. Instead, we identify a previously unrecognized link between AWP hydroclimate and the North Atlantic Subtropical High (NASH) operating on millennial to orbital timescales. During intervals of high summer insolation coupled with cooler tropical North Atlantic sea surface temperatures (SSTs), NASH strengthens and expands westward, reducing rainfall across the AWP. This SST-NASH coupling amplified precessional-scale hydroclimate variability between 130 and 60 ka, when insolation amplitude was nearly twice that of the 60- to 12-ka interval. Our data further show that particularly strong insolation peaks at 105 and 126 ka caused pronounced westward NASH expansion, triggering two extreme dry events, similar to the process observed during modern midsummer dry spells.


Summer insolation shapes Caribbean hydrology via North Atlantic Subtropical High, challenging the direct-precipitation paradigm.

INTRODUCTION

Understanding the dynamical mechanisms driving hydroclimate variability in the Caribbean region, Central America, and Western Tropical North Atlantic, which constitute the Atlantic Warm Pool (AWP) (~15°N to 24°N and 85°W to 70°W), is of vital interest. The region’s climate interacts with atmospheric and oceanic circulation systems that shape local weather patterns, regulate hurricane activity, and modulate oceanic heat transport associated with the Atlantic Meridional Overturning Circulation (AMOC) (13). Although not a classical monsoon region per se, the AWP hydroclimate is heavily influenced by tropical convection associated with the Mesoamerican monsoon and the seasonal movement of the intertropical convergence zone (ITCZ) (4, 5). Notably, the North Atlantic Subtropical High (NASH), a permanent anticyclone over the North Atlantic, plays a dominant role in driving the region’s interannual precipitation variability, including the midsummer dry spells (69).

The centennial- to millennial-scale drying trends in the AWP have been previously linked to the weakening of the AMOC (5, 1015), southward shifts in the mean location of the ITCZ, reduced Mesoamerican monsoon intensity (4, 5, 10), and cooler tropical sea surface temperatures (SSTs) below the convective threshold (4, 14, 16). On orbital timescale, the speleothem δ18O records covering the past 100 kyr from the greater Mesoamerican region, such as Guatemala (17), Costa Rica (18), Colombia (13), and Cuba (4), demonstrate that when boreal summer insolation was high, the tropical-subtropical convective region became drier, challenging the conventional paradigm that monsoon and tropical convection intensity are in step with summer insolation, as observed in East Asia (19, 20), South America (21) and southwestern Mexico (10).

In this study, we present stable oxygen isotope (δ18O) records of a speleothem (CUST-1) collected from the Santo Tomás cave system (22.52°N and 83.85°W; ~170 meters above sea level) in western Cuba. The δ18O record spans the Holocene (0 to 12 ka BP, where BP is defined as years before 1950 CE) and an earlier interval from 66 to 129 ka BP. By combining our record with a previously published speleothem δ18O record from the same cave system (4) that spans from 81 to 12.7 ka BP, we are able to characterize the AWP hydroclimate variability over the past 129 kyr. Our data, combined with results from a model simulation, highlight the influence of the NASH and its interactions with other key climate drivers, such as the AMOC, tropical SSTs, and the ITCZ, in shaping the AWP hydroclimate variability across millennial to orbital timescales.

RESULTS

Precipitation-NASH-SST teleconnections

The western part of Cuba (~12°N to 26°N and 85°W to 75°W), the largest island within the Greater Antilles, is located in the core region of the AWP and positioned at the nexus of the Gulf of Mexico and the Caribbean Sea. Western Cuba has a mean annual temperature of 25.5°C and a mean annual evaporation of ~1730 mm, nearly twice the mean annual precipitation (~930 mm) between 1979 and 2020 CE, as the island is persistently under the influence of relatively dry easterly and southeasterly winds (Fig. 1A and figs. S1A and S2). Approximately 68% of annual rainfall occurs during boreal summer (May through October) when northward migration of the ITCZ and warming of the tropical North Atlantic enhance convective rainfall (figs. S1A and S3A). During this period, NASH (often referred to as the Bermuda High) strengthens and expands westward, with its core reaching ~34°N and its western boundary extending across the subtropical Atlantic into the southeastern United States (Fig. 1C and figs. S1A and S2). In contrast, during boreal winter, the system weakens and shifts southward (as the Azores High) with its core generally south of 30°N and reduced westward extent (figs. S1A and S2) (see Supplementary Text). A prominent feature of Caribbean summer rainfall is the midsummer dry spell (Fig. 1B), which coincides with the westward expansion and peak intensity of NASH. This expansion strengthens the Caribbean low-level jet and enhances subsidence over the region, leading to a reduction of up to 23% in wet-season rainfall (Fig. 1B and fig. S1A and S2) (6, 8). Notably, rainfall and SSTs over the ITCZ core region remain relatively stable during this period (fig. S1A).

Fig. 1. Climatology of the AWP and NASH.

Fig. 1.

(A) The spatial pattern of JJA mean precipitation amount (millimeters per day) (shaded), mean SST (°C) (contour dash line), and mean location of NASH indicated by 1590 and 1570–geopotential meter (gpm) isobar of geopotential height at 850 hPa (brown shaded) between 1979 and 2020. (B) Annual cycle of AWP precipitation (blue) [15°N to 24°N and 85°W to 70°W; purple box in (A)] and highest geopotential height at 850 hPa of NASH (green). The midsummer dry spell in JJA in the AWP is highlighted by a vertical bar. (C) Spatial correlation between the western boundary of the NASH (Wb-NASH) and JJA precipitation between 1979 and 2020. Stippling indicates a significant correlation (P < 0.1). Sorted by the longitude of Wb-NASH, the brown (green) line and dot are the mean of 10-year extreme westward (eastward) positions of the NASH and its western boundary, respectively. The black line and dot are the mean position of the NASH and its western boundary from 1979 to 2020. Locations of Santo Tomás cave in west Cuba (star) and other proxy records for comparison (yellow circles) are marked in [(A) and (C)]. Prcp., precipitation.

In the modern climate, MJJASO precipitation in western Cuba is significantly correlated with rainfall across much of the AWP, including the eastern Gulf of Mexico, the southeast United States, and the Caribbean Sea (fig. S4A). In contrast, only weak negative correlations are observed between western Cuba and the core ITCZ-monsoon regions, including Central America, northern South America, and the northeastern tropical Pacific (fig. S4A). This suggests that while the ITCZ contributes moisture and convective activity to the region, its direct influence on AWP rainfall is limited. In addition, the first mode of empirical orthogonal function (EOF) analysis of North Atlantic SST anomalies (STTAs) explains 32% of the variance and reveals a tripole-like pattern with positive anomalies in the tropical and high-latitude Atlantic and negative anomalies in the mid-latitudes (fig. S3B). This tripole pattern resembles the spatial correlation pattern between AWP rainfall and STTAs (fig. S3C). Instrumental and reanalysis data also link year-to-year variability in AWP rainfall to zonal shifts in the western boundary of the NASH (Wb-NASH), defined by the isopiestic contour of 850-hPa geopotential height. During the 1979–2020 CE period, negative precipitation anomalies across the AWP consistently coincided with westward shifts in Wb-NASH position (Fig. 1C and fig. S4, C and H).

Notably, the spatial pattern associated with westward shifts in the Wb-NASH closely mirrors the tripole-like SST mode identified in EOF analysis (Fig. 1C and figs. S1C and S4, C and E), indicating a strong coupling between tropical Atlantic SSTs and NASH position. A westward displacement of the Wb-NASH is consistently linked to cooling of the tropical Atlantic, and vice versa [(22) and Supplementary Text]. One prominent example of this phenomenon occurred between 2013 and 2020, when the Wb-NASH shifted abruptly to ~75°W to 7°W of its mean position, coinciding with anomalously cool tropical SSTs (fig. S4H). This shift has been linked to the persistent “pan-Caribbean drought” during this period that occurred across the Caribbean, Central America, and northern South America (23, 24). In addition, modern observations reveal a precipitation dipole between the AWP and eastern North America (Fig. 1C and figs. S1C and S4, C and E) (9). A stronger, westward-expanded NASH intensifies the Caribbean trade winds, enhancing evaporation and cooling SSTs, which further suppresses convection over the AWP (25). At the same time, northerly moisture transport into eastern North America increases, promoting wetter conditions in that region (fig. S1C).

Oxygen isotopes in precipitation

The monthly observations and simulated data from an isotope-enabled climate model (26) show that the annual cycle of δ18Op from western Cuba mirrors the precipitation cycle, with heavier values during winter and lighter values during summer (fig. S1, A, E, and F). Nevertheless, δ18Op positive excursions up to ~2 per mil (‰) can occur during the midsummer dry spell, approaching the values that are typical of wintertime (fig. S1, E and F). On interannual scales, the amount-weighted MJJASO-simulated δ18Op displays significant negative correlations with AWP precipitation (fig. S4, B, G, and H) and regional atmospheric circulation patterns (figs. S1A and S4B). Warm STTAs in the tropical and high-latitude Atlantic, coupled with cool STTAs in the mid-latitude and eastward shifting Wb-NASH, lead to convective activity over the AWP and subsequently to negative δ18Op values and vice versa (figs. S3D and S4, D, F, and H). Moreover, continental and altitude impacts on δ18Op in western Cuba are minor, given that the location is remote from large landmasses and low topographic settings (4, 27).

Speleothem δ18O records and interpretation

We generated a new speleothem δ18O record (CUST-1) from the Santo Tomás cave system (Fig. 1A) and combined it with a previously published δ18O record (CM) (4) from the same cave (Fig. 2A; see Materials and Methods, Supplementary Text, and dataset S1). Stalagmite CUST-1 (337 mm with two hiatuses detected at 89 and 94.4 mm) grew during the Holocene [~88 to 9102 and 11196 to 11860 years before the present (yr B.P.)] and Marine Isotope Stage 5 (MIS5) to MIS4 (~129,300 to 66,600 yr B.P.), which includes the Eemian interglacial (130 and 116 ka BP; Fig. 2A and fig. S5). The chronology for the three sections is constrained by 22, 4, and 53 230Th/U dates, respectively. The age versus depth relation is established using the age model algorithm StalAge (28). The large age uncertainty in the section between 186 and 193 mm is likely due to high detrital 232Th concentration, making it difficult to determine whether this interval represents a slow growth rate or an actual hiatus (see Materials and Methods, Supplementary Text, figs. S5 and S6, and dataset S2). We therefore plotted the δ18O time series between 186.5 and 193 mm (102.6 to 110.8 ka BP), with a dashed line indicating this uncertainty. The average temporal resolutions of the CUST-1 δ18O record are ~255 years for the Holocene epoch and ~33 years for the period from 66 to 129 ka BP. The CUST-1 records between 66 and 129 ka BP were further interpolated to 33-year time steps to facilitate comparisons with other proxy records. The composite record over the past 130 kyr was constructed by splicing together δ18O records spanning ~0.9 to 11.9, 12.7 to 81.4, and 81.5 to 129.3 ka BP from CUST-1, CM, and CUST-1, respectively (Supplementary Text, Fig. 2A, and fig. S7). The close alignment in δ18O values between CUST-1 and other speleothem records from Cuba on centennial to orbital scales suggests that speleothem CUST-1 δ18O primarily reflects climate signals (Supplementary Text and fig. S6, A and C) (4, 27, 29). Combined with the modern δ18Op dynamics mentioned above, we interpret the CUST-1 δ18O record to indicate regional rainfall variability with heavier (lighter) δ18O corresponding to less (more) rainfall over the AWP (fig. S8B).

Fig. 2. Orbital scale variability.

Fig. 2.

(A) The stacked Cuba speleothem δ18O records over the past 130 ka with CM (blue) (4) during glacial and CUST-1 (purple; this study) during interglacial. The summer insolation curves at 30°N, 35°N, and 40°N (34) in blue dashed lines are shown with higher value in the bottom. The mean of the entire stacked record (−1.65‰) is marked in a horizontal dashed line. (B) The same as (A) but for speleothem δ18O records from Guatemala (17). (C) The same as (A) but for speleothem δ18O records from Colombia (13) and Costa Rica (18). (D) Comparison of speleothem δ18O records from Cuba [same as (A)] and Buckeye Creek caves in North America (36). (E) The 3000-year low-pass–filtered Pa/Th records (37) with values lower than the mean in blue shaded area. (F) Tropical SST record from MD03-2616 (38) in the western North Atlantic. The values higher than 27°C are shaded in 3000-year low-pass–filtered SST record. The vertical light and dark orange bars cover the periods that summer insolation changes from the maximum to minimum. The vertical light and dark blue bars cover the periods that summer insolation changes from the minimum to maximum with lighter color for periods between 12 and 60 ka BP and darker color for the rest periods. The climate meaning of each proxy is labeled in arrows aside. The dashed line of the CUST-1 δ18O record between 103 and 111 ka BP in [(A) and (D)] indicates large age uncertainty. inso, insolation.

DISCUSSION

Orbital-scale hydroclimate changes in the AWP

We used EC-Earth-2-2 (30), an idealized Earth system model simulation, to investigate the differences between precession-scale precipitation-NASH relationships between high-insolation (precession minimum) and low-insolation (precession maximum) states (see Materials and Methods). EC-Earth is a fully coupled ocean-atmosphere general model that has been shown to reproduce climatology, interannual, spatial, and temporal variability in present-day climate with good skill. In tropical regions, the atmospheric circulation and the mean sea level pressure are well simulated (31, 32). We use the pmin and pmax simulations of EC-Earth-2-2 as performed and reported in (30).

In the model, the annual cycle of AWP precipitation shows a substantial reduction during precession minima, when the NASH intensifies markedly in JJA (Fig. 3, A and B). During precession maxima, the NASH is relatively weak, and its western boundary remains near ~57°W, close to the center of the North Atlantic (Fig. 3, C and D). In this configuration, the AWP lies largely outside the direct influence of NASH, and the annual cycle features a single broad summer peak with no substantial midsummer dry spell (Fig. 3B). In contrast, during precession minima, there is a sharp increase in NASH intensity, exceeding 25 geopotential meter (gpm), accompanied by a 17° westward shift in Wb-NASH to ~74°W (Fig. 3, A, C, and D), primarily in June and July. This expansion strongly suppresses rainfall over the AWP, with summer precipitation falling by 1.6 mm/day, even lower than winter precipitation amounts (Fig. 3B).

Fig. 3. Model simulation of difference between precession minimum and maximum in NASH and AWP precipitation.

Fig. 3.

(A) Annual cycle of highest geopotential height at 850 hPa of NASH in precession pmin (brown) and pmax (green). (B) Annual cycle of AWP precipitation in precession pmin (brown) and pmax (green). Yellow vertical bars mark the midsummer dry spell during JJA. (C) Spatial pattern of differences of JJA geopotential height and wind at 850 hPa between precession pmin and pmax. Locations of NASH indicated by 1560-gpm isobar of geopotential height are marked by brown and green lines for pmin and pmax, respectively. (D) The same as (C) but for precipitation difference. The location of Santo Tomás cave is depicted by a star. The speleothem records from Guatemala (17), Colombia (13), Costa Rica (18), and Buckeye Creek caves in North America (36) are marked in yellow circles. (E) A Hovmöller diagram of JJA vertical velocity difference between pmin and pmax averaged over a subtropical (30°N to 40°N) zone. (F) the same as (E) but for the tropical (0° to 10°N) zone.

The spatial precipitation response also differs markedly between pmin and pmax states. Precession minima are marked by a dipole pattern, with drying over the AWP and western tropical Atlantic, and enhanced rainfall over eastern North America (Fig. 3D), consistent with the westward-expanded NASH, displacing convection equatorward in the tropics and drawing moisture northward. A similar dipole pattern in precipitation difference also stood out in ensemble mean of simulations between lig127k (pmin) and piControl (pmax) (33). Notably, the ITCZ over the tropical Pacific and along the Mesoamerican monsoon domain shows enhanced convection during JJA. Together with an intensified and westward-expanded NASH, a stronger tropical and subtropical circulation during precessional minima appears to be the dominant mode of the AWP hydroclimate, especially in the zonal direction (Fig. 3, E and F). The EC-Earth-2-2 simulation results suggest that the NASH acts as a nontrivial counterpart to monsoons, as it progressively intensifies and, at the same time, extends westward, which has the effect of diminishing the impact of the Mesoamerican monsoon and ITCZ (Fig. 3, C and D).

The speleothem records from the AWP support model simulation results. Our composite Cuba speleothem record exhibits a consistent pattern throughout ~5.5 precession cycles (~19 to 23 kyr) over the past 130 kyr. On the precessional scale, the dry (wet) periods indicated by the positive (negative) δ18O values are in sync with high (low) boreal summer insolation at 30°N, 35°N, and 40°N (34)—the latitudinal range over which summer insolation exerts influence on NASH variability (Fig. 2A and fig. S7). Similar relationships over the past 100 kyr hold for three other speleothem δ18O records from Bombil Pek and Jul Iq caves in Guatemala (17), Terciopelo Cave in Costa Rica (18), and Caracos Cave in Colombia (Fig. 2) (13). While these records are located in the Mesoamerican monsoon and ITCZ core region, respectively (13, 17, 18), they, together with speleothem records from Cuba, indicate drier conditions when insolation was high, manifesting a strong relation to a large regional extent equivalent to the pan-Caribbean drought region in present-day climate (23). Notably, the land in the Mesoamerican monsoon region, essentially represented by the Central American isthmus, is narrow, further restraining the affected area of the Mesoamerican monsoon/ITCZ (35) but expanding the affected area of NASH. In addition, a comparison of speleothem δ18O records between Cuba and Buckeye Creek Cave in eastern North America (36) shows a dipole pattern of dry (wet) AWP and wet (dry) eastern North America on precession minimum (precession maxima) (Fig. 2D), indicating the NASH impacts on precession scale. The spatial hydroclimate pattern from these records is broadly consistent with model simulations (Fig. 3D). Considering the proxy data-model agreements, we posit that intensified NASH and westward-expanded Wb-NASH driven by high insolation played a dominant role in regulating rainfall variability in AWP and eastern North America on precession scales and vice versa (Fig. 2).

The coupled insolation-SST impact on NASH and AWP rainfall variability

Speleothem δ18O records from Cuba and Guatemala (17) indicate that wetter conditions in the AWP occurred primarily during MIS 5, the early part of MIS 4, and the Holocene—periods broadly coinciding with intensified AMOC (37) and sustained tropical Atlantic SSTs (38) above the convective threshold (~27°C) (Fig. 2). Before ~60 ka BP, the AMOC transported warm surface waters into the high-latitude North Atlantic, maintaining a robust upper ocean limb that helped sustain both high tropical SSTs and atmospheric moisture supply to the AWP, supporting stronger convective activity and higher rainfall (39). In contrast, the transition to drier conditions after 60 ka BP coincided with increased iceberg discharge (40, 41), weakening of the AMOC (37), and a marked reduction in tropical Atlantic SSTs (38), indicating a shift to a glacial hydroclimate regime (17). The combination of reduced moisture transport and subconvective SSTs effectively suppressed rainfall in the AWP from ~65 to 12 ka BP. These observations underscore the critical role of tropical SSTs and ocean circulation in modulating hydroclimate in the region.

Over the past 130 kyr, our data reveal a persistent interaction between tropical Atlantic SSTs and the position and intensity of NASH that modulate AWP rainfall at precession timescales. Specifically, intervals of decreasing summer insolation at 128 to 116, 104 to 94, 83 to 72, 59 to 46, 34 to 22, and 12 to 0 ka BP coincide with tropical Atlantic warming, NASH weakening, and an eastward retreat of its western boundary (Fig. 2), producing negative excursions in the Cuban speleothem δ18O, which reflect a wetter AWP (Fig. 2, A and F). In contrast, intervals of increasing summer insolation at 116 to 104, 94 to 83, 72 to 59, 46 to 34, and 22 to 12 ka BP are linked to tropical SST cooling and enhanced and westward-expanded NASH, consistent with heavier cave δ18O, indicating drier conditions (Fig. 2). The land-sea thermal contrast is the most predominant driver of NASH intensity (42). During high summer insolation, the continental warming outpaces that of the ocean, enhancing this land-sea thermal contrast and strengthening NASH, thereby suppressing convection over the AWP. This effect is amplified when the tropical North Atlantic SSTs are cooler, whereas the opposite likely occurs, where configuration arises under low summer solar radiation and warmer tropical SSTs.

Crucially, the magnitude of precession-driven hydroclimate variability in the AWP was not uniform over the last 130 kyr. Our δ18O record exhibits a distinct regime shift around 60 ka BP, marking a transition from a period of pronounced precessional variability to one with markedly damped fluctuations (Fig. 2A). This shift is accompanied by reductions in both summer insolation amplitude and the magnitude of tropical Atlantic SST cooling (Fig. 2F). Between 129 and 60 ka BP, summer insolation varied by more than 30 W/m2, tropical SSTs cooled by over 1.2°C, and δ18O amplitudes averaged ~0.87‰. In contrast, between 60 and 12 ka BP, insolation amplitude declined to less than 15 W/m2, SST cooling fell below 1°C, and δ18O excursions diminished to ~0.44‰ (Fig. 2, A and F). These observations suggest that precessional variability of NASH is controlled jointly by insolation amplitude and tropical SSTs. When both insolation and SST variability are high, the resulting thermal gradient strengthens NASH and extends the westward reach of its western boundary, thereby amplifying AWP rainfall variability. When both are weak, NASH is reduced, and its longitudinal range contracts, weakening rainfall variability. On glacial-interglacial timescales, Cuban speleothem δ18O records remain coherent with boreal summer insolation (fig. S7D), but the amplitude of this signal is markedly damped between 60 and 12 ka BP compared with 129 to 60 ka BP. This reduced sensitivity to insolation indicates that during glacial periods, AWP hydroclimate was less directly driven by orbital forcing and more strongly influenced by AMOC and SST changes (43).

Millennial events between 66 and 129 ka BP

Between 129 and 66 ka BP, the CUST-1 δ18O profiles exhibit prominent positive excursions (Fig. 4F). To emphasize the variability on multicentennial to millennial timescales, we removed long-term (>10 kyr) nonlinear trends from the CUST-1 δ18O record (see Materials and Methods, Fig. 4G, and fig. S8). The detrended record reveals 12 dry intervals (detrended δ18O > 0) centered at 70.8, 77.2, 87.1, 90.3, 94, 98, 99.8, 105.1, 110.3 to 115.9, 120, 122.6, and 126.0 ka BP (Fig. 4G). Most of these events are observed in the Guatemala speleothem δ18O records (17), generally coinciding with Greenland cold stadials from GS20 to GS26 (Fig. 4A, F and G) (44, 45), cooler SSTs over the western and eastern tropical-subtropical North Atlantic (Fig. 4B) (37, 38, 4649), relatively weakened AMOC states (Fig. 4D) (37), and southward-shifted ITCZ (Fig. 4E and fig. S9) (5052). The interrelated ocean-atmospheric changes among high-mid– and low latitudes for the cold millennial events are consistent with previously suggested atmospheric/oceanic teleconnections among the Caribbean Sea, tropical Atlantic, AMOC, and ITCZ (11, 12, 14, 51, 53, 54). These climate patterns are also akin to Heinrich Stadials in glacial periods (4, 11, 12, 14). Model simulations suggest that cooler SSTs suppress convective rainfall, leading to broad dry conditions in the Mesoamerican monsoon domain and influencing the summer hydroclimate of the AWP (14).

Fig. 4. Millennial variability between 66 and 130 ka BP.

Fig. 4.

(A) δ18O records between 60 and 122 ka from The North Greenland Ice Core Project (NGRIP) ice core in GICC05modelext age scale (44) and between 114 and 130 ka from North Greenland Eemian Ice Drilling (NEEM) in EPICA Dronning Maud Land (EDML) age scale (45). (B) The first principal component (PC1) of eight z-scored SST records [MD01-2444 (46), MD03-2616 (38), MD02-2575 (47), ODP 999A (48), KNR191-CDH19 (37), MD95-2040 (49), MD95-2042 (49), and MD99-2331 (49)] over North Atlantic between 60 and 130 ka BP (see fig. S10). SST record of MD01-2444 (46) from North Atlantic at mid-latitude. (C) Ice-rafted debris (IRD) records from North Atlantic marine core (ODP 980; brown shaded) (40) and Gardar Drift (ODP983; purple dot line) (41). (D) Pa/Th records with 2σ error shaded (37). (E) Mo record of MD03-2622 (blue) (51) in Cariaco Basin. (F) Speleothem δ18O records from Guatemala (17). (G) The detrended CUST-1 δ18O record. The inferred dry (detrend δ18O > 0) and pluvial interval (detrend δ18O < 0) are shaded with red and blue, respectively. The one SD of detrended δ18O is marked with a horizontal dashed line at +1σ and –1σ, respectively. The highest value of summer insolation at 126 and 105 ka was marked by yellow circles. The dashed line of the CUST-1 δ18O record between 103 and 111 ka BP indicates large age uncertainty. Six gray bars show the dry events corresponding to the Greenland Stadials 20 to 26. Red thin bars show the centennial dry events. The discrete droughts with detrended δ18O higher than 1σ are highlighted by dark blue bars in the bottom. The climate meaning of each proxy is labeled in arrows aside. NAtlantic, North Atlantic; VSMOW, Vienna standard mean ocean water. cps, counts per second.

The two extreme dry events at 126 and 105 ka BP in our detrended record stand out as the most anomalous of the past 130 kyr, exceeding +3σ above the long-term mean, with δ18O values more than 0.8‰ heavier than those of other contiguous dry events (detrended δ18O > 1σ) (Fig. 4G). Both events occurred during periods of exceptionally high summer insolation (34), cooler North Atlantic SSTs (46), and weakened AMOC conditions triggered by freshwater inputs (Fig. 4, B, C, and G) (41). While similar oceanic conditions were associated with other events, such as those at 115.5 and 87.1 ka BP, the δ18O anomalies at 126 and 105 ka BP are notably more extreme (Fig. 4 and fig. S10). This suggests that oceanic forcing alone cannot explain their magnitude. We propose that the unusually high summer insolation plays a role in amplifying and prolonging midsummer drying across the AWP. At 35°N, summer insolation peaked at ~534.5 W/m2 around 126 ka BP and ~530 W/m2 around 105 ka BP, values that exceeded other precession maxima by at least 10 W/m2 (Fig. 4G). This extreme insolation likely extended the seasonal intensification and westward reach of the NASH, increasing subsidence and suppressing convection over the AWP, thereby leading to overall drier summers.

In summary, our speleothem data integrated with climate model simulation highlight the central role of the NASH in modulating hydroclimate variability across the AWP over the past 129,000 years. On orbital timescales, enhanced rainfall coincided with strong AMOC and sustained tropical Atlantic SSTs above the convective threshold. On precessional scales, the amplitude of rainfall variability was governed by a coupled response to summer insolation and SST, which amplified land-sea thermal gradients and intensified the NASH. This mechanism reduced rainfall across a broad swath of the Caribbean while promoting wetter conditions over eastern North America.

Notably, two exceptionally dry events at 126 and 105 ka BP were linked to both cooler SSTs and peak summer insolation, suggesting that insolation extremes may prolong and intensify the midsummer dry spell via westward NASH expansion. These findings underscore how subtropical atmospheric dynamics, typically underemphasized in orbital-scale hydroclimate studies, can modulate tropical rainfall responses. As modern warming trends continue, projections point to a strengthening and westward-shifting NASH (55), with implications for increasingly arid summers across the AWP (56). Future work should investigate how this dynamical regime may interact with anthropogenic forcing, especially in a warming world where land-sea contrasts are expected to intensify further.

MATERIALS AND METHODS

230Th dating

A total of 79 dates of the speleothem sample CUST-1 between 0 and 337 mm was performed at Xi’an Jiaotong University, China. Twenty-six dates were measured between 0 and 94 mm for the Holocene section, and 53 dates were measured between 94 and 337 mm for the late Pleistocene. We used standard chemistry procedures (57) to separate uranium and thorium by using Thermo Fisher Scientific Finnigan Neptune Plus multicollector inductively coupled plasma mass spectrometers. A triple-spike (229Th-233U-236U) isotope dilution method was used to correct instrumental fractionation and to determine U/Th isotopic ratios and concentrations (58). U and Th isotopes were measured on a MasCom multiplier behind the retarding potential quadrupole in the peak-jumping mode using standard procedures. Uncertainties in U and Th isotopic measurements were calculated offline at the 2σ levels, including corrections for blanks, multiplier dark noise, abundance sensitivity, and contents of the same nuclides in the spike solution. 234U and 230Th decay constants of Cheng et al. (58) were used. Corrected 230Th ages assume an initial 230Th/232Th atomic ratio of (4.4 ± 2.2) × 10−6, and the value for material at secular equilibrium with the bulk earth 232Th/238U value is 3.8. The correction for the powder samples with too high detrital 232Th concentrations (>10,000 parts per thousand) leads to large errors of ages centered around 189 to 193 and 292 mm (dataset S2).

High detrital 232Th concentrations may have introduced large uncertainties in the initial 230Th estimates and resulted in large uncertainties in the corrected ages. To a certain extent, in this case, a proper 230Th/232Th atomic ratio used would be crucial for the high 232Th dates to be close to their true ages. Hence, we constrain the 230Th/232Th ratio as the value of (22 ± 6) × 10−6 and then recalculate the ages for the three high 232Th dates between 189 to 193 mm (see Supplementary Text, dataset S2, and fig. S5A).

Stable isotope analysis

Stable isotope (δ18O) records of CUST-1 were established by ~1950 measurements. We used a computer-controlled triaxial micromill (New Wave MicroMill) to obtain subsamples (~80 μg) that were continuously milled along the growth axes. The increments are 2 mm between 0 and 94 mm and range between 0.1 and 0.2 mm for 94.4 and 336.8 mm. The subsamples were measured using a Thermo Fisher Scientific Delta V mass spectrometer coupled with an online carbonate preparation system (Kiel IV) in the Isotope Laboratory, Xi’an Jiaotong University. All results are reported in per mil (‰) relative to the Vienna Pee Dee Belemnite (VPDB) standard. Duplicate measurements of standards show a long-term reproducibility of 0.1‰ (1σ) or better for O isotope (Supplementary Text and dataset S1).

Age models

We used MOD-AGE (59) and StalAge (28) to constrain the age model of CUST-1 (fig. S5). The results are insensitive to the choice of age model. We used the age model conducted by StalAge for further analysis (see Supplementary Text). The age model of CUST-1 was conducted with 26 dates for the Holocene. The age model of CUST-1 between 129 and 66 ka BP was established by three age depth curves using (i) 47 ages with initial 230Th/232Th atomic ratio as (4.4 ± 2.2) × 10−6 and no set hiatus (fig. S6A) and (ii) 47 ages as mentioned above, setting an hiatus at 188.2 mm. Introducing this hiatus into the age model results in a ~6.7-kyr gap between 109.5 and 102.9 ka BP (fig. S6A); (iii) 47 ages as mentioned, three additional recalculated ages at 189, 192, and 192.1 mm with initial 230Th/232Th ratio as (22 ± 6) × 10−6 and an hiatus set at 188.2 mm (figs. S5 and S6B). Introducing this hiatus and three additional ages into the age model results in a 1.7-kyr gap between 104.5 and 102.8 ka BP. Considering that the uncertainty of all age model remains large between 112 and 103 ka BP, we show the curves for this period (from 193 to 186.5 mm) as dashed line instead of solid line in all figures to indicate the substantial uncertainty.

X-ray diffraction analysis

The dry intervals indicated by positive excursion in orbital detrended speleothem δ18O records coincide with light-yellow thin bands in the sample surface. To make sure of the minerology composition of these layers, we conducted x-ray diffraction (XRD) analyses in Xi’an Jiaotong University. Powder XRD patterns of samples were recorded on a Bruke D8 ADVANCE operating at 40-KV voltage and 15-mA current with Cu-Kα radiation (λ = 0.15406 nm). The results confirm that these bands are calcite, indicating no change in mineralogy during these dry intervals (fig. S5).

Idealized precession simulation, EC-Earth-2-2

We use the pmin and pmax simulations of EC-Earth-2-2 as performed in (30) to investigate the annual cycle and spatial pattern changes in precipitation and NASH. EC-Earth-2-2 is a fully coupled model with an atmospheric horizontal resolution of T159 (roughly 1.125° × 1.125°) and 62 vertical levels and an ocean model at 1° resolution and 42 vertical levels. The pmin simulation sets obliquity at 22.08°, eccentricity at 0.056, and longitude of perihelion at 95.96°; the pmax simulation has the same obliquity, eccentricity at 0.058, and longitude of perihelion at 273.50°. In addition, orbital extreme experiments used here are in general agreement with other modeling work (57) used to investigate North African monsoon (30) and Asian monsoon (60).

Time series analysis

Detrending

We applied ensemble empirical mode decomposition (EEMD) to identify the orbital and millennial variability of 33-year evenly spaced CUST-1 δ18O. EEMD uses an ensemble distribution of data generated by adding white noise to the source data and then iteratively removing the highest-frequency oscillatory components superimposed on lower frequency signals (61). EEMD decomposed the time series into 10 intrinsic mode functions (IMF) with IMF 1 to 3 as noise or high-frequency variability, IMF 4 to 7 as centennial to millennial quasiperiodic oscillations, and a sum of IMF 8 to 10 as the orbital trend. We remove the orbital trend for CUST-1 δ18O to get detrended data.

Wavelet analysis

Wavelet power spectrum of the detrended CUST-1 δ18O and cross wavelet coherence of Cuba speleothem δ18O and June-July insolation at 35°N were conducted in MATLAB (62) (fig. S7D). The cone of influence, where edge effects become important, is shown by the black arc. Irregular black curves delineate time-frequency regions that are greater than 95% confidence for a red-noise (AR1) process.

Insolation

We calculate the mean summer insolation averaged over June and July (1 June to 30 July) at 30°N, 35° N, 40°N, and 65°N (34). The comparison of Cuba δ18O records with insolation curves shows the similarity between Cuba δ18O records and insolation curves at mid-latitude rather than at 65°N, which excludes the high latitude impacts (fig. S7).

Modern climate analysis

We used the SST, precipitation, geopotential height at 850 hPa, and oxygen isotope of precipitation from globally gridded SSTs from Extended Reconstructed SST (ERSST) version 5 (63), globally gridded precipitation from the Global Precipitation Climatology Project (GPCP) monthly analysis version 2.3 (64), globally gridded monthly averaged data on pressure levels from 1940 to present at 850 hPa from the European Centre for Medium Range Weather Forecasts Reanalysis fifth-generation (ERA5) (65), and globally gridded monthly isotope-enabled general circulation climate model version 2 (IsoGSM2) (26), respectively. The u and v winds at 850 hPa were derived from the National Centers for Environmental Prediction/National Center for Atmospheric Research (NCEP/NCAR) Reanalysis 1 data products (66).

The EOF analysis of normalized MJJASO SSTA was set over the region 0° to 70°N and 90°W to 0°W (fig. S3B). The EOF-1 (explained variance ~32%) shows an SST triple pattern akin to the spatial pattern of correlation between regional precipitation and SSTA over the North Atlantic in fig. S3C.

Acknowledgments

We thank X. Zhang and J. Cheng for discussion and comments on this work. We acknowledge the support from Q. Yin and C. J. R. Williams for sharing the model data. We sincerely thank the editors and reviewers for the constructive comments and suggestions.

Funding:

Funding information is listed as follows: the National Natural Science Foundation of China grant 42488201 (H.C.), the National Natural Science Foundation of China grant 42261144753 (H.Z.), the US National Science Foundation grant ATM-1003502 (A.W.), the Young Scientists Fund of the National Natural Science Foundation of China grant 42102229 (H.L.), the China Postdoctoral Science Foundation grant 2021 M692522 (H.L.), and the Fundamental Research Funds for the Central Universities (H.L.).

Author contributions:

Conceptualization: A.W., A.S., H.C., H.L., and G.K. Methodology: A.S., H.C., J.H., P.D., H.L., L.Y., and G.K. Investigation: S.L., A.S., A.W., H.L., Y.N., G.K., and H.C. Resources: A.W., H.C., Y.N., and L.S. Visualization: A.S., H.L., and G.K. Supervision: H.C., A.S., and A.W. Validation: H.Z., H.L., G.K., H.C., and L.S. Data Curation: H.L. and H.C. Formal analysis: A.S., H.L., and L.S. Project administration: A.W., H.C., and H.L. Writing—original draft: H.L., A.S., and H.C. Writing—review and editing: H.L. A.W., A.S., S.W., H.C., X.D., L.S., G.K., and H.Z. Funding acquisition: H.C., H.L., A.W., and H.Z.

Competing interests:

All authors declare that they have no competing interests.

Data and materials availability:

All data are available in the public domain as follows: (i) globally gridded SSTs from ERSST version 5, available at www.ncei.noaa.gov/products/extended-reconstructed-sst; (ii) globally gridded precipitation from the GPCP monthly analysis version 2.3, available at https://climatedataguide.ucar.edu/climate-data/gpcp-monthly-global-precipitation-climatology-project; (iii) globally gridded monthly averaged data on pressure levels from 1940 to present at 850 hPa from the ERA5, available at https://cds.climate.copernicus.eu/datasets/reanalysis-era5-pressure-levels?tab=overview; (iv) globally gridded monthly IsoGSM2, available at http://isotope.iis.u-tokyo.ac.jp/~kei/tmp/isogsm2; (v) the idealized precession simulation EC-Earth-2-2, available at https://zenodo.org/records/5054419; (vi) temperature, precipitation, and δ18Op from GNIP Havana (CPHR) station at 2008 CE and δ18Op data from GNIP stations in the Antilles of the West Indies are available at https://www.iaea.org/services/networks/gnip; and (vii) globally gridded u and v winds at 850 hPa from https://psl.noaa.gov/data/gridded/data.ncep.reanalysis.html. All links were deposited on 25 September 2025. All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials.

Supplementary Materials

The PDF file includes:

Supplementary Text

Figs. S1 to S10

Legends for datasets S1 and S2

References

sciadv.aea5042_sm.pdf (5.8MB, pdf)

Other Supplementary Material for this manuscript includes the following:

Datasets S1 and S2

REFERENCES AND NOTES

  • 1.Lachniet M. S., Burns S. J., Piperno D. R., Asmerom Y., Polyak V. J., Moy C. M., Christenson K., A 1500-year El Niño/Southern Oscillation and rainfall history for the Isthmus of Panama from speleothem calcite. J. Geophys. Res. 109, D20117 (2004). [Google Scholar]
  • 2.Winte A., Miller T., Kushnir Y., Shina A., Timmermann A., Jury M. R., Gallup C., Cheng H., Edwards R. L., Evidence for 800 years of North Atlantic multi-decadal variability from a Puerto Rican speleothem. Earth Planet. Sci. Lett. 308, 23–28 (2011). [Google Scholar]
  • 3.Winter A., Zanchettin D., Miller T., Kushnir Y., Black D., Lohmann G., Burnett A., Haug G. H., Estrella-Martínez J., Breitenbach S. F. M., Beaufort L., Rubino A., Cheng H., Persistent drying in the tropics linked to natural forcing. Nat. Commun. 6, 7627 (2015). [DOI] [PubMed] [Google Scholar]
  • 4.Warken S. F., Scholz D., Spötl C., Jochum K. P., Pajón J. M., Bahr A., Mangini A., Caribbean hydroclimate and vegetation history across the last glacial period. Quat. Sci. Rev. 218, 75–90 (2019). [Google Scholar]
  • 5.Warken S. F., Vieten R., Winter A., Spötl C., Miller T. E., Jochum K. P., Schröder-Ritzrau A., Mangini A., Scholz D., Persistent link between Caribbean precipitation and Atlantic Ocean circulation during the last glacial revealed by a speleothem record from Puerto Rico. Paleoceanogr. Paleoclimatol. 35, e2020PA003944 (2020). [Google Scholar]
  • 6.Martinez C., Goddard L., Kushnir Y., Ting M., Seasonal climatology and dynamical mechanisms of rainfall in the Caribbean. Clim. Dyn. 53, 825–846 (2019). [Google Scholar]
  • 7.Bregy J. C., Maxwell J. T., Robeson S. M., Harley G. L., Elliott E. A., Heeter K. J., US Gulf Coast tropical cyclone precipitation influenced by volcanism and the North Atlantic subtropical high. Commun. Earth Environ. 3, 164 (2022). [Google Scholar]
  • 8.Small R. J. O., De Szoeke S. P., Xie S.-P., The Central American midsummer drought: Regional aspects and large-scale forcing. J. Clim. 20, 4853–4873 (2007). [Google Scholar]
  • 9.Li L., Li W., Kushnir Y., Variation of the North Atlantic subtropical high western ridge and its implication to southeastern US summer precipitation. Clim. Dyn. 39, 1401–1412 (2012). [Google Scholar]
  • 10.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. U.S.A. 110, 9255–9260 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Arienzo M. M., Swart P. K., Broad K., Clement A. C., Pourmand A., Kakuk B., Multi-proxy evidence of millennial climate variability from multiple Bahamian speleothems. Quat. Sci. Rev. 161, 18–29 (2017). [Google Scholar]
  • 12.Arienzoe M. M., Swart P. K., Pourmand A., Broad K., Clement A. C., Murphy L. N., Vonhof H. B., Kakuk B., Bahamian speleothem reveals temperature decrease associated with Heinrich stadials. Earth Planet. Sci. Lett. 430, 377–386 (2015). [Google Scholar]
  • 13.Ramirez V. M., Cruz F. W., Vuille M., Novello V. F., Stríkis N. M., Cheng H., Zhang H. W., Bernal J. P., Du W. J., Ampuero A., Deininger M., Chiessi C. M., Tejedor E., Campos J. L., Brahim Y. A., Edwards R. L., Summer insolation controlled movements of intertropical convergence zone during last glacial cycle in Northern South America. Commun. Earth Environ. 4, 495 (2023). [Google Scholar]
  • 14.Wright K. T., Johnson K. R., Marks G. S., McGee D., Bhattacharya T., Goldsmith G. R., Tabor C. R., Lacaille-Muzquiz J.-L., Lum G., Beramendi-Orosco L., Dynamic and thermodynamic influences on precipitation in Northeast Mexico on orbital to millennial timescales. Nat. Commun. 14, 2279 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Warken S. F., Weißbach T., Kluge T., Vonhof H., Scholz D., Vieten R., Schmidt M., Winter A., Frank N., Last glacial millennial-scale hydro-climate and temperature changes in Puerto Rico constrained by speleothem fluid inclusion δ18O and δ2H values. Clim. Past 18, 167–181 (2022). [Google Scholar]
  • 16.Winter A., Zanchettin D., Lachniet M., Vieten R., Pausata F. S. R., Ljungqvist F. C., Cheng H., Edwards R. L., Miller T., Rubinetti S., Rubino A., Taricco C., Initiation of a stable convective hydroclimatic regime in Central America circa 9000 years BP. Nat. Commun. 11, 716 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Lucia G., Zanchettin D., Winter A., Cheng H., Rubino A., Vásquez O. J., Bernal J. P., Cu-xi M., Lachniet M., Atlantic Ocean thermal forcing of Central American rainfall over 140,000 years. Nat. Commun. 15, 10586 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Lachniet M. S., Johnson L., Asmerom Y., Burns S. J., Polyak V., Patterson W. P., Burt L., Azouz A., Late quaternary moisture export across Central America and to Greenland: Evidence for tropical rainfall variability from Costa Rican stalagmites. Quat. Sci. Rev. 28, 3348–3360 (2009). [Google Scholar]
  • 19.Cheng H., Edwards R. L., Sinha A., Spötl C., Yi L., Chen S., Kelly M., Kathayat G., Wang X., Li X., Kong X., Wang Y., Ning Y., Zhang H., The Asian monsoon over the past 640,000 years and ice age terminations. Nature 534, 640–646 (2016). [DOI] [PubMed] [Google Scholar]
  • 20.Cheng H., Li H., Sha L., Sinha A., Shi Z., Yin Q., Lu Z., Zhao D., Cai Y., Hu Y., Hao Q., Tian J., Kathayat G., Dong X., Zhao J., Zhang H., Milankovitch theory and monsoon. Innovation 3, 100338 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Cruz F. W. Jr., Burns S. J., Karmann I., Sharp W. D., Vuille M., Cardoso A. O., Ferrari J. A., Dias P. L. S., Viana O., Insolation-driven changes in atmospheric circulation over the past 116,000 years in subtropical Brazil. Nature 434, 63–66 (2005). [DOI] [PubMed] [Google Scholar]
  • 22.Wang C., Enfield D. B., Lee S. K., Landsea C. W., Influences of the Atlantic warm pool on western hemisphere summer rainfall and Atlantic hurricanes. J. Clim. 19, 3011–3028 (2006). [Google Scholar]
  • 23.Herrera D., Ault T., Insights from a new high-resolution drought atlas for the Caribbean spanning 1950–2016. J. Clim. 30, 7801–7825 (2017). [Google Scholar]
  • 24.Pascale S., Kapnick S. B., Delworth T. L., Hidalgo H. G., Cooke W. F., Natural variability vs forced signal in the 2015–2019 Central American drought. Clim. Change 168, 16 (2021). [Google Scholar]
  • 25.S. P. Xie, J. A. Carton, “Tropical Atlantic variability: Patterns, mechanisms, and impacts,” in Earth’s Climate: The Ocean–Atmosphere Interaction, S. P. Xie, J. A. Carton, Eds. (Geophysical Monograph Series, American Geophysical Union, 2004), pp.121–142. [Google Scholar]
  • 26.Yoshimura K., Kanamitsu M., Noone D., Oki T., Historical isotope simulation using reanalysis atmospheric data. J. Geophys. Res. Atmos. 113, D19108 (2008). [Google Scholar]
  • 27.Fensterer C., Scholz D., Hoffmann D. L., Spötl C., Schröder-Ritzrau A., Horn C., Pajón J. M., Mangini A., Millennial-scale climate variability during the last 12.5 ka recorded in a Caribbean speleothem. Earth Planet. Sci. Lett. 361, 143–151 (2013). [Google Scholar]
  • 28.Scholz D., Hoffmann D. L., StalAge – An algorithm designed for construction of speleothem age models. Quat. Geochronol. 6, 369–382 (2010). [Google Scholar]
  • 29.Travis-Taylor L., Medina-Elizalde M., Pajón J. M., Karmalkar A. V., Polanco-Martínez J. M., Kinsley C. W., McGee D., Hydroclimate variability in the northern Caribbean during the last deglaciation was modulated by large-scale atmospheric circulation and climate events. Commun. Earth Environ. 6, 498 (2025). [Google Scholar]
  • 30.Bosmans J. H. C., Drijfhout S. S., Tuenter E., Hilgen F. J., Lourens L. J., Response of the North African summer monsoon to precession and obliquity forcings in the EC-Earth GCM. Clim. Dyn. 44, 279–297 (2015). [Google Scholar]
  • 31.Hazeleger W., Severijns C., Semmler T., Ştefǎnescu S., Yang S., Wang X., Wyser K., Dutra E., Baldasano J. M., Bintanja R., Bougeault P., Caballero R., Ekman A. M. L., Christensen J. H., Hurk B. V. D., Jimenez P., Jones C., Kållberg P., Koenigk T., Miranda R. M., Noije T. V., Palmer T., Parodi J. A., Schmith T., Selten F., Storelvmo T., Sterl A., Tapamo H., Vancoppenolle M., Viterbo P., Willén U., EC-Earth: A seamless earth system prediction approach in action. Bull. Am. Meteorol. Soc. 91, 1357–1364 (2010). [Google Scholar]
  • 32.Hazeleger W., Wang X., Severijns C., Ştefănescu S., Bintanja R., Sterl A., Wyser K., Semmler T., Yang S., van den Hurk B., van Noije T., van der Linden E., van der Wiel K., EC-Earth V2.2: Description and validation of a new seamless earth system prediction model. Clim. Dyn. 39, 2611–2629 (2011). [Google Scholar]
  • 33.Otto-Bliesner B. L., Brady E. C., Zhao A., Brierley C. M., Axford Y., Capron E., Govin A., Hoffman J. S., Isaacs E., Kageyama M., Scussolini P., Tzedakis P. C., Williams C. J. R., Wolff E., Abe-Ouchi A., Braconnot P., Buarque S. R., Cao J., de Vernal A., Guarino M. V., Guo C., LeGrande A. N., Lohmann G., Meissner K. J., Menviel L., Morozova P. A., Nisancioglu K. H., O'ishi R., Mélia D. S., Shi X., Sicard M., Sime L., Stepanek C., Tomas R., Volodin E., Yeung N. K. H., Zhang Q., Zhang Z., Zheng W., Large-scale features of Last Interglacial climate: Results from evaluating the lig127k simulations for the Coupled Model Intercomparison Project (CMIP6)–Paleoclimate Modeling Intercomparison Project (PMIP4). Clim. Past 17, 63–94 (2021). [Google Scholar]
  • 34.Laskar J., Robutel P., Joutel F., Gastineau M., Correia A. C. M., Levrard B., A long-term numerical solution for the insolation quantities of the Earth. Astron. Astrophys. 428, 261–285 (2004). [Google Scholar]
  • 35.Singarayer J. S., Valdes P. J., Roberts W. H. G., Ocean dominated expansion and contraction of the late Quaternary tropical rainbelt. Sci. Rep. 7, 9382 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Cheng H., Springer G. S., Sinha A., Hardt B. F., Yi L., Li H., Tian Y., Li X., Rowe H. D., Kathayat G., Ning Y., Edwards R. L., Eastern North American climate in phase with fall insolation throughout the last three glacial-interglacial cycles. Earth Planet. Sci. Lett. 522, 125–134 (2019). [Google Scholar]
  • 37.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, 470–474 (2016). [DOI] [PubMed] [Google Scholar]
  • 38.Rama-Corredor O., Martrat B., Grimalt J. O., López-Otalvaro G. E., Flores J. A., Sierro F., Parallelisms between sea surface temperature changes in the western tropical Atlantic (Guiana Basin) and high latitude climate signals over the last 140 000 years. Clim. Past 11, 1297–1311 (2015). [Google Scholar]
  • 39.Nascimento R. A., Johnstone H. J. H., Kuhnert H., Santos T. P., Venancio I. M., Chiessi C. M., Ballalai J. M., Campos C., Govin A., Mulitza S., Albuquerque A. L. S., Long-term variability of the western tropical Atlantic sea surface temperature driven by greenhouse gases and AMOC. Quat. Sci. Rev. 322, 108431 (2023). [Google Scholar]
  • 40.McManus J. F., Oppo D. W., Cullen J. L., A 0.5-million-year record of millennial-scale climate variability in the North Atlantic. Science 283, 971–975 (1999). [DOI] [PubMed] [Google Scholar]
  • 41.Barker S., Knorr G., Conn S., Lordsmith S., Newman D., Thornalley D., Early interglacial legacy of deglacial climate instability. Paleoceanogr. Paleoclimatol. 34, 1455–1475 (2019). [Google Scholar]
  • 42.Miyasaka T., Nakamura H., Structure and formation mechanisms of the Northern Hemisphere summertime subtropical highs. J. Clim. 18, 5046–5065 (2005). [Google Scholar]
  • 43.James A., Emile-Geay J., Partin J. W., Khider D., Global speleothem analysis reveals state-dependent hydrological response to orbital forcing. Paleoceanogr. Paleoclimatol. 40, e2024PA005098 (2025). [Google Scholar]
  • 44.Rasmussen S. O., Bigler M., Blockley S. P., Blunier T., Buchardt S. L., Clausen H. B., Cvijanovic I., Dahl-Jensen D., Johnsen S. J., Fischer H., Gkinis V., Guillevic M., Hoek W. Z., Lowe J. J., Pedro J. B., Popp T., Seierstad I. K., Steffensen J. P., Svensson A. M., Vallelonga P., Vinther B. M., Walker M. J. C., Wheatley J. J., Winstrup M., A stratigraphic framework for abrupt climatic changes during the last glacial period based on three synchronized Greenland ice-core records: Refining and extending the INTIMATE event stratigraphy. Quat. Sci. Rev. 106, 14–28 (2014). [Google Scholar]
  • 45.Dahl-Jensen D., Albert M. R., Aldahan A., Azuma N., Balslev-Clausen D., Baumgartner M., Berggren A.-M., Bigler M., Binder T., Blunier T., Bourgeois J. C., Brook E. J., Buchardt S. L., Buizert C., Capron E., Chappellaz J., Chung J., Clausen H. B., Cvijanovic I., Davies S. M., Ditlevsen P., Eicher O., Fischer H., Fisher D. A., Fleet L. G., Gfeller G., Gkinis V., Gogineni S., Goto-Azuma K., Grinsted A., Gudlaugsdottir H., Guillevic M., Hansen S. B., Hansson M., Hirabayashi M., Hong S., Hur S. D., Huybrechts P., Hvidberg C. S., Iizuka Y., Jenk T., Johnsen S. J., Jones T. R., Jouze J., Karlsson N. B., Kawamura K., Keegan K., Kettner E., Kipfstuhl S., Kjær H. A., Koutnik M., Kuramoto T., Köhler P., Laepple T., Landais A., Langen P. L., Larsen L. B., Leuenberger D., Leuenberger M., Leuschen C., Li J., Lipenkov V., Martinerie P., Maselli O. J., Masson-Delmotte V., McConnell J. R., Miller H., Mini O., Miyamoto A., Montagnat-Rentier M., Mulvaney R., Muscheler R., Orsi A. J., Paden J., Panton C., Pattyn F., Petit J.-R., Pol K., Popp T., Possnert G., Prié F., Prokopiou M., Quiquet A., Rasmussen S. O., Raynaud D., Ren J., Reutenauer C., Ritz C., Röckmann T., Rosen J. L., Rubino M., Rybak O., Samyn D., Sapart C. J., Schilt A., Schmidt A. M. Z., Schwander J., Schüpbach S., Seierstad I., Severinghaus J. P., Sheldon S., Simonsen S. B., Sjolte J., Solgaard A. M., Sowers T., Sperlich P., Steen-Larsen H. C., Steffen K., Steffensen J. P., Steinhage D., Stocker T. F., Stowasser C., Sturevik A. S., Sturges W. T., Sveinbjörnsdottir A., Svensson A., Tison J.-L., Uetake J., Vallelonga P., van de Wal R. S. W., van der Wel G., Vaughn B. H., Vinther B., Waddington E., Wegner A., Weikusat I., White J. W. C., Wilhelms F., Winstrup M., Witrant E., Wolff E. W., Xiao C., Zheng J., Eemian interglacial reconstructed from a Greenland folded ice core. Nature 493, 489–494 (2013). [DOI] [PubMed] [Google Scholar]
  • 46.Martrat B., Grimalt J. O., Shackleton N. J., de Abreu L., Hutterli M. A., Stocker T. F., Four climate cycles of recurring deep and surface water destabilizations on the iberian margin. Science 317, 502–507 (2007). [DOI] [PubMed] [Google Scholar]
  • 47.Nürnberg D., Ziegler M., Karas C., Tiedemann R., Schmidt M. W., Interacting loop current variability and Mississippi river discharge over the past 400 kyr. Earth Planet. Sci. Lett. 272, 278–289 (2008). [Google Scholar]
  • 48.Schmidt M. W., Spero H. J., Lea D. W., Links between salinity variation in the Caribbean and North Atlantic thermohaline circulation. Nature 428, 160–163 (2004). [DOI] [PubMed] [Google Scholar]
  • 49.Davtian N., Bard E., Darfeuil S., Ménot G., Rostek F., The novel hydroxylated tetraether index RI-OH′ as a sea surface temperature proxy for the 160-45 ka BP period off the Iberian Margin. Paleoceanogr. Paleoclimatol. 36, e2020PA004077 (2021). [Google Scholar]
  • 50.Deplazes G., Lückge A., Peterson L. C., Timmermann A., Hamann Y., Hughen K. A., Röhl U., Laj C., Cane M. A., Sigman D. M., Haug G. H., Links between tropical rainfall and North Atlantic climate during the last glacial period. Nat. Geosci. 6, 213–217 (2013). [Google Scholar]
  • 51.Gibson K. A., Peterson L. C., A 0.6 million year record of millennial-scale climate variability in the tropics. Geophys. Res. Lett. 41, 969–975 (2014). [Google Scholar]
  • 52.Zhuravleva A., Hüls M., Tiedemann R., Bauch H. A., A 125-ka record of northern South American precipitation and the role of high-to-low latitude teleconnections. Quat. Sci. Rev. 270, 107159 (2021). [Google Scholar]
  • 53.Bhattacharya T., Chiang J. C., Cheng W., Ocean-atmosphere dynamics linked to 800–1050 CE drying in Mesoamerica. Quat. Sci. Rev. 169, 263–277 (2017). [Google Scholar]
  • 54.Ait Brahim Y., Peros M. C., Viau A. E., Liedtke M., Pajón J. M., Valdes J., Li X., Edwards R. L., Reinhardt E. G., Oliva F., Hydroclimate variability in the Caribbean during North Atlantic Heinrich cooling events (H8 and H9). Sci. Rep. 12, 20719 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Li W., Li L., Ting M., Liu Y., Intensification of Northern Hemisphere subtropical highs in a warming climate. Nat. Geosci. 5, 830–834 (2012). [Google Scholar]
  • 56.Campbell J. D., Taylor M. A., Bezanilla-Morlot A., Stephenson T. S., Centella-Artola A., Clarke L. A., Stephenson K. A., Generating projections for the Caribbean at 1.5, 2.0 and 2.5°C from a high-resolution ensemble. Atmosphere 12, 328 (2021). [Google Scholar]
  • 57.Edwards R. L., Chen J. H., Wasserburg G. J., 238U–234U–230Th–232Th systematics and the precise measurement of time over the past 500,000 years. Earth Planet. Sci. Lett. 81, 175–192 (1987). [Google Scholar]
  • 58.Cheng H., Edwards R. L., Shen C.-C., Polyak V. J., Asmerom Y., Woodhead J., Hellstrom J., Wang Y., Kong X., Spötl C., Wang X., Alexander E. C. Jr., Improvements in 230Th dating, 230Th and 234U half-life values, and U–Th isotopic measurements by multi-collector inductively coupled plasma mass spectrometry. Earth Planet. Sci. Lett. 371, 82–91 (2013). [Google Scholar]
  • 59.Hercman H., Pawlak J., MOD-AGE: An age-depth model construction algorithm. Quat. Geochronol. 12, 1–10 (2012). [Google Scholar]
  • 60.Bosmans J. H. C., Erb M. P., Dolan A. M., Drijfhout S. S., Tuenter E., Hilgen F. J., Edge D., Pope J. O., Lourens L. J., Response of the Asian summer monsoons to idealized precession and obliquity forcing in a set of GCMs. Quat. Sci. Rev. 188, 121–135 (2018). [Google Scholar]
  • 61.Wu Z., Huang N. E., Ensemble empirical mode decomposition: A noise-assisted data analysis method. Adv. Adapt. Data Anal. 01, 1–41 (2009). [Google Scholar]
  • 62.Grinsted A., Moore J. C., Jevrejeva S., Application of the cross wavelet transform and wavelet coherence to geophysical time series. Nonlinear Processes Geophys. 11, 561–566 (2004). [Google Scholar]
  • 63.Huang B., Thorne P. W., Banzon V. F., Boyer T., Chepurin G., Lawrimore J. H., Menne M. J., Smith T. M., Vose R. S., Zhang H.-M., Extended reconstructed sea surface temperature, version 5 (ERSSTv5): Upgrades, validations, and intercomparisons. J. Clim. 30, 8179–8205 (2017). [Google Scholar]
  • 64.Adler R. F., Sapiano M. R. P., Huffman G. J., Wang J.-J., Gu G., Bolvin D., Chiu L., Schneider U., Becker A., Nelkin E., Xie P., Ferraro R., Shin D., The Global Precipitation Climatology Project (GPCP) monthly analysis (new version 2.3) and a review of 2017 global precipitation. Atmosphere 9, 138 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Hersbach H., Bell B., Berrisford P., Hirahara S., Horányi A., Muñoz-Sabater J., Nicolas J., Peubey C., Radu R., Schepers D., Simmons A., Soci C., Abdalla S., Abellan X., Balsamo G., Bechtold P., Biavati G., Bidlot J., Bonavita M., De Chiara G., Dahlgren P., Dee D., Diamantakis M., Dragani R., Flemming J., Forbes R., Fuentes M., Geer A., Haimberger L., Healy S., Hogan R. J., Hólm E., Janisková M., Keeley S., Laloyaux P., Lopez P., Lupu C., Radnoti G., de Rosnay P., Rozum I., Vamborg F., Villaume S., Thépaut J., The ERA5 global reanalysis. Q. J. Roy. Meteorol. Soc. 146, 1999–2049 (2020). [Google Scholar]
  • 66.Kalnay E., Kanamitsu M., Kistler R., Collins W., Deaven D., Gandin L., Iredell M., Saha S., White G., Woollen J., Zhu Y., Chelliah M., Ebisuzaki W., Higgins W., Janowiak J., Mo K. C., Ropelewski C., Wang J., Leetmaa A., Reynolds R., Jenne R., Joseph D., The NCEP/NCAR 40-year reanalysis project. Bull. Am. Meteorol. Soc. 77, 437–471 (1996). [Google Scholar]
  • 67.Li H., Sinha A., André A. A., Spötl C., Vonhof H. B., Meunier A., Kathayat G., Duan P., Li X., Sha L., Zhao J., Edwards R. L., Cheng H., A multimillennial climatic context for the megafaunal extinctions in Madagascar and Mascarene Islands. Sci. Adv. 6, eabb2459 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Moseley G. E., Spötl C., Brandstätter S., Erhardt T., Luetscher M., Edwards R. L., NALPS19: Sub-orbital-scale climate variability recorded in northern Alpine speleothems during the last glacial period. Clim. Past 16, 29–50 (2020). [Google Scholar]
  • 69.Kim S.-T., O’Neil J. R., Hillaire-Marcel C., Mucci A., Oxygen isotope fractionation between synthetic aragonite and water: Influence of temperature and Mg2+ concentration. Geochim. Cosmochim. Acta 71, 4704–4715 (2007). [Google Scholar]
  • 70.O’Neil J. R., Clayton R. N., Mayeda T. K., Oxygen isotope fractionation in divalent metal carbonates. J. Chem. Phys. 51, 5547–5558 (1969). [Google Scholar]
  • 71.Tremaine D. M., Froelich P. N., Wang Y., Speleothem calcite farmed in situ: Modern calibration of δ18O and δ13C paleoclimate proxies in a continuously-monitored natural cave system. Geochim. Cosmochim. Acta 75, 4929–4950 (2011). [Google Scholar]
  • 72.Coplen T. B., Calibration of the calcite–water oxygen-isotope geothermometer at Devils Hole, Nevada, a natural laboratory. Geochim. Cosmochim. Acta 71, 3948–3957 (2007). [Google Scholar]
  • 73.Gabitov R. I., Watson E. B., Sadekov A., Oxygen isotope fractionation between calcite and fluid as a function of growth rate and temperature: An in situ study. Chem. Geol. 306-307, 92–102 (2012). [Google Scholar]
  • 74.Demény A., Kele S., Siklósy Z., Empirical equations for the temperature dependence of calcite-water oxygen isotope fractionation from 10 to 70°C: Empirical equations for the calcite-water oxygen isotope fractionation. Rapid Commun. Mass Spectrom. 24, 3521–3526 (2010). [DOI] [PubMed] [Google Scholar]
  • 75.C. Fensterer, “Holocene Caribbean climate variability reconstructed from speleothems from Western Cuba,” thesis, Ruperto-Carola University of Heidelberg (2011). [Google Scholar]
  • 76.Nilsson-Kerr K., Anand P., Holden P. B., Clemens S. C., Leng M. J., Dipole patterns in tropical precipitation were pervasive across landmasses throughout marine isotope stage 5. Commun. Earth Environ. 2, 64 (2021). [Google Scholar]
  • 77.Li Y., Rao Z., Xu Q., Zhang S., Liu X., Wang Z., Cheng H., Edwards R. L., Chen F., Inter-relationship and environmental significance of stalagmite δ13C and δ18O records from Zhenzhu Cave, north China, over the last 130 ka. Earth Planet. Sci. Lett. 536, 116149 (2020). [Google Scholar]
  • 78.Zhang R., Yang X., Zhang H., Zhang R., Yan Y., Zuli S., Wang Y., Centennial-to millennial-scale Asian summer monsoon changes during the MIS 5/4 transition revealed by high-resolution stalagmite records from southwestern China. Palaeogeogr. Palaeoclimatol. Palaeoecol. 571, 110390 (2021). [Google Scholar]
  • 79.Wu Y., Li T.-Y., Yu T.-L., Shen C.-C., Chen C.-J., Zhang J., Li J.-Y., Wang T., Huang R., Xiao S.-Y., Variation of the Asian summer monsoon since the last glacial-interglacial recorded in a stalagmite from southwest China. Quat. Sci. Rev. 234, 106261 (2020). [Google Scholar]
  • 80.Yin Q. Z., Wu Z. P., Berger A., Goosse H., Hodell D., Insolation triggered abrupt weakening of Atlantic circulation at the end of interglacials. Science 373, 1035–1040 (2021). [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary Text

Figs. S1 to S10

Legends for datasets S1 and S2

References

sciadv.aea5042_sm.pdf (5.8MB, pdf)

Datasets S1 and S2

Data Availability Statement

All data are available in the public domain as follows: (i) globally gridded SSTs from ERSST version 5, available at www.ncei.noaa.gov/products/extended-reconstructed-sst; (ii) globally gridded precipitation from the GPCP monthly analysis version 2.3, available at https://climatedataguide.ucar.edu/climate-data/gpcp-monthly-global-precipitation-climatology-project; (iii) globally gridded monthly averaged data on pressure levels from 1940 to present at 850 hPa from the ERA5, available at https://cds.climate.copernicus.eu/datasets/reanalysis-era5-pressure-levels?tab=overview; (iv) globally gridded monthly IsoGSM2, available at http://isotope.iis.u-tokyo.ac.jp/~kei/tmp/isogsm2; (v) the idealized precession simulation EC-Earth-2-2, available at https://zenodo.org/records/5054419; (vi) temperature, precipitation, and δ18Op from GNIP Havana (CPHR) station at 2008 CE and δ18Op data from GNIP stations in the Antilles of the West Indies are available at https://www.iaea.org/services/networks/gnip; and (vii) globally gridded u and v winds at 850 hPa from https://psl.noaa.gov/data/gridded/data.ncep.reanalysis.html. All links were deposited on 25 September 2025. All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials.


Articles from Science Advances are provided here courtesy of American Association for the Advancement of Science

RESOURCES