Hostname: page-component-78c5997874-ndw9j Total loading time: 0 Render date: 2024-10-27T19:23:34.859Z Has data issue: false hasContentIssue false

Precessional hydroclimatic synchronicity changes in the Indo-Pacific Warm Pool driven by the intertropical convergence zone over the past 450 kyr

Published online by Cambridge University Press:  22 October 2024

Fang Qian
Affiliation:
Key Laboratory of Ocean Observation and Forecasting, Key Laboratory of Marine Geology and Environment, Institute of Oceanology, Chinese Academy of Sciences, Qingdao, China Department of Earth and Environmental Sciences, Tulane University, New Orleans, LA, USA Department of Geology and Geophysics, Woods Hole Oceanographic Institution, Woods Hole, MA, USA
Fengming Chang*
Affiliation:
Key Laboratory of Ocean Observation and Forecasting, Key Laboratory of Marine Geology and Environment, Institute of Oceanology, Chinese Academy of Sciences, Qingdao, China Laboratory for Marine Geology, Qingdao Marine Science and Technology Center, Qingdao, China University of Chinese Academy of Sciences, Beijing, China
Dirk Nürnberg
Affiliation:
GEOMAR Helmholtz Center for Ocean Research Kiel, Kiel, Germany
Shuai Zhang
Affiliation:
College of Oceanography, Hohai University, Nanjing, China
Yi Wang
Affiliation:
Department of Earth and Environmental Sciences, Tulane University, New Orleans, LA, USA
Junru Zhang
Affiliation:
Key Laboratory of Ocean Observation and Forecasting, Key Laboratory of Marine Geology and Environment, Institute of Oceanology, Chinese Academy of Sciences, Qingdao, China University of Chinese Academy of Sciences, Beijing, China
Luyao Tang
Affiliation:
Key Laboratory of Ocean Observation and Forecasting, Key Laboratory of Marine Geology and Environment, Institute of Oceanology, Chinese Academy of Sciences, Qingdao, China University of Chinese Academy of Sciences, Beijing, China
Tiegang Li*
Affiliation:
Laboratory for Marine Geology, Qingdao Marine Science and Technology Center, Qingdao, China Key Laboratory of Marine Geology and Metallogeny, First Institute of Oceanography, MNR, Qingdao, China
*
Corresponding authors: Fengming Chang; Email: [email protected]; Tiegang Li; Email: [email protected]
Corresponding authors: Fengming Chang; Email: [email protected]; Tiegang Li; Email: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

The Indo-Pacific Warm Pool (IPWP) significantly influences the global hydrological cycle through its impact on atmospheric-oceanic circulation. However, gaining a comprehensive understanding of the hydrologic climate dynamics within the IPWP and its broader effects on the global climate have been hindered by spatial and temporal limitations in paleoclimate records on orbital timescales. In this study, we reconstructed precipitation records (approximated from δ18Osw-ivc) over the past 450 kyr, based on planktonic foraminiferal Mg/Ca and δ18O data obtained from International Ocean Discovery Program Site U1486 in the western tropical Pacific. The δ18Osw-ivc record revealed a generally consistent pattern with precession variations over the past 450 kyr, closely corresponding to changes in boreal summer insolation at the equator. The δ18Osw-ivc record displayed an anti-phased relationship with Chinese speleothem δ18O records on the precession band, with lower precipitation in the western tropical Pacific and higher precipitation in the East Asia summer monsoon region during periods of high Northern Hemisphere summer insolation. This anti-phased correlation primarily resulted from the north-south migration of the Intertropical Convergence Zone (ITCZ), influenced by the interhemispheric insolation contrast. By considering additional δ18Osw-ivc records from various locations within the IPWP region, we identified synchronous precipitation changes within the IPWP on the precession band. The synchronization of precipitation on both margins of the ITCZ’s seasonal range and differences between central and marginal regions of the ITCZ within the IPWP revealed the expansion and contraction of the ITCZ on precession band.

Type
Original Article
Creative Commons
Creative Common License - CCCreative Common License - BYCreative Common License - NCCreative Common License - ND
This is an Open Access article, distributed under the terms of the Creative Commons Attribution-NonCommercial-NoDerivatives licence (http://creativecommons.org/licenses/by-nc-nd/4.0/), which permits non-commercial re-use, distribution, and reproduction in any medium, provided that no alterations are made and the original article is properly cited. The written permission of Cambridge University Press must be obtained prior to any commercial use and/or adaptation of the article.
Copyright
© The Author(s), 2024. Published by Cambridge University Press

1. Introduction

The Indo-Pacific Warm Pool (IPWP), acknowledged as possessing the world’s largest tropical warm water mass, is characterized by sea surface temperatures surpassing 28°C (Clement et al. Reference Clement, Seager and Murtugudde2005; Kim et al. Reference Kim, Yu and Lu2012; De Deckker, Reference De Deckker2016) and plays a crucial role in atmospheric deep convection (Russell et al. Reference Russell, Vogel, Konecky, Bijaksana, Huang, Melles, Wattrus, Costa and King2014). Variations in sea surface temperatures, rainfall patterns and oceanic circulation within the IPWP significantly impact the distribution of heat and moisture, thus influencing global atmospheric circulation patterns (Cane, Reference Cane1998; Hoerling et al. Reference Hoerling, Hurrell and Xu2001; Yan et al. Reference Yan, Ho, Zheng and Klemas1992; Qian et al. Reference Qian, Chang, Li, Li, Sun and Zhang2023). Modern observations highlight the substantial impact of the seasonal latitudinal shifts of the Intertropical Convergence Zone (ITCZ) and its associated monsoon system, as well as the El Niño-Southern Oscillation (ENSO) variability, on shaping hydrologic characteristics in the IPWP region (Leech et al. Reference Leech, Lynch-Stieglitz and Zhang2013; Schneider et al. Reference Schneider, Bischoff and Haug2014). Therefore, studying hydrological climate variations in the IPWP is crucial for unraveling its intricate role in regional and global climate systems.

On geological timescales, the impact and mechanisms of ITCZ variations on the regional precipitation changes within the IPWP present conflicting findings. As a narrow band of convective winds and intense precipitation located near the equator, the ITCZ regulates the distribution of precipitation in the middle and low latitudes by seasonally migrating in response to interhemispheric temperature gradients (Schneider et al. Reference Schneider, Bischoff and Haug2014; Denniston et al. Reference Denniston, Ummenhofer, Wanamaker, Lachniet, Villarini, Asmerom, Polyak, Passaro, Cugley, Woods and Humphreys2016). With the southward migration of the ITCZ, precipitation decreases in the Northern Hemisphere summer monsoon area, while increasing in the Southern Hemisphere summer monsoon area, and vice versa (Schneider et al. Reference Schneider, Bischoff and Haug2014). Several studies propose that tropical precipitation changes were primarily linked to the precession cycles in Earth’s orbit (Clement et al. Reference Clement, Hall and Broccoli2004; Wang et al. Reference Wang, Cheng, Edwards, Kong, Shao, Chen, Wu, Jiang, Wang and An2008; Merlis et al. Reference Merlis, Schneider, Bordoni and Eisenman2013; Jalihal et al. Reference Jalihal, Bosmans, Srinivasan and Chakraborty2019). Accordingly, the interhemispheric anti-phasing precipitation pattern is shaped by contrasting insolation between the Northern and Southern Hemispheres, causing latitudinal ITCZ migrations (Berger, Reference Berger1978; Kutzbach et al. Reference Kutzbach, Liu, Liu and Chen2008). Findings from stalagmite δ18O records in both hemispheres strongly support this hypothesis (e.g., Wang et al. Reference Wang, Cheng, Edwards, Kong, Shao, Chen, Wu, Jiang, Wang and An2008; Cheng et al. Reference Cheng, Edwards, Sinha, Spotl, Yi, Chen, Kelly, Kathayat, Wang, Li, Kong, Wang, Ning and Zhang2016). Reconstructions of precipitation, runoff and sea-surface salinity (SSS) from the IPWP further reveal notable responses to precession-induced alterations (e.g., Kissel et al. Reference Kissel, Laj, Kienast, Bolliet, Holbourn, Hill, Kuhnt and Braconnot2010; Tachikawa et al. Reference Tachikawa, Cartapanis, Vidal, Beaufort, Barlyaeva and Bard2011; Dang et al. Reference Dang, Jian, Kissel and Bassinot2015; Jian et al. Reference Jian, Wang, Dang, Mohtadi, Rosenthal, Lea, Liu, Jin, Ye, Kuhnt and Wang2022). Notably, Earth’s obliquity also appears to be important in driving the latitudinal migration of the ITCZ, similarly resulting in anti-phase variations in interhemispheric precipitation in the IPWP (Liu et al. Reference Liu, Lo, Shi, Wei, Chou, Chen, Chuang, Wu, Mii, Peng, Amakawa, Burr, Lee, DeLong, Elderfield and Shen2015; Zhang et al. Reference Zhang, Xu, Holbourn, Kuhnt, Beil, Li, Xiong, Dang, Yan, Pei, Ran and Wu2020, Reference Zhang, Xu, Holbourn, Kuhnt, Xiong and Li2022). Liu et al. (Reference Liu, Lo, Shi, Wei, Chou, Chen, Chuang, Wu, Mii, Peng, Amakawa, Burr, Lee, DeLong, Elderfield and Shen2015) suggest that obliquity forcing may have a greater impact on the western Pacific ITCZ migration, based on planktonic foraminifera REE/Ca records. Therefore, the reconstruction of precipitation records on longer timescales is of great significance for exploring the effects of diverse orbital forcing on the hydroclimate variability.

In contrast to the observed anti-phase precipitation patterns between hemispheres on different timescales, an increasing number of studies demonstrated synchronized precipitation variations between hemispheres in tropical regions on (sub-)millennial timescales (e.g., Konecky et al. Reference Konecky, Russell, Rodysill, Vuille, Bijaksana and Huang2013; Yan et al. Reference Yan, Wei, Soon, An, Zhou, Liu, Wang and Carter2015; Denniston et al. Reference Denniston, Ummenhofer, Wanamaker, Lachniet, Villarini, Asmerom, Polyak, Passaro, Cugley, Woods and Humphreys2016; Scroxton et al. Reference Scroxton, Burns, McGee, Hardt, Godfrey, Ranivoharimanana and Faina2017; Yang et al. Reference Yang, Zhang, Yi, Zhong, Lu, Wan, Du and Xiang2023). Such synchronized pattern would contradict the traditional migration theory of the ITCZ. The in-phase precipitation changes can be well explained by the expansion and contraction of the latitudinal extent of the ITCZ (Konecky et al. Reference Konecky, Russell, Rodysill, Vuille, Bijaksana and Huang2013; Yan et al. Reference Yan, Wei, Soon, An, Zhou, Liu, Wang and Carter2015; Denniston et al. Reference Denniston, Ummenhofer, Wanamaker, Lachniet, Villarini, Asmerom, Polyak, Passaro, Cugley, Woods and Humphreys2016; Scroxton et al. Reference Scroxton, Burns, McGee, Hardt, Godfrey, Ranivoharimanana and Faina2017; Yang et al. Reference Yang, Zhang, Yi, Zhong, Lu, Wan, Du and Xiang2023). During times of ITCZ expansion, precipitation intensifies synchronously at both northern and southern edges, while precipitation declines simultaneously at both edges when the ITCZ contracts (Yuan et al. Reference Yuan, Chiang, Liu, Bijaksana, He, Jiang, Imran, Wicaksono and Wang2023). The expansion and contraction of the ITCZ is likely triggered by various interacting factors, such as symmetrical changes in solar irradiance (Yan et al. Reference Yan, Wei, Soon, An, Zhou, Liu, Wang and Carter2015; Scroxton et al. Reference Scroxton, Burns, McGee, Hardt, Godfrey, Ranivoharimanana and Faina2017), strengthening Walker circulation (Konecky et al. Reference Konecky, Russell, Rodysill, Vuille, Bijaksana and Huang2013) and changes in the meridional atmospheric circulation (Denniston et al. Reference Denniston, Ummenhofer, Wanamaker, Lachniet, Villarini, Asmerom, Polyak, Passaro, Cugley, Woods and Humphreys2016). Further studies, hence, are warranted to explore the impact of spatial ITCZ variations on alterations in precipitation patterns between hemispheres.

As the region with the highest water vapor exchange globally (Pierrehumbert, Reference Pierrehumbert2000), the IPWP is ideal to provide insight into the impact of ITCZ migration and expansion/contraction on local precipitation. Recently, Xiong et al. (Reference Xiong, Zhai, Algeo, Lu, Li, Meyer, Jiang, Zhang, Qin, Gong and Wang Zand Jia2022) and Yu et al. (Reference Yu, Tang, Colin, Wilson, Zhou, Song, Chang, Zhang, Bassinot and Wan2023) examined the spatial pattern of precipitation variations across the IPWP on millennial scales, highlighting the importance of the ITCZ on the precipitation patterns in different regions of the IPWP. The observed changes in precipitation during the Last Glacial Maximum (LGM) point to an overall drying trend across the IPWP, caused by both the zonal shifts of ENSO and the meridional migration of ITCZ (Xiong et al. Reference Xiong, Zhai, Algeo, Lu, Li, Meyer, Jiang, Zhang, Qin, Gong and Wang Zand Jia2022). Yu et al. (Reference Yu, Tang, Colin, Wilson, Zhou, Song, Chang, Zhang, Bassinot and Wan2023) identified noticeably spatial heterogeneity of precipitation within the IPWP over the past 40 kyr, attributed to the predominant effects of the ENSO-like system in the western Pacific and the migration of the ITCZ in the Indian Ocean. Unfortunately, studies on past ITCZ expansion/contraction changes are not only scarce but also primarily confined to the last few millennia for the IPWP region (e.g., Denniston et al. Reference Denniston, Ummenhofer, Wanamaker, Lachniet, Villarini, Asmerom, Polyak, Passaro, Cugley, Woods and Humphreys2016; Konecky et al. Reference Konecky, Russell, Rodysill, Vuille, Bijaksana and Huang2013; Scroxton et al. Reference Scroxton, Burns, McGee, Hardt, Godfrey, Ranivoharimanana and Faina2017; Yan et al. Reference Yan, Wei, Soon, An, Zhou, Liu, Wang and Carter2015; Yang et al. Reference Yang, Zhang, Yi, Zhong, Lu, Wan, Du and Xiang2023), significantly limiting our understanding of the impact of ITCZ variability on spatial patterns of precipitation variations in IPWP.

To determine the temporal and spatial variability of the ITCZ on orbital timescales, we attempted to reconstruct the precipitation patterns within the IPWP region. Past precipitation, approximated from the residual oxygen isotope compositions of seawater, was deduced from combined δ18O and Mg/Ca compositions of sea surface-dwelling planktonic foraminifera from International Ocean Discovery Program (IODP) Site U1486 within the central IPWP (Bismarck Sea). By integrating published ice-volume corrected sea surface salinity records (δ18Osw-ivc) from the IPWP region, we assessed the spatial and temporal patterns of precipitation across the IPWP over the past 450 kyr and explored its driving mechanisms on the precession band.

2. Materials and methods

2.a. Location of site U1486 and modern climate

IODP Expedition 363 Site U1486 (02º22.34ʹS, 144º36.08ʹE) was drilled at a water depth of 1332 m in the Bismarck Sea, located ∼150 km away from the mouth of the Sepik River in Papua New Guinea (Fig. 1a; Rosenthal et al. Reference Rosenthal, Holbourn and Kulhanek2018). This study focused on samples from Site U1486, collected at 10-cm intervals, covering a depth range from the core-top down to 31 m Core Composite Depth Below Sea Floor. The sediment lithology in this section remained consistently homogenous, primarily composed of fine-grained, grayish-white sediment rich in planktonic microfossils. No volcanic ash layers or signs of disturbance were observed.

Figure 1. The Indo-Pacific Warm Pool and regional circulation patterns, alongside the modern climatology of the study area. (a) Map of the IPWP displaying modern annual sea surface temperatures (SST at 0 m; World Ocean Atlas 2013 dataset; Locarnini et al. Reference Locarnini, Mishonov, Antonov, Boyer, Garcia, Baranova, Zweng, Paver, Reagan, Johnson, Hamilton and Seidov2013) of the studied Site U1486 (yellow star) and other relevant sites discussed in the study (white circles) is shown. Bright yellow shading lines denote the mean positions of the Intertropical Convergence Zone (ITCZ) in July and January, respectively (Lutgens & Tarbuck, Reference Lutgens and Tarbuck2001). Black dotted lines represent annual mean 28°C SST isotherms. Key ocean currents include the North Equatorial Current (NEC), North Equatorial Counter Current (NECC), South Equatorial Current (SEC), Kuroshio Current (KC) and Mindanao Current (MC). (b) The Southern Oscillation Index (SOI, 5-point running average) between 1980 and 2020 (https://www.cpc.ncep.noaa.gov/data/indices/soi) together with the monthly rainfall anomaly over the Site U1486 source area (2.5° spatial resolution centered on 2°S, 144°E, 11-point running average, https://climatedataguide.ucar.edu/climate-data/gpcp-monthly-global-precipitation-climatology-project). (c) Depiction of mean rainfall rate and the 1000 hPa wind field in January (left) and July (right). Rainfall data sourced from https://psl.noaa.gov/data/gridded/data.cmap.html, while land wind data obtained from http://iridl.ldeo.columbia.edu/expert/ds:/SOURCES/.NOAA/.NCEP-NCAR/.CDAS-1.

The hydrological cycle in the studied area is influenced both by the Asian-Australian monsoon and ENSO (Webster et al. Reference Webster, Magana, Palmer, Shukla, Tomas, Yanai and Yasunari1998; Wang et al. Reference Wang, Wu and Li2003). During the boreal summer, southeasterly trade winds prevail, with the ITCZ shifting northward, thereby reducing the precipitation rate in the study area (Fig. 1a, c). Conversely, northwesterly winds dominate during the boreal winter, leading to a southward shift of the ITCZ and an increase in precipitation (Fig. 1a, c). On inter-annual timescales, precipitation is closely linked to ENSO events (Fig. 1b). Unusual dry (wet) conditions occur in the Papua New Guinea region during El Niño (La Niña) phases (Dai & Wigley, Reference Dai and Wigley2000).

2.b. Stable oxygen isotope and Mg/Ca analyses

Stable oxygen isotope (δ18O) and Mg/Ca analyses were conducted on the surface-dwelling planktonic foraminiferal species Globigerinoides ruber (white) extracted from 335 samples collected at 10-cm intervals downcore. Approximately 50 shells of G. ruber were selected from the 250–355 μm size fraction. Subsequently, these foraminiferal shells were gently crushed into several fragments.

For stratigraphical purposes, δ18O was measured on the benthic foraminifer Cibicidoides wuellerstorfi. Approximately 3–5 individuals were picked from both the 300–355 μm and the 355–500 μm size fractions.

For stable isotope analysis, all fragments from the benthic samples and one-third of the fragments from the planktonic foraminifer G. ruber were utilized, while the remaining two-thirds of the G. ruber fragments were allocated for Mg/Ca analysis. The isotope analysis was performed using a GV IsoPrime mass spectrometer at the Key Laboratory of Marine Geology and Environment, Institute of Oceanography, Chinese Academy of Sciences, Qingdao. Results are presented in δ18O notation (‰ relative to the Vienna Pee Dee Belemnite standard), calibrated to the National Bureau of Standards 18 standard, analyzed approximately every 10 samples, with a long-term standard deviation of 0.06‰.

For Mg/Ca analysis, the G. ruber underwent pretreatment and cleaning procedures following the methods described by Barker et al. (Reference Barker, Greaves and Elderfield2003), with the exclusion of a reductive step. The shells were sequentially cleaned using Milli-Q water rinses, methanol rinses and an oxidizing solution. Organic matter was eliminated through oxidation using a heated 1% NaOH-buffered H2O2 solution. Samples were dissolved in 0.075-M QD HNO3 and analyzed using an inductively coupled plasma–optical emission spectrometer (ICP-OES) at the same laboratory. The average standard deviation for replicate measurements was approximately 0.52%. The contamination assessment included Fe/Ca, Mn/Ca and Al/Ca ratios, with most samples registering values below detection limits (Fig. S1). Furthermore, according to Barker et al. (Reference Barker, Greaves and Elderfield2003), the presence of detrital contamination affecting the measured Mg/Ca ratios can be inferred by evaluating the correlation between estimated Mg/Ca and Fe/Ca or Al/Ca data. No significant correlation was found between Fe/Ca and Mg/Ca (R2 = 0.006), Mn/Ca and Mg/Ca (R2 = 0.02), or Al/Ca and Mg/Ca (R2 = 0.004), ensuring the reliability of the Mg/Ca measurements (Fig. S1).

2.c. Estimates of sea-surface temperature and seawater δ18O

The planktonic foraminifer G. ruber calcifies within the depths of 30–95 m in the mixed layer of the Western Pacific Warm Pool (WPWP) (Rippert et al. Reference Rippert, Nürnberg, Raddatz, Maier, Hathorne, Bijma and Tiedemann2016; Raddatz et al. Reference Raddatz, Nürnberg, Tiedemann and Rippert2017). By comparing the reconstructed sea-surface temperature (SSTMg/Ca) based on the Mg/Ca ratio in G. ruber shells using available calibration equations (Fig. S2; Table S1), we selected the equation (1) established in the western Pacific by Dekens et al. (Reference Dekens, Lea, Pak and Spero2002). This decision was based on the fact that its constructed average core-top SSTMg/Ca value (∼29.3°C) aligns closely with the modern SST (∼29.1°C; Locarnini et al. Reference Locarnini, Mishonov, Antonov, Boyer, Garcia, Baranova, Zweng, Paver, Reagan, Johnson, Hamilton and Seidov2013).

(1) $$\matrix{ {{\rm{SST}}(^\circ {\rm{C}}) = \;{\rm{ln}}\;\left[ {{\rm{Mg}}/{\rm{Ca}}\;\left( {{\rm{mmol}}/{\rm{mol}}} \right)/0.38} \right]/0.09\;} \hfill \cr {\quad \quad \quad \quad \quad \quad + 0.61\;\left( {{\rm{core}}\;{\rm{depth}}\;{\rm{km}}} \right) + 1.6} \hfill \cr } $$

Calcite dissolution might lead to lowered Mg/Ca-temperature estimates due to the selective removal of Mg-ions from the foraminiferal calcite (e.g., Nürnberg et al. Reference Nürnberg, Bijma and Hemleben1996; Dekens et al. Reference Dekens, Lea, Pak and Spero2002; Regenberg et al. Reference Regenberg, Nürnberg, Steph, Groeneveld, Garbe-Schönberg, Tiedemann and Dullo2006). Equation (1) incorporates a depth correction to account for dissolution effects on Mg/Ca. As indicated by Regenberg et al. (Reference Regenberg, Regenberg, Garbe-Schönberg and Lea2014), foraminiferal Mg/Ca in the study area is notably affected at depths below 2500 m. Given the limited impact of dissolution at our specific location (1332 m), we decided not to apply the dissolution correction in our Mg/Ca data. Moreover, recent calibrations derived from sediment traps and modern core tops in WPWP provide confidence that the dissolution effect is insignificant in our study area (Hollstein et al. Reference Hollstein, Mohtadi, Rosenthal, Moffa Sanchez, Oppo, Martínez Méndez, Steinke and Hebbeln2017).

The error in SSTMg/Ca reconstructions was assessed by propagating the uncertainties arising from both the Mg/Ca measurements and the Mg/Ca temperature calibration, following Mohtadi et al. (Reference Mohtadi, Prange, Oppo, De Pol-Holz, Merkel, Zhang, Steinke and Luckge2014). On average, the resulting errors are approximately ±0.97°C.

The oxygen isotope composition of seawater (δ18Osw) and SSS exhibit a linear relationship in the western Pacific (e.g., Fairbanks et al. Reference Fairbanks, Evans, Rubenstone, Mortlock, Broad, Moore and Charles1997; Gibbons et al. Reference Gibbons, Oppo, Mohtadi, Rosenthal, Cheng, Liu and Linsley2014). As a result, the residual δ18O of surface seawater (δ18Osw-ivc), which corrects for the influence of continental ice volume, has been widely utilized in the reconstruction of past regional hydrological changes. We derived δ18Osw using Equation (2) established by Bemis et al. (Reference Bemis, Spero, Bijma and Lea1998), adjusting values to the Vienna Standard Mean Ocean Water standard by adding 0.27‰ (Hut, Reference Hut1987). The estimated average core-top δ18Osw value (∼0.25‰) closely matches the modern seawater δ18O value at the core site (0.22‰; LeGrande & Schmidt, Reference LeGrande and Schmidt2006). Additionally, we compared the estimated δ18Osw results derived from different SSTMg/Ca calibration equations (Fig. S2; Table S1), validating the reliability of the calibration equations. Furthermore, we corrected δ18Osw for global ice volume changes based on the studies by Bintanja et al. (Reference Bintanja, van de Wal and Oerlemans2005) and Spratt & Lisiecki (Reference Spratt and Lisiecki2016), respectively. As shown in Fig. S2, the two estimated δ18Osw-ivc results exhibit a high level of consistency. Consequently, the estimated δ18Osw-ivc derived from simulated global marine isotopes from Bintanja et al. (Reference Bintanja, van de Wal and Oerlemans2005) is reliable and feasible.

(2) $${{\rm{\delta }}^{18}}{{\rm{O}}_{{\rm{sw}}}} = \left[ {{\rm{SS}}{{\rm{T}}_{{\rm{Mg}}/{\rm{Ca}}}}\left( {^\circ {\rm{C}}} \right) - 16.5 + 4.8 \times {{\rm{\delta }}^{18}}{{\rm{O}}_{G.\;ruber}}} \right]/4.8 + 0.27$$

We calculated uncertainties in δ18Osw-ivc using the method described by Mohtadi et al. (Reference Mohtadi, Prange, Oppo, De Pol-Holz, Merkel, Zhang, Steinke and Luckge2014), accounting for uncertainties from Mg/Ca and δ18O measurements, temperature equations and global ice volume removal (Bintanja et al. Reference Bintanja, van de Wal and Oerlemans2005). The resulting uncertainties averaged 0.21‰ for δ18Osw-ivc.

2.d. Chronology

The age model for Site U1486 was constructed based on the δ18O record obtained from the benthic foraminifer Cibicidoides wuellerstorfi18O C. wuellerstorfi ). The δ18O C. wuellerstorfi values ranged from 1.68‰ to 3.85‰ over the past 450 kyr BP, revealing clear glacial-interglacial cycles (Fig. 2a). Comparison of δ18O C. wuellerstorfi values with the global benthic foraminiferal δ18O stack LR04 (Lisiecki & Raymo, Reference Lisiecki and Raymo2005) identified 15 tie points for age control (Fig. 2a). The age model for the entire core was calculated using a MATLAB linear interpolation optimization method. Site U1486 provides a continuous record spanning the past 450 kyr, encompassing five glacial-interglacial cycles (Fig. 2a). The average sedimentation rate over this period was approximately 7.17 cm/kyr the past 450 kyr (Fig. 2b). With a sampling interval of 10 cm, the average sample resolution is approximately 1.3 kyr.

Figure 2. Age model of Site U1486. (a) Benthic foraminifer C. wuellerstorfi δ18O (blue line) and the global benthic foraminiferal δ18O stack LR04 (Lisiecki & Raymo, Reference Lisiecki and Raymo2005) (red line). Black triangles indicate tie points for δ18O C. wuellerstorfi . (b) Depth-age model (green line) and sedimentation rates (black line) of sediment Site U1486. Gray vertical shadings indicate even-numbered Marine Isotope Stages (MIS).

3. Results

The oxygen isotope values of G. ruber18O G. ruber ) exhibit a range from −3.18‰ to −0.93‰ over the past 450 kyr, displaying distinct glacial-interglacial cycles (Fig. 3a). Generally, δ18O G. ruber values are higher/heavier during glacial periods (MIS 2, 4, 6, 8, 10 and 12) and lower/lighter during interglacial stages. The highest δ18O G. ruber value (−0.93‰) is observed at ∼433.4 kyr (MIS 12), while the lowest δ18O G. ruber value (∼−3.18‰) is found at the top of the core.

Figure 3. Planktonic foraminiferal proxy records of Site U1486. (a) G. ruber δ18O, (b) G. ruber Mg/Ca-based SST, (c) the ice volume corrected-seawater δ18O (δ18Osw-ivc) and (d) C. wuellerstorfi δ18O (black), superimposed is the global ice volume δ18Osw (blue; Bintanja et al. Reference Bintanja, van de Wal and Oerlemans2005). Gray vertical shadings indicate even-numbered marine isotope stages (MIS), while colored shadings for individual records represent their respective error ranges (see Sections 2.2 and 2.3). (e) Spectral (left) and wavelet transform (right) analyses of the δ18Osw-ivc record of Site U1486. The green line indicates the precession cycle.

G. ruber Mg/Ca varies from 3.22 to 5.24 mmol/mol, which yields sea-surface temperature (SSTMg/Ca) at Site U1486 spanning a range of 25.3 to 30.8 °C, with an average of approximately 27.7 °C. As shown in Figure 3b, the SSTMg/Ca record shows pronounced glacial-interglacial variations over the past 450 kyr, with higher temperatures occurring during interglacial stages and lower temperatures during peak glacial stages.

Modern seawater δ18O and SSS display a positive correlation, with higher δ18Osw-ivc values corresponding to increased salinity in surface seawater, while lower δ18Osw-ivc values indicate fresher seawater (Fairbanks et al. Reference Fairbanks, Evans, Rubenstone, Mortlock, Broad, Moore and Charles1997; Rosenthal et al. Reference Rosenthal, Oppo and Linsley2003; Stott et al. Reference Stott, Cannariato, Thunell, Haug, Koutavas and Lund2004; Gibbons et al. Reference Gibbons, Oppo, Mohtadi, Rosenthal, Cheng, Liu and Linsley2014). Model simulations also support the assumption that the tropical Pacific has maintained a consistent relationship between δ18Osw-ivc and SSS (Holloway et al. Reference Holloway, Sime, Singarayer, Tindall and Valdes2016). In comparison with δ18O G. ruber and SSTMg/Ca, δ18Osw-ivc does not exhibit obvious glacial-interglacial cycles, ranging from −0.66‰ to 1.20‰ (Fig. 3c). However, during MIS 2, MIS 6, MIS 10 and MIS 12, δ18Osw-ivc is relatively lower compared to the interglacial stages, with the highest value observed near the MIS 10/9 transition (Fig. 3c). During deglacial episodes of rapid SSTMg/Ca warming (e.g., MIS 12/11, MIS 10/9, MIS 6/5 and MIS 2/1), sea-surface δ18Osw-ivc values simultaneously increase, implying a change to more saline conditions at sea surface (Fig. 3b, c).

4. Discussion

4.a. Using site U1486 δ18Osw-ivc as a proxy to reconstruct local precipitation variability over the past 450 kyr

δ18Osw serves as a common proxy for SSS, which is a crucial parameter in the hydrological cycle (Stott et al. Reference Stott, Cannariato, Thunell, Haug, Koutavas and Lund2004; Leech et al. Reference Leech, Lynch-Stieglitz and Zhang2013). Although δ18Osw is influenced by multiple factors, including precipitation, evaporation, upwelling and meridional and zonal currents (Conroy et al. Reference Conroy, Thompson, Cobb, Noone, Rea and LeGrande2017), the existence of a linear correlation between δ18Osw and SSS is due to both being primarily influenced by evaporation and precipitation (Fairbanks et al. Reference Fairbanks, Evans, Rubenstone, Mortlock, Broad, Moore and Charles1997; LeGrande & Schmidt, Reference LeGrande and Schmidt2006). Thus, the δ18Osw variability is often interpreted in the context of hydroclimate variability (Fairbanks et al. Reference Fairbanks, Evans, Rubenstone, Mortlock, Broad, Moore and Charles1997; LeGrande & Schmidt, Reference LeGrande and Schmidt2006). Additionally, Conroy et al. (Reference Conroy, Thompson, Cobb, Noone, Rea and LeGrande2017) found that the δ18Osw-salinity linear relationship appears stronger in the western tropical Pacific based on both observational data and results from isotope-enabled climate model simulations. Therefore, δ18Osw primarily reflects changes in the precipitation-evaporation balance. Given the minimal fluctuations in modern SST in the western Pacific (Kim et al. Reference Kim, Yu and Lu2012), δ18Osw likely reflects local precipitation patterns (Gibbons et al. Reference Gibbons, Oppo, Mohtadi, Rosenthal, Cheng, Liu and Linsley2014). Recent studies by Yang et al. (Reference Yang, Zhou, Yan, Ma, Cheng, Hu, Shen, Zhao, Hou, Lu and Liu2024) analyzing modern reanalysis data in the South China Sea revealed an inverse correlation between regional SSS and precipitation rate. Additionally, Conroy et al. (Reference Conroy, Thompson, Cobb, Noone, Rea and LeGrande2017) identified a weak relationship between SST and SSS in the western Pacific based on modern data.

For the reconstruction of past local precipitation changes in the western Pacific, we consider the ice-volume corrected δ18Osw-ivc record from Site U1486. Positive deviations in δ18Osw-ivc values indicating more saline conditions at sea surface, suggest reduced local precipitation in the western Pacific, while negative values point to intensified local precipitation (Rosenthal et al. Reference Rosenthal, Oppo and Linsley2003; Stott et al. Reference Stott, Cannariato, Thunell, Haug, Koutavas and Lund2004). The Site U1486 δ18Osw-ivc record displays relatively muted glacial-interglacial cycles over the past 450 kyr, which fluctuate between −0.66 ‰ and 1.20 ‰ and differ substantially from the SSTMg/Ca pattern (Fig. 3c). Moreover, the variation in Site U1486 δ18Osw-ivc is similar to the change in the ln(Ti/Ca) record obtained from the adjacent core MD05-2920 (Fig. 4i; Tachikawa et al. Reference Tachikawa, Cartapanis, Vidal, Beaufort, Barlyaeva and Bard2011, Reference Tachikawa, Timmermann, Vidal, Sonzogni and Timm2014). The ln(Ti/Ca) values were utilized as an indicator of rainfall over the drainage basin of the Sepik River in New Guinea (Tachikawa et al. Reference Tachikawa, Cartapanis, Vidal, Beaufort, Barlyaeva and Bard2011, Reference Tachikawa, Timmermann, Vidal, Sonzogni and Timm2014). Cross-spectral analysis reveals a significant coherence between Site U1486 δ18Osw-ivc and ln(Ti/Ca) from core MD05-2920 on the precession band (Fig. S3), supporting the dominance of precession cycles in controlling precipitation variations in the study area.

Figure 4. The orbital-scale precipitation pattern in the Indo-Pacific Warm Pool over the past 450 kyr. (a) The variation in precession (Laskar et al. Reference Laskar, Robutel, Joutel, Gastineau, Correia and Levrard2004). (b) GISS_ModelE2-R simulated δ18Osw-ivc records of the IPWP (Jian et al. Reference Jian, Wang, Dang, Mohtadi, Rosenthal, Lea, Liu, Jin, Ye, Kuhnt and Wang2022). (c, d, e) The δ18Osw-ivc records of cores MD98-2162, SO18480-3 and U1483 in the eastern tropical Indian Ocean (Jian et al. Reference Jian, Wang, Dang, Mohtadi, Rosenthal, Lea, Liu, Jin, Ye, Kuhnt and Wang2022; Zhang et al. Reference Zhang, Xu, Holbourn, Kuhnt, Xiong and Li2022). (f, g) The δ18Osw-ivc records of cores MD06-3047B and MD06-3067 in the northern sector of western equatorial Pacific (Bolliet et al. Bolliet et al., Reference Bolliet, Holbourn, Kuhnt, Laj, Kissel, Beaufort, Kienast, Andersen and Garbe-Schönberg2011; Jia et al. Jia et al., Reference Jia, Li, Xiong, Steinke, Jiang, Chang and Qin2018). (h) The δ18Osw-ivc records of core MD05-2925 (Lo et al. Reference Lo, Shen, Zeeden, Tsai, Yin, Yang, Chang, Su, Mii, Chuang and Chen2022). (i) The ln(Ti/Ca) record of core MD05-2920 (Tachikawa et al. Reference Tachikawa, Timmermann, Vidal, Sonzogni and Timm2014). (j) The δ18Osw-ivc record of Site U1486, superimposed by the July–August insolation at the equator (black) (Laskar et al. Reference Laskar, Robutel, Joutel, Gastineau, Correia and Levrard2004). The overlain, bold, dark-colored lines represent the precession band-pass filtered results. Band-pass filter with a central frequency of 0.043 kyr−1 and a bandwidth of 0.01 kyr−1. Vertical grey bars indicate lower precession values.

To study the underlying mechanisms driving local precipitation changes in the western Pacific over the past 450 kyr, we employed redfit spectral and wavelet analysis using the Past3 software with a Hanning window (Hammer et al. Reference Hammer, Harper and Ryan2001). The spectral analysis reveals that the Site U1486 δ18Osw-ivc record is characterized by prominent 100 kyr eccentricity and 23 kyr precession cycles. Wavelet analyses further shows that the 23 kyr cycle is most prominent during ∼260−180 and ∼370−300 ka BP (Fig. 3e). The phase difference between the maximum in δ18Osw-ivc and the maximum in precession is ∼−148° (Fig. 5h). On the precession band, the Site U1486 δ18Osw-ivc record displays precession band-pass filtered patterns that are mostly in phase with a few exceptions, such as δ18Osw-ivc changes preceding precession shifts during the periods of ∼175−145 and ∼110−75 ka BP (Fig. 4j). The discrepancies between the δ18Osw-ivc and precession may be attributed to terrestrial influences (Hollstein et al. Reference Hollstein, Mohtadi, Kienast, Rosenthal, Groeneveld, Oppo, Southon and Lückge2020), or due to a relatively weakened 23 kyr cycle during these specific time intervals (Fig. 3e). Nonetheless, the δ18Osw-ivc record exhibits an overall consistency with precession variations over the past 450 kyr, where almost every low precession value corresponds to high δ18Osw-ivc values (Fig. 4j). We conclude that the past local precipitation variations at Site U1486 were associated with precessional changes.

Figure 5. Comparing the precipitation records in the IPWP and East Asia with the ENSO-like proxy and the migration of the ITCZ. (a) The variation in precession (Laskar et al. Reference Laskar, Robutel, Joutel, Gastineau, Correia and Levrard2004). (b) The speleothem δ18O records in China (Cheng et al. Reference Cheng, Edwards, Sinha, Spotl, Yi, Chen, Kelly, Kathayat, Wang, Li, Kong, Wang, Ning and Zhang2016), superimposed on the summer insolation at 15°N (Laskar et al. Reference Laskar, Robutel, Joutel, Gastineau, Correia and Levrard2004). (c) The summer interhemispheric tropical insolation gradient, determined by the variation in June insolation between 23° N and 23° S (Laskar et al. Reference Laskar, Robutel, Joutel, Gastineau, Correia and Levrard2004). (d) CESM-simulated Niño3 SST of DJF (Zhang et al. Reference Zhang, Yu, Gong, Wang, Chang, Lohmman, Qi and Li2021). (e) Differences between SST and thermocline water temperature (ΔT) for Site U1486. (f) GISS_ModelE2-R simulated δ18Osw-ivc records of the IPWP (Jian et al. Reference Jian, Wang, Dang, Mohtadi, Rosenthal, Lea, Liu, Jin, Ye, Kuhnt and Wang2022). (g) The δ18Osw-ivc records of Site U1486 superimposed by the July–August insolation at the equator (black) (Laskar et al. Reference Laskar, Robutel, Joutel, Gastineau, Correia and Levrard2004). The records are overlaid by the precession band-pass filtered output (bold, dark-colored lines). Band-pass filter with a central frequency of 0.043 kyr−1 and a bandwidth of 0.01 kyr−1. Vertical grey bars indicate lower precession values (high Northern Hemisphere summer insolation). (h) Phase relationships among U1486 δ18Osw-ivc, simulated IPWP δ18Osw-ivc, speleothem δ18O and the precession parameter, illustrated through coherences (left) and phase angles (right).

It is noteworthy that, although Site U1486 δ18Osw-ivc record and the ln(Ti/Ca) record from adjacent core MD05-2920 exhibit nearly coherent variation on the precession band (Fig. 4i, j), the ln(Ti/Ca) record is characterized by not only precession but also obliquity signals (Tachikawa et al. Reference Tachikawa, Timmermann, Vidal, Sonzogni and Timm2014). The coarse terrigenous fraction represented by Ti varied with the precession cycle, while CaCO3 exhibited a pronounced obliquity band (Tachikawa et al. Reference Tachikawa, Cartapanis, Vidal, Beaufort, Barlyaeva and Bard2011). Additionally, precipitation records based on ln(K/Ca) at Site U1483 (offshore north-western Australia) contain an additional obliquity component (Zhang et al. Reference Zhang, Xu, Holbourn, Kuhnt, Beil, Li, Xiong, Dang, Yan, Pei, Ran and Wu2020). The orbital forcing of elemental ratios is evidently different from the predominant precession control observed in the δ18Osw-ivc records of Site U1486. This discrepancy in dominant cyclicity may suggest that different precipitation proxies respond differently to climatic forcing. The SSS proxy (δ18Osw-ivc), primarily controlled by the precipitation-evaporation balance (Conroy et al. Reference Conroy, Thompson, Cobb, Noone, Rea and LeGrande2017), tends to exhibit stronger precession-driven variability due to its direct linkage with monsoon dynamics and seasonal changes in precipitation (Stott et al. Reference Stott, Cannariato, Thunell, Haug, Koutavas and Lund2004).

Orbital precession is triggered by the Earth’s rotational axis oscillation, leading to a continuous change in the relative positioning of the Earth and the Sun (Berger & Loutre, Reference Berger, Loutre, Duplessy and Spyridakis1994). This alteration modifies the seasonal and meridional distribution of solar radiation across the Earth’s surface with a 23 kyr periodicity (Berger & Loutre, Reference Berger, Loutre, Duplessy and Spyridakis1994; Merlis et al. Reference Merlis, Schneider, Bordoni and Eisenman2013). The equatorial insolation variation is primarily influenced by precession changes on orbital timescales, with an increased summer solar insolation in the equatorial region during low precession periods (Yin et al. Reference Yin, Wu, Berger, Goosse and Hodell2021). Our δ18Osw-ivc record reveals a similar pattern with July–August insolation changes at the equator over the past 450 kyr. The elevated insolation levels corresponded to increased δ18Osw-ivc values, implying reduced local precipitation in Papua New Guinea, and vice versa (Fig. 4j). The core MD05-2920 ln(Ti/Ca) also exhibits significant coherence with insolation at precessional periodicities, consistent with our δ18Osw-ivc record of Site U1486 (Fig. 4i, j). Additionally, modeled precipitation results further support the simultaneous changes in precipitation and local summer insolation in the western Pacific (Kutzbach et al. Reference Kutzbach, Liu, Liu and Chen2008). These findings suggest that variations in local precipitation in the Papua New Guinea region were likely driven by the equatorial insolation changes in line with precessional variations. The precession signal might have influenced the local precipitation patterns through northern and southern hemispheric thermal gradients (Braconnot et al. Reference Braconnot, Marzin, Gregoire, Mosquet and Marti2008) or the temperature contrast between land and sea surfaces (Ruddiman, Reference Ruddiman2008; Merlis et al. Reference Merlis, Schneider, Bordoni and Eisenman2013).

4.b. Precession-driven synchronous precipitation changes in the IPWP over the past 450 kyr

The IPWP as Earth’s largest reservoir of atmospheric water vapor experiences the highest rainfall levels (Cane, Reference Cane1998; Leung et al. Reference Leung, Zhang, Gan, Wang, Qian and Hu2022). Modern mean rainfall data show a central region for rainfall within the IPWP (Fig. 1c). To investigate precipitation variations on orbital timescales across the entire IPWP region, we compared precipitation records from different areas within its central precipitation zone (Fig. 1a). These areas encompass the northern IPWP (cores MD06-3047B and MD06-3067; Bolliet et al. Bolliet et al., Reference Bolliet, Holbourn, Kuhnt, Laj, Kissel, Beaufort, Kienast, Andersen and Garbe-Schönberg2011; Jia et al. Jia et al., Reference Jia, Li, Xiong, Steinke, Jiang, Chang and Qin2018), the southern sector of the IPWP (cores U1486 and MD05-2925; this study and Lo et al. Reference Lo, Shen, Zeeden, Tsai, Yin, Yang, Chang, Su, Mii, Chuang and Chen2022) and the Indian Ocean (cores MD98-2162, SO18480-3 and U1483; Jian et al. Reference Jian, Wang, Dang, Mohtadi, Rosenthal, Lea, Liu, Jin, Ye, Kuhnt and Wang2022; Zhang et al. Reference Zhang, Xu, Holbourn, Kuhnt, Xiong and Li2022). We distinguish the northern and southern divisions of the IPWP based on the equator. Detailed information about the original datasets for these cores is provided in Table S2. Conroy et al. (Reference Conroy, Thompson, Cobb, Noone, Rea and LeGrande2017) identified a linear relationship between δ18Osw and SSS in various IPWP areas across multiple sites. δ18Osw-ivc likely serves as a reliable proxy to reflect past local precipitation variations due to the limited variability of SST in the IPWP (Gibbons et al. Reference Gibbons, Oppo, Mohtadi, Rosenthal, Cheng, Liu and Linsley2014). All δ18Osw-ivc records are similarly based on G. ruber Mg/Ca and δ18O and are considered to reliably reflect local precipitation through time.

All δ18Osw-ivc records from various regions of the western tropical Pacific across the equator suggest overall synchronous precipitation patterns on orbital timescales. The δ18Osw-ivc record of core MD05-2925 (Lo et al. Reference Lo, Shen, Zeeden, Tsai, Yin, Yang, Chang, Su, Mii, Chuang and Chen2022), situated in the southern region of the western equatorial Pacific, shows a fundamental synchrony with variations in boreal summer insolation at the equator driven by precession over the past 450 kyr (Fig. 4h). Similarly, the precipitation record of core MD06-3047B (Jia et al. Jia et al., Reference Jia, Li, Xiong, Steinke, Jiang, Chang and Qin2018), located in the northernmost part of the WPWP, is consistently related to precession variations over the past 450 kyr, except during the ∼410-370 ka BP period (Fig. 4f) (likely due to missing original data in the core MD06-3047B δ18Osw-ivc record during this specific period). Furthermore, the filtered 23-kyr component of the δ18Osw-ivc record from core MD06-3067 in the northern IPWP also demonstrates synchronous changes with precession variations over the past ∼160 kyr, where each peak in SSS corresponds to both a low in precession and a peak in equatorial insolation during boreal summer (Fig. 4g; Bolliet et al. Bolliet et al., Reference Bolliet, Holbourn, Kuhnt, Laj, Kissel, Beaufort, Kienast, Andersen and Garbe-Schönberg2011).

In contrast, the δ18Osw-ivc record from the central IPWP core KX97322-4 exhibits limited coherence with precession variability over the past 360 kyr (Zhang et al. Reference Zhang, Yu, Gong, Wang, Chang, Lohmman, Qi and Li2021). This lack of coherence might be attributed to an ENSO-like state or potentially to the fact that δ18Osw-ivc is influenced by both precession and obliquity (Zhang et al. Reference Zhang, Li, Chang, Yu and Wang2015, Reference Zhang, Yu, Gong, Wang, Chang, Lohmman, Qi and Li2021). Obliquity may affect the atmospheric circulation and the associated meridional heat and moisture fluxes by regulating the latitudinal distribution of insolation (Mantsis et al. Reference Mantsis, Clement, Broccoli and Erb2011, Reference Mantsis, Lintner, Broccoli, Erb, Clement and Park2014; Bosmans et al. Reference Bosmans, Drijfhout, Tuenter, Hilgen, Lourens and Rohling2015). This might explain why obliquity bands are present in the δ18Osw-ivc record of core KX97322-4, as obliquity could affect local precipitation through atmospheric circulation and heat and moisture fluxes. Additionally, variations in the δ18Osw-ivc signals may be due in part to differences in core locations capturing divergent information. Moreover, Liu et al. (Reference Liu, Lo, Shi, Wei, Chou, Chen, Chuang, Wu, Mii, Peng, Amakawa, Burr, Lee, DeLong, Elderfield and Shen2015) proposed that precipitation variations in the western Pacific are more affected by obliquity than by precession-paced changes in the ITCZ, based on foraminiferal rare earth elements (indicator of precipitation-dependent river runoff). This change in dominant cyclicity could be attributed to differences in proxy-specific bias or tendency. In the eastern IPWP, the δ18Osw-ivc records from cores SO225-08 (unpublished data from Nürnberg) and ODP806 (Lea et al. Reference Lea, Pak and Spero2000), located in the eastern IPWP, did not exhibit synchronous changes with precession parameters but rather showed distinct glacial-interglacial cycles. This discrepancy might be attributed to the relatively low sample resolution in these records. Cores SO225-08 and ODP806, being situated close to the central equatorial Pacific, might experience more influence on their hydrological conditions by atmospheric circulation, such as the meridional Hadley circulation and the Walker circulation (Lea et al. Reference Lea, Pak and Spero2000). In conclusion, the variations in δ18Osw-ivc across various regions of the WPWP are not entirely identical, possibly due to regional differences, varying impact of driving factors, and/or sample resolution. However, with respect to precessional cyclicity, all precipitation records within the central precipitation zone across the WPWP generally show similar patterns (Fig. 4).

Notably, the pattern of precipitation changes in the western part of the IPWP (eastern tropical Indian Ocean) is similar to in the western tropical Pacific. The filtered 23-kyr component of the δ18Osw-ivc records from the southern part of the Makassar Strait (core MD98-2162) and Timor Sea (cores SO18480-3 and U1483) close to the equator demonstrate a nearly synchronous response to precession-driven boreal summer insolation variations (Fig. 4c–e; Jian et al. Reference Jian, Wang, Dang, Mohtadi, Rosenthal, Lea, Liu, Jin, Ye, Kuhnt and Wang2022; Zhang et al. Reference Zhang, Xu, Holbourn, Kuhnt, Xiong and Li2022). Higher insolation levels correspond to higher δ18Osw-ivc values, hence more saline sea-surface conditions, while lower insolation levels correspond to lowered δ18Osw-ivc values and fresher conditions (Fig. 4c–e). This covariance is consistent with the precipitation record of Site U1486. Overall, nearly all records are characterized by higher δ18Osw-ivc values (higher SSS; lower precipitation) at precession minima, implying that both the eastern tropical Indian Ocean and the western tropical Pacific within the IPWP experienced drier conditions simultaneously during lower precession periods (Fig. 4).

Consistently, the simulated annual mean δ18Osw-ivc derived from a water isotope (δ18Osw)-enabled air-sea coupled climate model (GISS_ModelE2-R; Jian et al. Reference Jian, Wang, Dang, Mohtadi, Rosenthal, Lea, Liu, Jin, Ye, Kuhnt and Wang2022) reveals a dominant 23-kyr cycle with higher δ18Osw-ivc values and hence, more saline conditions at sea surface during precession minima (Fig. 4b), supporting our notion on synchronous precession-driven precipitation changes in the IPWP. We conclude that the variations in equatorial insolation, driven by precession parameters, likely influence the SSS through precipitation variability within the IPWP region.

4.c. Effects of the ITCZ changes on the precipitation variations in IPWP

4.c.1. Variations in ITCZ latitudinal migration

The IPWP is a significant moisture source for the East Asian summer monsoon (EASM). Numerous studies demonstrated a strong association between EASM variability and the hydrological conditions as well as convective activity in the IPWP (e.g., Lau et al. Reference Lau, Kim and Yang2000; Lu, Reference Lu2001; Kawamura et al. Reference Kawamura, Matsuura and Iizuka2001; Yin et al. Reference Yin, Singh, Berger, Guo and Crucifix2014). The records of Chinese stalagmite δ18O, commonly used to reconstruct EASM variability, display distinct precession-related changes with intense precipitation during higher Northern Hemisphere summer insolation over the past 450 kyr (Fig. 5b; Cheng et al. Reference Cheng, Edwards, Sinha, Spotl, Yi, Chen, Kelly, Kathayat, Wang, Li, Kong, Wang, Ning and Zhang2016). The strong precession signal observed in Chinese speleothem records corresponding to insolation patterns is consistent with climate model simulations (e.g., Battisti et al. Reference Battisti, Ding and Roe2014; Rachmayani et al. Reference Rachmayani, Prange and Schulz2016). The Site U1486 δ18Osw-ivc variations generally show a pronounced anti-phased relationship with Chinese stalagmite δ18O on precessional timescales (Fig. 5b, f). This observation is further corroborated by the coherence analysis (Fig. 5h) and by most of the precipitation records within the IPWP region (Fig. 4). When δ18Osw-ivc values increase (lower precipitation) in the IPWP, the δ18O values in Chinese speleothems decrease (higher precipitation). Conversely, periods of high precipitation in the IPWP correspond to a weakening of the EASM. Additionally, the simulated δ18Osw-ivc of the IPWP and measured δ18O values in Chinese speleothems are anti-correlated on precessional timescales (Fig. 5b, e; Jian et al. Reference Jian, Wang, Dang, Mohtadi, Rosenthal, Lea, Liu, Jin, Ye, Kuhnt and Wang2022), which is further confirmed by cross-spectral analysis (Fig. 5h).

The anti-phased relationship between δ18Osw-ivc of the IPWP and Chinese speleothem δ18O is associated with the ocean heat content of the IPWP (Jian et al. Reference Jian, Wang, Dang, Mohtadi, Rosenthal, Lea, Liu, Jin, Ye, Kuhnt and Wang2022). During low precession phases (high Northern Hemisphere summer insolation), La Niña-like conditions prevail in the western tropical Pacific. This is indicated by both the CESM-simulated Niño3 SST of December–January–February (DJF) (Zhang et al. Reference Zhang, Yu, Gong, Wang, Chang, Lohmman, Qi and Li2021), and the temperature gradient between sea surface and thermocline waters (ΔT). A deep thermocline at Site U1486 deduced from a small ΔT (unpublished data) implies enhanced ocean heat storage in the mixed layer (Fig. 5d, e). This higher ocean heat content likely amplifies the strength of the EASM, enhancing moisture transport from ocean to continent and resulting in a negative deviation in Chinese speleothem δ18O values (Jian et al. Reference Jian, Wang, Dang, Mohtadi, Rosenthal, Lea, Liu, Jin, Ye, Kuhnt and Wang2022). The precipitation changes in the IPWP and EASM regions exhibit a strong coherence to local insolation changes. Therefore, their precipitation changes may be attributed to the redistribution of heat between the Northern and Southern Hemispheres triggered/amplified by precessional forcing, a conclusion which is consistent with model simulations (Braconnot et al. Reference Braconnot, Marzin, Gregoire, Mosquet and Marti2008; Battisti et al. Reference Battisti, Ding and Roe2014).

The IPWP, located within the belt of the modern ITCZ’s seasonal migration (Fig. 1a), experiences precipitation primarily regulated by the seasonal latitudinal movement of the ITCZ (Trenberth et al. Reference Trenberth, Stepaniak and Caron2000; Schmidt & Spero, Reference Schmidt and Spero2011; Schneider et al. Reference Schneider, Bischoff and Haug2014; Jo et al. Reference Jo, Woo, Yi, Yang, Lim, Wang, Cheng and Edwards2014). During periods of elevated summer insolation in the Northern Hemisphere, the ITCZ migrates northward (e.g., McGee et al. Reference McGee, Donohoe, Marshall and Ferreira2014), lowering precipitation in IPWP (Schmidt & Spero, Reference Schmidt and Spero2011; Jo et al. Reference Jo, Woo, Yi, Yang, Lim, Wang, Cheng and Edwards2014). This low in IPWP precipitation is supported by the high values of δ18Osw-ivc records from the IPWP region during the lower precession periods, which correspond to the high summer insolation in the Northern Hemisphere (Fig. 4). Seasonal variations in the thermal contrast between the Northern and Southern Hemispheres prompted the migration of the ITCZ, shifting toward the warmer hemisphere (Schneider et al. Reference Schneider, Bischoff and Haug2014). During the boreal summer, increasing insolation warms the subtropical water, strengthening the meridional pressure gradient and enhancing the southerlies flowing across the equator into the ITCZ, leading the ITCZ to move northward (Chang & Philander, Reference Chang and Philander1994; Geen et al. Reference Geen, Bordoni, Battisti and Hui2020). Therefore, the migration of the mean position of the ITCZ could be indicated by the disparity in summer insolation between hemispheres (Chiang & Friedman Reference Chiang and Friedman2012; Lu et al. Reference Lu, Yi, Liu, Mason, Jiang, Cheng, Stevens, Xu, Zhang, Jin, Zhang, Guo, Wang and Otto-Bliesner2013; Schneider et al. Reference Schneider, Bischoff and Haug2014). We calculated the summer interhemispheric insolation gradient between 23° N and 23° S (Fig. 5c; Laskar et al. Reference Laskar, Robutel, Joutel, Gastineau, Correia and Levrard2004). A higher summer interhemispheric gradient during the precessional minima points to the northward migration of the ITCZ and reduced precipitation in the IPWP (Fig. 5f, g). Conversely, a decreased meridional summer interhemispheric insolation gradient results in the southward shift of the ITCZ, causing increased precipitation in the IPWP (Fig. 5c, f, g). During periods of Northern Hemisphere warming, such as the Bølling–Allerød and the early Holocene epochs characterized by a strong EASM, the ITCZ shifted northward (Wang et al. Reference Wang, Cheng, Edwards, An, Wu, Shen and Dorale2001; Yancheva et al. Reference Yancheva, Nowaczyk, Mingram, Dulski, Schettler, Negendank, Liu, Sigman, Peterson and Haug2007). Recent research further suggests that the location and migrations of the ITCZ, where the trade winds of the Northern and Southern Hemispheres converge, are closely associated with variations in regional monsoon patterns (Gadgil, Reference Gadgil2018, and references therein). Hence, the strength of the EASM is closely linked to the migrations of the ITCZ (e.g., Ding et al. Reference Ding, Liu, Rutter, Yu, Guo and Zhu1995; An, Reference An2000; Yancheva et al. Reference Yancheva, Nowaczyk, Mingram, Dulski, Schettler, Negendank, Liu, Sigman, Peterson and Haug2007). Chinese stalagmite δ18O records revealed that the intensification of the EASM during the precessional minima, accompanied by the northward migration of the ITCZ, resulting in high precipitation in the EASM region (Fig. 5b, c; Cheng et al. Reference Cheng, Edwards, Sinha, Spotl, Yi, Chen, Kelly, Kathayat, Wang, Li, Kong, Wang, Ning and Zhang2016). These findings emphasize the significant impact of ITCZ variations on hydrological changes within the IPWP and EASM regions. Orbital precession played a pivotal role in determining the seasonal distribution of incoming insolation in both hemispheres. This drives the north-south migration of the ITCZ and causes the subtropical high-pressure anomalies in the Indian and Pacific Oceans (Merlis et al. Reference Merlis, Schneider, Bordoni and Eisenman2013; Schneider et al. Reference Schneider, Bischoff and Haug2014), leading to anti-correlated changes in precipitation between the IPWP and EASM region.

The mechanisms described above, supported by other paleoclimatic and modeling studies spanning millennial to orbital timescales (e.g., Wang et al. Reference Wang, Auler, Edwards, Cheng, Ito and Solheid2006; Merlis et al. Reference Merlis, Schneider, Bordoni and Eisenman2013; Dang et al. Reference Dang, Jian, Kissel and Bassinot2015; Huang et al. Reference Huang, Wan, Li and Li2019), provide a robust explanation for the interhemispheric anti-phased precipitation pattern between the IPWP and EASM region. Nevertheless, precipitation records in the northern equatorial region of the IPWP (e.g., cores MD06-3047B and MD06-3067; Bolliet et al. Bolliet et al., Reference Bolliet, Holbourn, Kuhnt, Laj, Kissel, Beaufort, Kienast, Andersen and Garbe-Schönberg2011; Jia et al. Jia et al., Reference Jia, Li, Xiong, Steinke, Jiang, Chang and Qin2018) exhibit nearly synchronous interhemispheric changes with the southern IPWP region (Fig. 4). The migration of the ITCZ alone might not adequately account for the synchronous interhemispheric precipitation changes observed. Today, the ITCZ shifts between 9°N in boreal summer and 2°N in boreal winter across the modern Pacific Ocean (Schneider et al. Reference Schneider, Bischoff and Haug2014). We speculate that during periods of high precession when the ITCZ migrates southward, the ITCZ might still remain in the Northern Hemisphere, resulting in heightened precipitation over both the northern and southern IPWP regions. Besides, we propose that the synchronized interhemispheric precipitation changes in the northern and southern WPWP could be linked to the expansion and contraction of the ITCZ on the precession band (see below).

4.c.2. Variations in ITCZ expansion and contraction

Numerous studies have revealed the in-phase changes in precipitation between hemispheres over (sub-)millennial timescales, which can be effectively explained by the expansion and contraction of the ITCZ, specifically its width (e.g., Collins et al. Reference Collins, Schefuß, Heslop, Mulitza, Prange, Zabel, Tjallingii, Dokken, Huang, Mackensen, Schulz, Tian, Zarriess and Wefer2010; Konecky et al. Reference Konecky, Russell, Rodysill, Vuille, Bijaksana and Huang2013; Yan et al. Reference Yan, Wei, Soon, An, Zhou, Liu, Wang and Carter2015; Denniston et al. Reference Denniston, Ummenhofer, Wanamaker, Lachniet, Villarini, Asmerom, Polyak, Passaro, Cugley, Woods and Humphreys2016; Scroxton et al. Reference Scroxton, Burns, McGee, Hardt, Godfrey, Ranivoharimanana and Faina2017; Yang et al. Reference Yang, Zhang, Yi, Zhong, Lu, Wan, Du and Xiang2023). Yang et al. (Reference Yang, Zhang, Yi, Zhong, Lu, Wan, Du and Xiang2023) discovered that both the southern and northern margins of the ITCZ experienced simultaneously arid tropical Indian Ocean hydrological conditions during the early stage of the deglacial Heinrich Stadial 1 period, indicating a contracted tropical precipitation belt during that time. The precession bandpass filtered δ18Osw-ivc records of core MD06-3047B (located near the northern margin of the ITCZ), and Site U1483 and core MD05-2925 (located in the southern margin of the ITCZ) show reduced precipitation during most lower precession periods (Fig. 6b–d; Jia et al. Jia et al., Reference Jia, Li, Xiong, Steinke, Jiang, Chang and Qin2018; Zhang et al. Reference Zhang, Xu, Holbourn, Kuhnt, Xiong and Li2022; Lo et al. Reference Lo, Shen, Zeeden, Tsai, Yin, Yang, Chang, Su, Mii, Chuang and Chen2022), possibly indicating the contraction of the ITCZ in the IPWP region and vice versa. While the precipitation records in the northern and southern marginal ITCZ are not consistently in-phase during certain periods (e.g., 390–360 ka BP), this discrepancy may be attributed to regional influencing factors or stratigraphical issues. However, collectively, these records consistently demonstrate high coherence with precession variability, suggesting that the ITCZ contracts during lower precession periods and expands during higher precession periods. We hypothesize that the synchronization of precipitation in both northern and southern regions of the IPWP on the precession band may also be attributed to the expansion and contraction of the ITCZ. Coincident with the ascending branch of the Hadley circulation, the ITCZ is typically delineated as the convergence zone of the trade winds from the Northern and Southern Hemispheres, thus strongly impacted by regional tropical climate dynamics (Byrne & Schneider, Reference Byrne and Schneider2016; Geen et al. Reference Geen, Bordoni, Battisti and Hui2020). Studies suggest that on (sub)-millennial timescales, the ITCZ width may be linked to variations in the Walker circulation (Konecky et al. Reference Konecky, Russell, Rodysill, Vuille, Bijaksana and Huang2013) and meridional atmospheric circulation (Denniston et al. Reference Denniston, Ummenhofer, Wanamaker, Lachniet, Villarini, Asmerom, Polyak, Passaro, Cugley, Woods and Humphreys2016). Consequently, we propose that synchronous precipitation changes between the hemispheres on orbital timescales might be intricately tied to changes in regional atmospheric circulation, which could affect the expansion and contraction of the ITCZ.

Figure 6. Effects of the ITCZ migration and width on the precipitation variations in the IPWP on the precession band. A. The records of ITCZ migration and width variation: (a) The variation in precession (Laskar et al. Reference Laskar, Robutel, Joutel, Gastineau, Correia and Levrard2004). (b, c, d) Precession bandpass filtering curves of δ18Osw-ivc records in northern marginal ITCZ (core MD06-3047B; Jia et al. Jia et al., Reference Jia, Li, Xiong, Steinke, Jiang, Chang and Qin2018) and southern marginal ITCZ (core MD05-2925 and Site U1483; Lo et al. Reference Lo, Shen, Zeeden, Tsai, Yin, Yang, Chang, Su, Mii, Chuang and Chen2022; Zhang et al. Reference Zhang, Xu, Holbourn, Kuhnt, Xiong and Li2022). (e) Gradients of δ18Osw−ivc between Site U1486 and core MD06-3047B (Jia et al. Jia et al., Reference Jia, Li, Xiong, Steinke, Jiang, Chang and Qin2018). (f) Gradients of δ18Osw−ivc between Site U1486 and core MD05-2925 (Lo et al. Reference Lo, Shen, Zeeden, Tsai, Yin, Yang, Chang, Su, Mii, Chuang and Chen2022). (g) The summer interhemispheric tropical insolation gradient, determined by the variation in June insolation between 23° N and 23° S (Laskar et al. Reference Laskar, Robutel, Joutel, Gastineau, Correia and Levrard2004). Vertical grey bars indicate lower precession values. B. Schematic diagrams showing ITCZ variability during low-precession (left) and high-precession (right) intervals. The pink shaded band indicates the position and size of the ITCZ, referring to the belt of the modern ITCZ’s seasonal migration.

Yuan et al. (Reference Yuan, Chiang, Liu, Bijaksana, He, Jiang, Imran, Wicaksono and Wang2023) suggested that the larger (smaller) discrepancy in rainfall between the central and marginal regions of the ITCZ corresponds to a more contracted (expanded) ITCZ. In our study, we calculated the differences in δ18Osw-ivc (Δδ18Osw-ivc) among Site U1486 (central ITCZ), core MD06-3047B (near the northern marginal ITCZ) and core MD05-2925 (southern marginal ITCZ) to explore the variations in ITCZ contraction/expansion on orbital timescales (Fig. 6e, f). It should be noted that although δ18Osw-ivc may be influenced by various factors, we hypothesize that it primarily reflects changes in the precipitation-evaporation balance at the core sites in the IPWP, as discussed in Section 4.a. However, we acknowledge the presence of a leading or lagging relationship between Δδ18Osw-ivc and the expansion and contraction of the ITCZ on precession bands over the past 450 kyr (Fig. 6a, e, f), which does not exclude the possibility of δ18Osw-ivc being influenced by other factors or differences in core age models. We use this approach of Δδ18Osw-ivc to solely qualitatively describe the changes in ITCZ contraction and expansion during low-precession and high-precession intervals. Overall, the Δδ18Osw-ivc records between central and marginal ITCZ regions both increase during lower precession periods, while smaller precipitation differences (reduced Δδ18Osw-ivc) indicate an expanded ITCZ during higher precession periods (Fig. 6e, f).

Yan et al. (Reference Yan, Wei, Soon, An, Zhou, Liu, Wang and Carter2015) proposed that the ITCZ’s expansion and contraction correspond with symmetrical changes in insolation between hemispheres on sub-millennial timescales. Furthermore, recent modeling work by Singarayer et al. (Reference Singarayer, Valdes and Roberts2017) indicate that precessional changes in the ITCZ width may result from different responses of its northern and southern extremes to variations in the interhemispheric temperature gradient. Thus, we suggest that variations in interhemispheric insolation due to precession may influence the expansion and contraction of the ITCZ. These hemispheric insolation variations could affect the interactions between radiation and clouds and thus ITCZ width, as suggested by the climate models (Voigt & Shaw, Reference Voigt and Shaw2015; Lau & Kim, Reference Lau and Kim2015). The contracted and expanded positions of the ITCZ, as depicted in Figure 6B, were determined based on various precipitation records across different IPWP regions, providing a qualitative overview of the ITCZ positions. Some studies suggests that the ITCZ might be influenced by extratropical forcing, such as the North Hemisphere ice sheet and the Atlantic meridional overturning circulation (Kang et al. Reference Kang, Held, Frierson and Zhao2008). If so, variations in the low/high precession parameter under different climate conditions could lead to differing latitudinal shifts and sizes of the ITCZ. However, in our study we hypothesize that position and size of the ITCZ remain constant under all low (high) precession periods. Here, the position and size refer to the belt of the modern ITCZ’s seasonal migration (Fig. 1a). In summary, we suggest that the precessional hydroclimatic synchronicity changes in the IPWP might be driven by both the latitudinal migration of the ITCZ and the variations in its width over the past 450 kyr. The ITCZ migrated northward (southward) and contracted (expanded) during the periods of low (high) precession, as indicated by enriched (depleted) precipitation in the EASM region and in-phase depleted (enriched) precipitation in the entire IPWP region (Fig. 6B).

5. Conclusions

We utilized high-resolution δ18Osw-ivc records approximating SSS from Site U1486, retrieved from the Sepik River mouth in Papua New Guinea, to reconstruct precipitation changes within the IPWP over the past 450 kyr. The δ18Osw-ivc records display a dominant 23 kyr periodicity without a distinct glacial-interglacial trend. Our analysis focused on examining the spatial pattern of precipitation variations across the IPWP over orbital timescales, revealing a synchronous precipitation pattern on precession band. We assert that changes in ITCZ played a dominant role in determining the reconstructed precipitation patterns in the IPWP region on precession band. The anti-phased precipitation changes between the IPWP and the EASM region unveiled variations in the ITCZ migration, which was determined by the seasonal distribution of incoming insolation in both hemispheres on precession band. The synchronous precipitation changes between the two hemispheres within the IPWP emphasized the significant impact of the ITCZ’s expansion and contraction on precipitation variability within the IPWP over precession timescales. Our findings suggest that during low precession periods, the ITCZ migrated northward and contracted, leading to increased precipitation in the EASM region and synchronous reduction of precipitation within the various locations across the IPWP.

Supplementary material

To view supplementary material for this article, please visit https://doi.org/10.1017/S0016756824000177

Acknowledgments

We thank all members of IODP Expedition 363 for their efforts in gathering the fundamental information and samples essential for this study. This study was financially supported by Laoshan Laboratory (No. LSKJ202204201), the Strategic Priority Research Program of the Chinese Academy of Sciences (Grant No. XDB42000000), and the National Natural Science Foundation of China (Grant Nos. 42076051, 42076050 and 41876041). We thank David McGee and Jiangnan Shi for providing useful feedback that improved the manuscript.

References

An, ZS (2000) The history and variability of the East Asian paleomonsoon climate. Quaternary Science Reviews 19, 171–87. doi: 10.1016/S0277-3791(99)00060-8 CrossRefGoogle Scholar
Barker, S, Greaves, M and Elderfield, H (2003) A study of cleaning procedures used for foraminiferal Mg/Ca paleothermometry. Geochemistry, Geophysics, Geosystems 4, 120. doi: 10.1029/2003gc000559 CrossRefGoogle Scholar
Battisti, DS, Ding, Q and Roe, GH (2014) Coherent pan-Asian climatic and isotopic response to orbital forcing of tropical insolation. Journal of Geophysical Research: Atmospheres 119, 11997. doi: 10.1002/2014jd021960 Google Scholar
Bemis, BE, Spero, HJ, Bijma, J and Lea, DW (1998) Reevaluation of the oxygen isotopic composition of planktonic foraminifera: experimental results and revised paleotemperature equations. Paleoceanography 13, 150–60.CrossRefGoogle Scholar
Berger, A and Loutre, MF (1994) Precession, eccentricity, obliquity, insolation and paleoclimates. In Long-Term Climatic Variations. NATO ASI Series, vol 22 (eds Duplessy, JC & Spyridakis, MT) Berlin, Heidelberg: Springer. doi: 10.1007/978-3-642-79066-9_5 Google Scholar
Berger, AL (1978) Long-term variations of daily insolation and quaternary climate changes. Journal of the Atmospheric Sciences 35, 2362–67.2.0.CO;2>CrossRefGoogle Scholar
Bintanja, R, van de Wal, RS and Oerlemans, J (2005) Modelled atmospheric temperatures and global sea levels over the past million years. Nature 437, 125–28. doi: 10.1038/nature03975 CrossRefGoogle ScholarPubMed
Bolliet, T, Holbourn, A, Kuhnt, W, Laj, C, Kissel, C, Beaufort, L, Kienast, M, Andersen, N and Garbe-Schönberg, D (2011) Mindanao Dome variability over the last 160 kyr: episodic glacial cooling of the West Pacific Warm Pool. Paleoceanography 26, PA1208. doi: 10.1029/2010pa001966 CrossRefGoogle Scholar
Bosmans, JHC, Drijfhout, SS, Tuenter, E, Hilgen, FJ, Lourens, LJ and Rohling, EJ (2015) Precession and obliquity forcing of the freshwater budget over the Mediterranean. Quaternary Science Reviews 123, 1630. doi: 10.1016/j.quascirev.2015.06.008 CrossRefGoogle Scholar
Braconnot, P, Marzin, C, Gregoire, L, Mosquet, E and Marti, O (2008) Monsoon response to changes in Earth’s orbital parameters: comparisons between simulations of the Eemian and of the Holocene. Climate of the Past 4, 281–94. doi: 10.5194/cp-4-281-2008 CrossRefGoogle Scholar
Byrne, MP and Schneider, T (2016) Narrowing of the ITCZ in a warming climate: physical mechanisms. Geophysical Research Letters 43, 11350. doi: 10.1002/2016gl070396 CrossRefGoogle Scholar
Cane, MA (1998) A role for the tropical Pacific. Science 282, 5961.CrossRefGoogle Scholar
Chang, P and Philander, SG (1994) A coupled ocean–atmosphere instability of relevance to the seasonal cycle. Journal of the Atmospheric Sciences 51, 3627–48. doi: 10.1175/1520-0469(1994)051<3627:ACOIOR>2.0.CO;2 2.0.CO;2>CrossRefGoogle Scholar
Cheng, H, Edwards, RL, Sinha, A, Spotl, C, Yi, L, Chen, S, Kelly, M, Kathayat, G, Wang, X, Li, X, Kong, X, Wang, Y, Ning, Y and Zhang, H (2016) The Asian monsoon over the past 640,000 years and ice age terminations. Nature 534, 640–46. doi: 10.1038/nature18591 CrossRefGoogle ScholarPubMed
Chiang, JCH and Friedman, AR (2012) Extratropical cooling, interhemispheric thermal gradients, and tropical climate change. Annual Review of Earth and Planetary Sciences 40, 383412. doi: 10.1146/annurev-earth-042711-105545 CrossRefGoogle Scholar
Clement, AC, Hall, A and Broccoli, A (2004) The importance of precessional signals in the tropical climate. Climate Dynamics 22, 327–41. doi: 10.1007/s00382-003-0375-8 CrossRefGoogle Scholar
Clement, AC, Seager, R and Murtugudde, R (2005) Why are there tropical warm pools? Journal of Climate 18, 5294–311. doi: 10.1175/JCLI3582.1 CrossRefGoogle Scholar
Collins, JA, Schefuß, E, Heslop, D, Mulitza, S, Prange, M, Zabel, M, Tjallingii, R, Dokken, TM, Huang, E, Mackensen, A, Schulz, M, Tian, J, Zarriess, M and Wefer, G (2010) Interhemispheric symmetry of the tropical African rainbelt over the past 23,000 years. Nature Geoscience 4, 42–5. doi: 10.1038/ngeo1039 CrossRefGoogle Scholar
Conroy, JL, Thompson, DM, Cobb, KM, Noone, D, Rea, S and LeGrande, AN (2017) Spatiotemporal variability in the delta (18)O-salinity relationship of seawater across the tropical Pacific Ocean. Paleoceanography 32, 484–97. doi: 10.1002/2016PA003073 CrossRefGoogle Scholar
Dai, A and Wigley, TML (2000) Global patterns of ENSO-induced precipitation. Geophysical Research Letters 27, 1283–86. doi: 10.1029/1999GL011140 CrossRefGoogle Scholar
Dang, H, Jian, Z, Kissel, C and Bassinot, F (2015) Precessional changes in the western equatorial Pacific Hydroclimate: a 240 kyr marine record from the Halmahera Sea, East Indonesia. Geochemistry, Geophysics, Geosystems 16, 148–64. doi: 10.1002/2014gc005550 CrossRefGoogle Scholar
De Deckker, P (2016) The Indo-Pacific Warm Pool: critical to world oceanography and world climate. Geoscience Letters 3, 20. doi: 10.1186/s40562-016-0054-3 CrossRefGoogle Scholar
Dekens, PS., Lea, DW, Pak, DK and Spero, HJ (2002) core top calibration of Mg/Ca in tropical foraminifera: refining paleotemperature estimation. Geochemistry, Geophysics, Geosystems 3, 129. doi: 10.1029/2001gc000200 CrossRefGoogle Scholar
Denniston, RF, Ummenhofer, CC, Wanamaker, AD, Lachniet, MS, Villarini, G, Asmerom, Y, Polyak, VJ, Passaro, KJ, Cugley, J, Woods, D and Humphreys, WF (2016) Expansion and contraction of the Indo-Pacific tropical rain belt over the last three Millennia. Scientific Reports 6, 34485. doi: 10.1038/srep34485 CrossRefGoogle ScholarPubMed
Ding, ZL, Liu, TS, Rutter, NW, Yu, ZW, Guo, ZT and Zhu, RX (1995) Ice-volume forcing of East-Asian winter monsoon variations in the past 800,000 years, Quaternary Research 44, 149–59.CrossRefGoogle Scholar
Fairbanks, RG, Evans, MN, Rubenstone, JL, Mortlock, RA, Broad, K, Moore, MD and Charles, CD (1997) Evaluating climate indices and their geochemical proxies measured in corals. Coral Reefs 16, S93S100.CrossRefGoogle Scholar
Gadgil, S (2018) The monsoon system: land-sea breeze or the ITCZ? Journal of Earth System Science 127, 129. doi: 10.1007/s12040-017-0916-x CrossRefGoogle Scholar
Geen, R, Bordoni, S, Battisti, DS and Hui, K (2020) Monsoons, ITCZs, and the concept of the global monsoon. Reviews of Geophysics 58, e2020RG000700. doi: 10.1029/2020rg000700 CrossRefGoogle Scholar
Gibbons, FT, Oppo, DW, Mohtadi, M, Rosenthal, Y, Cheng, J, Liu, Z and Linsley, BK (2014) Deglacial δ18O and hydrologic variability in the tropical Pacific and Indian Oceans. Earth and Planetary Science Letters 387, 240–51. doi: 10.1016/j.epsl.2013.11.032 CrossRefGoogle Scholar
Hammer, Ø, Harper, D and Ryan, P (2001) PAST-Paeontological Statistics, ver. 1.89. Palaeontology Electron 4, 19.Google Scholar
Hoerling, MP, Hurrell, JW and Xu, T (2001) Tropical origins for recent North Atlantic climate change. Science 292, 90–2. doi: 10.1126/science.1058582 CrossRefGoogle ScholarPubMed
Holloway, MD, Sime, LC, Singarayer, JS, Tindall, JC and Valdes, PJ (2016) Recon structing paleosalinity from δ18O: coupled model simulations of the last glacial maximum, last interglacial and late Holocene. Quaternary Science Reviews 131, 350–64. doi: 10.1016/j.quascirev.2015.07.007 CrossRefGoogle Scholar
Hollstein, M, Mohtadi, M, Kienast, M, Rosenthal, Y, Groeneveld, J, Oppo, DW, Southon, JR and Lückge, A (2020) The impact of astronomical forcing on surface and thermocline variability within the Western Pacific warm pool over the past 160 kyr. Paleoceanography and Paleoclimatology 35, e2019PA003832. doi: 10.1029/2019pa003832 CrossRefGoogle Scholar
Hollstein, M, Mohtadi, M, Rosenthal, Y, Moffa Sanchez, P, Oppo, D, Martínez Méndez, G, Steinke, S and Hebbeln, D (2017) Stable oxygen isotopes and Mg/Ca in planktic foraminifera from modern surface sediments of the western Pacific warm pool: Implications for thermocline reconstructions. Paleoceanography 32(11), 11741194. doi: 10.1002/2017pa003122 CrossRefGoogle Scholar
Huang, J, Wan, S, Li, A and Li, T (2019) Two-phase structure of tropical hydroclimate during Heinrich Stadial 1 and its global implications. Quaternary Science Reviews 222, 105900. doi: 10.1016/j.quascirev.2019.105900 CrossRefGoogle Scholar
Hut, G (1987) In: Consultants Group Meeting on Stable Isotopic Reference Samples for Geochemical and Hydrological Investigations. International Atomic Energy Agency, Vienna, p. 42.Google Scholar
Jalihal, C, Bosmans, JHC, Srinivasan, J and Chakraborty, A (2019) The response of tropical precipitation to earth’s precession: the role of energy fluxes and vertical stability. Climate of the Past 15, 449–62. doi: 10.5194/cp-15-449-2019 CrossRefGoogle Scholar
Jia, Q, Li, T, Xiong, Z, Steinke, S, Jiang, F, Chang, F and Qin, B (2018) Hydrological variability in the western tropical Pacific over the past 700 kyr and its linkage to Northern Hemisphere climatic change. Palaeogeography, Palaeoclimatology, Palaeoecology 493, 4454. doi: 10.1016/j.palaeo.2017.12.039 CrossRefGoogle Scholar
Jian, Z, Wang, Y, Dang, H, Mohtadi, M, Rosenthal, Y, Lea, DW, Liu, Z, Jin, H, Ye, L, Kuhnt, W and Wang, X (2022) Warm pool ocean heat content regulates ocean-continent moisture transport. Nature 612, 92–9. doi: 10.1038/s41586-022-05302-y CrossRefGoogle ScholarPubMed
Jo, KN, Woo, KS, Yi, S, Yang, DY, Lim, HS, Wang, Y, Cheng, H and Edwards, RL (2014) Mid-latitude interhemispheric hydrologic seesaw over the past 550,000 years. Nature 508, 378–82. doi: 10.1038/nature13076 CrossRefGoogle ScholarPubMed
Kang, SM, Held, IM, Frierson, DMW and Zhao, M (2008) The response of the ITCZ to extratropical thermal forcing: idealized slab-Ocean experiments with a GCM. Journal of Climate 21, 3521–32. doi: 10.1175/2007jcli2146.1 CrossRefGoogle Scholar
Kawamura, R, Matsuura, T and Iizuka, S (2001) Role of equatorial asymmetric sea surface temperature anomalies in the Indian Ocean on the Asian summer monsoon and El Nino–southern oscillation coupling. Journal of Geophysical Research 106, 4681–93. doi: 10.1029/2000JD900610 CrossRefGoogle Scholar
Kim, ST, Yu, JY and Lu, MM (2012) The distinct behaviors of Pacific and Indian Ocean warm pool properties on seasonal and interannual time scales. Journal of Geophysical Research: Atmospheres 117, D05128. doi: 10.1029/2011jd016557 CrossRefGoogle Scholar
Kissel, C, Laj, C, Kienast, M, Bolliet, T, Holbourn, A, Hill, P, Kuhnt, W and Braconnot, P (2010) Monsoon variability and deep oceanic circulation in the western equatorial Pacific over the last climatic cycle: insights from sedimentary magnetic properties and sortable silt. Paleoceanography 25, PA3215. doi: 10.1029/2010pa001980 CrossRefGoogle Scholar
Konecky, BL, Russell, JM, Rodysill, JR, Vuille, M, Bijaksana, S and Huang, Y (2013) Intensification of southwestern Indonesian rainfall over the past millennium. Geophysical Research Letters 40, 386–91. doi: 10.1029/2012gl054331 CrossRefGoogle Scholar
Kutzbach, JE, Liu, X, Liu, Z and Chen, G (2008) Simulation of the evolutionary response of global summer monsoons to orbital forcing over the past 280,000 years. Climate Dynamics 30, 567–79. doi: 10.1007/s00382-007-0308-z CrossRefGoogle Scholar
Laskar, J, Robutel, P, Joutel, F, Gastineau, M, Correia, ACM and Levrard, B (2004) A long-term numerical solution for the insolation quantities of the Earth. Astronomy & Astrophysics 428, 261–85.CrossRefGoogle Scholar
Lau, KM and Kim, KM (2015) Robust Hadley circulation changes and increasing global dryness due to CO2 warming from CMIP5 model projections. Proceedings of the National Academy of Sciences 112, 3630–35. doi: 10.1073/pnas.1418682112 CrossRefGoogle ScholarPubMed
Lau, KM, Kim, KM and Yang, S (2000) Dynamical and boundary forcing characteristics of regional components of the Asian summer monsoon. Journal of Climate 13, 2461–82. doi: 10.1175/1520-0442(2000)013<2461:DABFCO>2.0.CO;2 2.0.CO;2>CrossRefGoogle Scholar
Lea, DW, Pak, DK and Spero, HJ (2000) Climate impact of late quaternary equatorial pacific sea surface temperature variations. Science 289, 1719–24. doi: 10.1126/science.289.5485.1719 CrossRefGoogle ScholarPubMed
Leech, PJ, Lynch-Stieglitz, J and Zhang, R (2013) Western pacific thermocline structure and the pacific marine intertropical convergence zone during the last glacial maximum. Earth and Planetary Science Letters 363, 133–43. doi: 10.1016/j.epsl.2012.12.026 CrossRefGoogle Scholar
LeGrande, AN and Schmidt, GA (2006) Global gridded data set of the oxygen isotopic composition in seawater. Geophysical Research Letters 33, L12604. doi: 10.1029/2006GL026011 CrossRefGoogle Scholar
Leung, JCH, Zhang, B, Gan, Q, Wang, L, Qian, W and Hu, ZZ (2022) Differential expansion speeds of Indo-Pacific warm pool and deep convection favoring pool under greenhouse warming. npj Climate and Atmospheric Science 5, 97. doi: 10.1038/s41612-022-00315-w CrossRefGoogle Scholar
Lisiecki, L and Raymo, M (2005) A Pliocene-Pleistocene stack of 57 globally distributed benthic δ18O records. Paleoceanography 20, PA1003. doi: 10.1029/2004PA001071 Google Scholar
Liu, Y, Lo, L, Shi, Z, Wei, KY, Chou, CJ, Chen, YC, Chuang, CK, Wu, CC, Mii, HS, Peng, Z, Amakawa, H, Burr, GS, Lee, SY, DeLong, KL, Elderfield, H and Shen, CC (2015) Obliquity pacing of the western Pacific Intertropical Convergence Zone over the past 282,000 years. Nat Commun 6, 10018. doi: 10.1038/ncomms10018 CrossRefGoogle ScholarPubMed
Lo, L, Shen, CC, Zeeden, C, Tsai, YH, Yin, Q, Yang, CC, Chang, TL, Su, YC, Mii, HS, Chuang, CK and Chen, YC (2022) Orbital control on the thermocline structure during the past 568 kyr in the Solomon Sea, southwest equatorial Pacific. Quaternary Science Reviews 295, 107756. doi: 10.1016/j.quascirev.2022.107756 CrossRefGoogle Scholar
Locarnini, RA, Mishonov, AV, Antonov, JI, Boyer, TP, Garcia, HE, Baranova, OK, Zweng, MM, Paver, CR, Reagan, JR, Johnson, DR, Hamilton, M and Seidov, D (2013) World Ocean Atlas 2013. Volume 1, temperature. In NOAA Atlas NESDIS, vol. 73 (eds S Levitus, A Mishonov & Technical), p. 40, Publication. doi: 10.7289/V55X26VD CrossRefGoogle Scholar
Lu, H, Yi, S, Liu, Z, Mason, JA, Jiang, D, Cheng, J, Stevens, T, Xu, Z, Zhang, E, Jin, L, Zhang, Z, Guo, Z, Wang, Y and Otto-Bliesner, B (2013) Variation of East Asian monsoon precipitation during the past 21 k.y. and potential CO2 forcing. Geology 41, 1023–26. doi: 10.1130/G34488.1 CrossRefGoogle Scholar
Lu, RY (2001) Interannual variability of the summer time North Pacific subtropical high and its relation to atmospheric convection over the Warm Pool. Journal of the Meteorological Society of Japan Ser II 79, 771–83.CrossRefGoogle Scholar
Lutgens, FK and Tarbuck, EJ (2001) The atmosphere (8th ed.). Upper Saddle River, NJ: Prentice Hall.Google Scholar
Mantsis, DF, Clement, AC, Broccoli, AJ and Erb, MP (2011) Climate feedbacks in response to changes in obliquity. Journal of Climate, 24(11), 28302845. doi: 10.1175/2010JCLI3986.1 CrossRefGoogle Scholar
Mantsis, DF, Lintner, BR, Broccoli, AJ, Erb, MP, Clement, AC and Park, H (2014) The response of large-scale circulation to obliquity-induced changes in meridional heating gradients. Journal of Climate, 27(14), 55045516. doi: 10.1175/JCLI-D-13-00526.1 CrossRefGoogle Scholar
McGee, D, Donohoe, A, Marshall, J and Ferreira, D (2014) Changes in ITCZ location and cross-equatorial heat transport at the Last Glacial Maximum, Heinrich Stadial 1, and the mid-Holocene. Earth and Planetary Science Letters 390, 6979. doi: 10.1016/j.epsl.2013.12.043 CrossRefGoogle Scholar
Merlis, TM, Schneider, T, Bordoni, S and Eisenman, I (2013) The tropical precipitation response to orbital precession. Journal of Climate 26, 2010–21. doi: 10.1175/jcli-d-12-00186.1 CrossRefGoogle Scholar
Mohtadi, M, Prange, M, Oppo, DW, De Pol-Holz, R, Merkel, U, Zhang, X, Steinke, S and Luckge, A (2014) North Atlantic forcing of tropical Indian Ocean climate. Nature 509, 7680. doi: 10.1038/nature13196 CrossRefGoogle ScholarPubMed
Nürnberg, D, Bijma, J and Hemleben, C (1996) Assessing the reliability of magnesium in foraminiferal calcite as a proxy for water mass temperatures. Geochimica et Cosmochimica Acta 60, 803–14. doi: 10.1016/0016-7037(95)00446-7 CrossRefGoogle Scholar
Pierrehumbert, RT (2000) Climate change and the Tropical Pacific: the sleeping Dragon Wakes. Proceedings of the National Academy of Sciences 97, 1355–58. doi: 10.1073/pnas.97.4.1355 CrossRefGoogle ScholarPubMed
Qian, F, Chang, F, Li, T, Li, A, Sun, H and Zhang, J (2023) Thermocline variability in the subtropical northwestern Pacific since the last deglaciation. Palaeogeography, Palaeoclimatology, Palaeoecology 612, 111379. doi: 10.1016/j.palaeo.2022.111379 CrossRefGoogle Scholar
Rachmayani, R, Prange, M and Schulz, M (2016) Intra-interglacial climate variability: model simulations of Marine Isotope Stages 1, 5, 11, 13, and 15. Climate of the Past 12, 677–95. doi: 10.5194/cp-12-677-2016 CrossRefGoogle Scholar
Raddatz, J, Nürnberg, D, Tiedemann, R and Rippert, N (2017) Southeastern marginal West Pacific Warm Pool sea-surface and thermocline dynamics during the Pleistocene (2.5–0.5 Ma). Palaeogeography, Palaeoclimatology, Palaeoecology 471, 144–56. doi: 10.1016/j.palaeo.2017.01.024 CrossRefGoogle Scholar
Regenberg, M, Nürnberg, D, Steph, S, Groeneveld, J, Garbe-Schönberg, D, Tiedemann, R and Dullo, WC (2006) Assessing the effect of dissolution on planktonic foraminiferal Mg/Ca ratios: evidence from Caribbean core tops. Geochemistry, Geophysics, Geosystems 7, Q07P15. doi: 10.1029/2005gc001019 CrossRefGoogle Scholar
Regenberg, M, Regenberg, A, Garbe-Schönberg, D and Lea, DW (2014) Global dissolution effects on planktonic foraminiferal Mg/Ca ratios controlled by the calcite-saturation state of bottom waters. Paleoceanography 29, 127–42. doi: 10.1002/2013pa002492 CrossRefGoogle Scholar
Rippert, N, Nürnberg, D, Raddatz, J, Maier, E, Hathorne, E, Bijma, J and Tiedemann, R (2016) Constraining foraminiferal calcification depths in the western Pacific warm pool. Marine Micropaleontology 128, 1427. doi: 10.1016/j.marmicro.2016.08.004 CrossRefGoogle Scholar
Rosenthal, Y, Holbourn, AE, Kulhanek, DK and the Expedition 363 Scientists (2018) Western Pacific Warm Pool. Proceedings of the International Ocean Discovery Program, 363: College Station, TX (International Ocean Discovery Program). doi: 10.14379/iodp.proc.363.107.2018.CrossRefGoogle Scholar
Rosenthal, Y, Oppo, DW and Linsley, BK (2003) The amplitude and phasing of climate change during the last deglaciation in the Sulu Sea, western equatorial Pacific. Geophysical Research Letters 30, 1428. doi: 10.1029/2002gl016612 CrossRefGoogle Scholar
Ruddiman, WF (2008) Earth’s Climate: Past and Future. 2nd ed. W. H. New York: Freeman and Company, 388 p.Google Scholar
Russell, JM, Vogel, H, Konecky, BL, Bijaksana, S, Huang, Y, Melles, M, Wattrus, N, Costa, K and King, JW (2014) Glacial forcing of central Indonesian hydroclimate since 60,000 y B.P. Proceedings of the National Academy of Sciences (U.S.A.) 111, 5100–05. doi: 10.1073/pnas.1402373111 CrossRefGoogle ScholarPubMed
Schmidt, MW and Spero, HJ (2011) Meridional shifts in the marine ITCZ and the tropical hydrologic cycle over the last three glacial cycles. Paleoceanography 26, PA1206. doi: 10.1029/2010pa001976 CrossRefGoogle Scholar
Schneider, T, Bischoff, T and Haug, GH (2014) Migrations and dynamics of the intertropical convergence zone. Nature 513, 4553. doi: 10.1038/nature13636 CrossRefGoogle ScholarPubMed
Scroxton, N, Burns, SJ, McGee, D, Hardt, B, Godfrey, LR, Ranivoharimanana, L and Faina, P (2017) Hemispherically in-phase precipitation variability over the last 1700 years in a Madagascar speleothem record. Quaternary Science Reviews 164, 2536. doi: 10.1016/j.quascirev.2017.03.017 CrossRefGoogle Scholar
Singarayer, JS, Valdes, PJ and Roberts, WHG (2017) Ocean dominated expansion and contraction of the late Quaternary tropical rainbelt. Scientific Reports 7, 9382. doi: 10.1038/s41598-017-09816-8 CrossRefGoogle ScholarPubMed
Spratt, RM and Lisiecki, LE (2016) A late pleistocene sea level stack. Climate of the Past 12(4), 10791092. doi: 10.5194/cp-12-1079-2016 CrossRefGoogle Scholar
Stott, L, Cannariato, K, Thunell, R, Haug, GH, Koutavas, A and Lund, S (2004) Decline of surface temperature and salinity in the western tropical Pacific Ocean in the Holocene epoch. Nature 431, 56–9. doi: 10.1038/nature02903 CrossRefGoogle ScholarPubMed
Tachikawa, K, Cartapanis, O, Vidal, L, Beaufort, L, Barlyaeva, T and Bard, E (2011) The precession phase of hydrological variability in the Western Pacific Warm Pool during the past 400 ka. Quaternary Science Reviews 30, 3716–27. doi: 10.1016/j.quascirev.2011.09.016 CrossRefGoogle Scholar
Tachikawa, K, Timmermann, A, Vidal, L, Sonzogni, C and Timm, OE (2014) CO2 radiative forcing and Intertropical Convergence Zone influences on western Pacific warm pool climate over the past 400ka. Quaternary Science Reviews 86, 2434. doi: 10.1016/j.quascirev.2013.12.018 CrossRefGoogle Scholar
Trenberth, KE, Stepaniak, DP and Caron, JM (2000) The global monsoon as seen through the divergent atmospheric circulation. Journal of Climate 13, 3969–93. doi: 10.1175/1520-0442(2000)013<3969:Tgmast>2.0.Co;2 2.0.CO;2>CrossRefGoogle Scholar
Voigt, A and Shaw, T (2015) Circulation response to warming shaped by radiative changes of clouds and water vapour. Nature Geoscience 8, 102–6. doi: 10.1038/ngeo2345 CrossRefGoogle Scholar
Wang, B, Wu, R and Li, T (2003) Atmosphere-warm ocean interaction and its impacts on Asian-Australian monsoon variation. Journal of Climate 16, 1195–211. doi: 10.1175/1520-0442(2003)16<1195:AOIAII>2.0.CO;2 2.0.CO;2>CrossRefGoogle Scholar
Wang, X, Auler, AS, Edwards, RL, Cheng, H, Ito, E and Solheid, M (2006) Interhemispheric anti-phasing of rainfall during the last glacial period. Quaternary Science Reviews 25, 3391–403. doi: 10.1016/j.quascirev.2006.02.009 CrossRefGoogle Scholar
Wang, Y, Cheng, H, Edwards, RL, Kong, X, Shao, X, Chen, S, Wu, J, Jiang, X, Wang, X and An, Z (2008) Millennial- and orbital-scale changes in the East Asian monsoon over the past 224,000 years. Nature 451, 1090–93. doi: 10.1038/nature06692 CrossRefGoogle ScholarPubMed
Wang, YJ, Cheng, H, Edwards, RL, An, ZS, Wu, JY, Shen, CC and Dorale, JA (2001) A high-resolution absolute-dated late Pleistocene monsoon record from Hulu Cave, China. Science 294, 2345–48. doi: 10.1126/science.1064618 CrossRefGoogle ScholarPubMed
Webster, PJ, Magana, VO, Palmer, TN, Shukla, J, Tomas, RA, Yanai, M and Yasunari, T (1998) Monsoon: processes, predictability, and prospects for prediction. Journal of Geophysical Research 103, 14,451510.CrossRefGoogle Scholar
Xiong, Z, Zhai, B, Algeo, TJ, Lu, Z, Li, T, Meyer, H, Jiang, F, Zhang, P, Qin, B, Gong, X, Wang Zand Jia, Q (2022) Intensified aridity over the Indo-Pacific Warm Pool controlled by ice-sheet expansion during the Last Glacial Maximum. Global and Planetary Change 217, 103952. doi: 10.1016/j.gloplacha.2022.103952 CrossRefGoogle Scholar
Yan, H, Wei, W, Soon, W, An, Z, Zhou, W, Liu, Z, Wang, Y and Carter, RM (2015) Dynamics of the intertropical convergence zone over the western Pacific during the Little Ice Age. Nature Geoscience 8, 315–20. doi: 10.1038/ngeo23795 CrossRefGoogle Scholar
Yan, XH, Ho, CR, Zheng, Q and Klemas, V (1992) Temperature and size variabilities of the Western pacific warm pool. Science 258, 1643–45.CrossRefGoogle ScholarPubMed
Yancheva, G, Nowaczyk, NR, Mingram, J, Dulski, P, Schettler, G, Negendank, JF, Liu, J, Sigman, DM, Peterson, LC and Haug, GH (2007) Influence of the intertropical convergence zone on the East Asian monsoon. Nature 445, 74–7. doi: 10.1038/nature05431 CrossRefGoogle ScholarPubMed
Yang, L, Zhou, W, Yan, H, Ma, X, Cheng, P, Hu, B, Shen, X, Zhao, H, Hou, Y, Lu, X and Liu, C (2024). Step-wise shifted rainbelt throughout the early last deglaciation. Global and Planetary Change 232, 104325. doi: 10.1016/j.gloplacha.2023.104325 CrossRefGoogle Scholar
Yang, Y, Zhang, L, Yi, L, Zhong, F, Lu, Z, Wan, S, Du, Y and Xiang, R (2023) A contracting intertropical convergence zone during the Early Heinrich Stadial 1. Nature Communications 14, 4695. doi: 10.1038/s41467-023-40377-9 CrossRefGoogle ScholarPubMed
Yin, QZ, Singh, UK, Berger, A, Guo, ZT and Crucifix, M (2014) Relative impact of insolation and the Indo-Pacific warm pool surface temperature on the East Asia summer monsoon during the MIS-13 interglacial. Climate of the Past 10, 1645–57. doi: 10.5194/cp-10-1645-2014 CrossRefGoogle Scholar
Yin, QZ, Wu, ZP, Berger, A, Goosse, H and Hodell, D (2021) Insolation triggered abrupt weakening of Atlantic circulation at the end of interglacials. Science 373, 1035–40. doi: 10.1126/science.abg1737 CrossRefGoogle ScholarPubMed
Yu, Z, Tang, X, Colin, C, Wilson, DJ, Zhou, X, Song, L, Chang, F, Zhang, S, Bassinot, F and Wan, S (2023) Millennial-scale precipitation variability in the Indo-Pacific region over the Last 40 Kyr. Geophysical Research Letters 50, e2022GL101646. doi: 10.1029/2022gl101646 CrossRefGoogle Scholar
Yuan, S, Chiang, H, Liu, G, Bijaksana, S, He, S, Jiang, X, Imran, A, Wicaksono, S and Wang, X (2023) The strength, position, and width changes of the intertropical convergence zone since the Last Glacial Maximum. Proceedings of the National Academy of Sciences 120, e2217064120. doi: 10.1073/pnas.2217064120 CrossRefGoogle ScholarPubMed
Zhang, P, Xu, J, Holbourn, A, Kuhnt, W, Beil, S, Li, T, Xiong, Z, Dang, H, Yan, H, Pei, R, Ran, Y and Wu, H (2020) Indo-Pacific hydroclimate in response to changes of the intertropical convergence zone: discrepancy on precession and obliquity bands over the last 410 kyr. Journal of Geophysical Research: Atmospheres 125, e2019JD032125. doi: 10.1029/2019jd032125 CrossRefGoogle Scholar
Zhang, P, Xu, J, Holbourn, A, Kuhnt, W, Xiong, Z and Li, T (2022) Obliquity induced latitudinal migration of the intertropical convergence zone during the past ∼410 kyr. Geophysical Research Letters 49, e2022GL100039. doi: 10.1029/2022gl100039 CrossRefGoogle Scholar
Zhang, S, Li, T, Chang, F, Yu, Z and Wang, H (2015) The relationship between the Tropical Pacific precipitation and the ITCZ variation since MIS 6. Quaternary Sciences 35, 390400. doi: 10.11928/j.issn.1001-7410.2015.02.14 Google Scholar
Zhang, S, Yu, Z, Gong, X, Wang, Y, Chang, F, Lohmman, G, Qi, Y and Li, T (2021) Precession cycles of the El Niño/Southern oscillation-like system controlled by Pacific upper-ocean stratification. Communications Earth & Environment 2, 239. doi: 10.1038/s43247-021-00305-5 CrossRefGoogle Scholar
Figure 0

Figure 1. The Indo-Pacific Warm Pool and regional circulation patterns, alongside the modern climatology of the study area. (a) Map of the IPWP displaying modern annual sea surface temperatures (SST at 0 m; World Ocean Atlas 2013 dataset; Locarnini et al. 2013) of the studied Site U1486 (yellow star) and other relevant sites discussed in the study (white circles) is shown. Bright yellow shading lines denote the mean positions of the Intertropical Convergence Zone (ITCZ) in July and January, respectively (Lutgens & Tarbuck, 2001). Black dotted lines represent annual mean 28°C SST isotherms. Key ocean currents include the North Equatorial Current (NEC), North Equatorial Counter Current (NECC), South Equatorial Current (SEC), Kuroshio Current (KC) and Mindanao Current (MC). (b) The Southern Oscillation Index (SOI, 5-point running average) between 1980 and 2020 (https://www.cpc.ncep.noaa.gov/data/indices/soi) together with the monthly rainfall anomaly over the Site U1486 source area (2.5° spatial resolution centered on 2°S, 144°E, 11-point running average, https://climatedataguide.ucar.edu/climate-data/gpcp-monthly-global-precipitation-climatology-project). (c) Depiction of mean rainfall rate and the 1000 hPa wind field in January (left) and July (right). Rainfall data sourced from https://psl.noaa.gov/data/gridded/data.cmap.html, while land wind data obtained from http://iridl.ldeo.columbia.edu/expert/ds:/SOURCES/.NOAA/.NCEP-NCAR/.CDAS-1.

Figure 1

Figure 2. Age model of Site U1486. (a) Benthic foraminifer C. wuellerstorfi δ18O (blue line) and the global benthic foraminiferal δ18O stack LR04 (Lisiecki & Raymo, 2005) (red line). Black triangles indicate tie points for δ18OC. wuellerstorfi. (b) Depth-age model (green line) and sedimentation rates (black line) of sediment Site U1486. Gray vertical shadings indicate even-numbered Marine Isotope Stages (MIS).

Figure 2

Figure 3. Planktonic foraminiferal proxy records of Site U1486. (a) G. ruber δ18O, (b) G. ruber Mg/Ca-based SST, (c) the ice volume corrected-seawater δ18O (δ18Osw-ivc) and (d) C. wuellerstorfi δ18O (black), superimposed is the global ice volume δ18Osw (blue; Bintanja et al. 2005). Gray vertical shadings indicate even-numbered marine isotope stages (MIS), while colored shadings for individual records represent their respective error ranges (see Sections 2.2 and 2.3). (e) Spectral (left) and wavelet transform (right) analyses of the δ18Osw-ivc record of Site U1486. The green line indicates the precession cycle.

Figure 3

Figure 4. The orbital-scale precipitation pattern in the Indo-Pacific Warm Pool over the past 450 kyr. (a) The variation in precession (Laskar et al. 2004). (b) GISS_ModelE2-R simulated δ18Osw-ivc records of the IPWP (Jian et al. 2022). (c, d, e) The δ18Osw-ivc records of cores MD98-2162, SO18480-3 and U1483 in the eastern tropical Indian Ocean (Jian et al. 2022; Zhang et al. 2022). (f, g) The δ18Osw-ivc records of cores MD06-3047B and MD06-3067 in the northern sector of western equatorial Pacific (Bolliet et al. Bolliet et al., 2011; Jia et al. Jia et al., 2018). (h) The δ18Osw-ivc records of core MD05-2925 (Lo et al. 2022). (i) The ln(Ti/Ca) record of core MD05-2920 (Tachikawa et al. 2014). (j) The δ18Osw-ivc record of Site U1486, superimposed by the July–August insolation at the equator (black) (Laskar et al. 2004). The overlain, bold, dark-colored lines represent the precession band-pass filtered results. Band-pass filter with a central frequency of 0.043 kyr−1 and a bandwidth of 0.01 kyr−1. Vertical grey bars indicate lower precession values.

Figure 4

Figure 5. Comparing the precipitation records in the IPWP and East Asia with the ENSO-like proxy and the migration of the ITCZ. (a) The variation in precession (Laskar et al. 2004). (b) The speleothem δ18O records in China (Cheng et al. 2016), superimposed on the summer insolation at 15°N (Laskar et al. 2004). (c) The summer interhemispheric tropical insolation gradient, determined by the variation in June insolation between 23° N and 23° S (Laskar et al. 2004). (d) CESM-simulated Niño3 SST of DJF (Zhang et al. 2021). (e) Differences between SST and thermocline water temperature (ΔT) for Site U1486. (f) GISS_ModelE2-R simulated δ18Osw-ivc records of the IPWP (Jian et al. 2022). (g) The δ18Osw-ivc records of Site U1486 superimposed by the July–August insolation at the equator (black) (Laskar et al. 2004). The records are overlaid by the precession band-pass filtered output (bold, dark-colored lines). Band-pass filter with a central frequency of 0.043 kyr−1 and a bandwidth of 0.01 kyr−1. Vertical grey bars indicate lower precession values (high Northern Hemisphere summer insolation). (h) Phase relationships among U1486 δ18Osw-ivc, simulated IPWP δ18Osw-ivc, speleothem δ18O and the precession parameter, illustrated through coherences (left) and phase angles (right).

Figure 5

Figure 6. Effects of the ITCZ migration and width on the precipitation variations in the IPWP on the precession band. A. The records of ITCZ migration and width variation: (a) The variation in precession (Laskar et al. 2004). (b, c, d) Precession bandpass filtering curves of δ18Osw-ivc records in northern marginal ITCZ (core MD06-3047B; Jia et al. Jia et al., 2018) and southern marginal ITCZ (core MD05-2925 and Site U1483; Lo et al. 2022; Zhang et al. 2022). (e) Gradients of δ18Osw−ivc between Site U1486 and core MD06-3047B (Jia et al. Jia et al., 2018). (f) Gradients of δ18Osw−ivc between Site U1486 and core MD05-2925 (Lo et al. 2022). (g) The summer interhemispheric tropical insolation gradient, determined by the variation in June insolation between 23° N and 23° S (Laskar et al. 2004). Vertical grey bars indicate lower precession values. B. Schematic diagrams showing ITCZ variability during low-precession (left) and high-precession (right) intervals. The pink shaded band indicates the position and size of the ITCZ, referring to the belt of the modern ITCZ’s seasonal migration.

Supplementary material: File

Qian et al. supplementary material

Qian et al. supplementary material
Download Qian et al. supplementary material(File)
File 2.3 MB