Supplementary Information for: The effect of sea level on glacial Indo-Pacific climate Pedro DiNezio1∗ & Jessica Tierney2 1 International Pacific Research Center, University of Hawaii, 2 ∗ Woods Hole Oceanographic Institution, To whom correspondence should be addressed; E-mail: [email protected]. 1. 1.1. Proxy data Proxy database and selection criteria We conducted a literature search, a perusal of publicly available data in paleoclimate archives (NOAA’s National Climatic Data Centre’s Paleoclimatology database, http://www.ncdc.noaa.gov/ paleo/paleo.html, and Pangaea, http://www.pangaea.de) and contacted some authors directly to collect proxy data for our synthesis. As stated in the Methods Summary, we applied the following criteria towards our proxy data selection: 1) the proxy used be interpreted to reflect hydroclimate, 2) the data must contain information concerning hydroclimate during both the Last Glacial Maximum (26.5–19 ka, corresponding to the timing of the sea-level lowstand 1 ) and the Late Holocene (0–4 ka) and 3) the proxy data site be located within our target region of 25◦ S–20◦ N, 25–170◦ E. The latitudinal bounds of the target region were chosen to include as much of the Indo-Pacific region as possible while excluding the northern tropics, where the presence of the glacial ice sheets likely affects hydroclimate. The longitudinal bounds were informed by the geographical extent of the Indian Ocean/Indo-Pacific Warm Pool (IPWP), and also reflect the lack of proxy data from the central tropical Pacific east of 170◦ E. We did not impose any criteria regarding the chronological control of the proxy archives in order to facilitate the inclusion of as many proxy records as possible, and under the assumption that in most cases, sedimentation rates and chronology are sufficient to constrain the data to within our broad target windows. Nevertheless, certain archives may have dating uncertainties that exceed our target windows. To test the sensitivity of our results to chronological uncertainties, we restricted our collection of proxy data to records that contain one or more radiocarbon, U/Th, or well-constrained OSL dates near to or during the LGM and repeated the analyses (Fig. S1). Imposing this restriction reduces the number of our terrestrial proxy sites from 53 to 37 (32 sites with robust data), and the marine sites from 54 to 26 (25 sites with robust data). For the terrestrial proxies, we get the same result with the reduced matrix: HadCM3 1 outperforms the other models. For the marine proxies, the pattern of kappa results (Fig. S2) is similar to the pattern when including all proxies (Fig. 3) but the kappa values are no longer statistically significant because the number of proxy sites (N) is now reduced to 25. The severe depopulation of the SSS proxy matrix reflects the fact that many of the marine records rely on wiggle-matching benthic δ 18 O data at their site to the global benthic δ 18 O stack (e.g., LR04 2 ) as a means of age control rather than relying on radiocarbon. It could be argued that this technique is sufficient to constrain LGM data (given sufficient sampling rates) as the LGM lowstand (26–19ka) is demarcated clearly by enriched benthic δ 18 O values, highlighting the invariably subjective nature of identifying records as “better-dated” and “poorly-dated” by any a priori metric. In sum, this test suggests that our results are largely insensitive to chronological differences between the records, but, in the case of the salinity proxies, it is clear that a critical number of sites is needed to get meaningful results. This latter observation, along with the inherent subjectivity in imposing a threshold for “well-dated” and “poorly-dated,” justifies our choice to use a generous proxy screening criteria with regards to chronological control. In addition, we emphasize that the core finding of the paper, which is that terrestrial and marine proxies – each subject to independent types of dating uncertainties and proxy noise – both point towards the same pattern of hydroclimatic change. Therfore the response of Indo-Pacific hydroclimate to the exposure of the Sunda Shelf appears to be robust. For many types of proxies, conversion of the proxy to a quantitative climate metric (precipitation, salinity) is difficult or else may carry large uncertainties and can complicate multi-proxy synthesis 3 . Thus, for the present synthesis we chose to place proxy data into three basic categories: evidence for drier or saltier conditions during the LGM (as compared to the Late Holocene), no evidence for change during the LGM, or evidence for wetter or fresher conditions during the LGM. This categorical classification makes the conservative assumption that the proxies can at the least give a robust indication of the direction of climatic change. We used 2 the interpretations of the authors of the individual proxy record studies to determine which category the data fall into, which in some cases is based on quantitative comparisons of the mean and variance of LGM and Late Holocene data, and in other cases is necessarily qualitative (e.g., “increased dune activity implies drier conditions relative to present at the LGM”). Tables S1 and S2 list the proxy data used in the synthesis, including their locations, source references, and the corresponding categorical classification of hydroclimatic or salinity change during the LGM relative to late Holocene conditions. 1.2. Heinrich Events Our target LGM window includes Heinrich Event 2 (ca. 23 ka), a millennial-scale climate event triggered by a sudden advance of sea-ice in the North Atlantic 4 . We also recognize that poor dating on some archives may conflate late-glacial climate with changes associated with another Heinrich Event, Heinrich Event 1 (ca. 17 ka). Although Heinrich Events are North Atlantic phenomena, it is well-known that tropical climate responded to these perturbations (e.g., 5 ). Thus, Heinrichs 2 and 1 have the potential to overprint the proxy signature of LGM conditions. In many continuous core sequences, the transition into Heinrich 1 is apparent and clearly distinguishable from LGM conditions. Heinrich 2 poses more of a challenge, but it is unlikely that in low-resolution, highly-integrated archives H2 would dominate the signal of mean late-glacial conditions (although there may be site-specific exceptions, c.f. the East African discussion below). In addition, the expected climatic response of a Heinrich Event – dry conditions in the northern tropics counteracted by wet conditions in the southern tropics (c.f. 6 ) – is not apparent in the spatial patterns of our proxy synthesis (Figs. 2a and 3a). 1.3. Proxy data in key regions Here we provide brief overviews of proxy evidence for LGM conditions in areas that are particularly diagnostic of a Sunda Shelf effect on tropical hydroclimate. 3 1.3.1. East Africa An early synthesis of proxy lake level data in East Africa suggested that some lakes in the region may have experienced highstands during the LGM 7 . Subsequently, some argued that the older lake chronologies are poorly dated and/or that the inferred LGM highstands likely occurred before the LGM, therefore the entirety of the East African region was drier than present 8 . However, in the last decade, well-dated sedimentary archives including cores recovered from Lake Challa (located just to the East of Kilimanjaro) 9 and near the Zambezi River delta 10, 11 as well as new proxy data from existing archives 12 provide well-constrained evidence of wetter conditions along the eastern and southeastern African coast, in agreement with some of the older pollen and lake stratigraphies from central Kenya 7 . Thus, we consider the older East African proxy data “as-is”, e.g., in keeping with its original proxy interpretation. Two sites in southeastern Africa contain conflicting evidence or interpretations for hydroclimatic change. Lake Malawi, a large rift lake near 12◦ S, has traditionally been interpreted to be lower during the LGM on the basis of diatom distributions and carbon isotopes 8, 13 . However, organic matter composition, hydrogen isotopes and mineralogical evidence suggests minimal drying/no change 14, 15, 16 . The proximity of Lake Masoko, which experienced a highstand during the LGM, has also promoted debate concerning the fidelity of diatoms as lake level indicators under glacial conditions 12 . Given the controversy and disagreement between proxies, we treat this site as “not robust”, and it is not used in the kappa calculations. The second site of interest is Lake Chilwa, in which an ancient shoreline was dated by the OSL technique to ca. 26-21 ka 17 . As this spans the majority of the LGM, we originally considered this evidence for wetter conditions at the time. However, the authors attribute this highstand to Heinrich Event 2 and do not consider it representative of LGM conditions 17 , therefore, we also treat this site as not robust. Readers are also referred to 18 for further discussion of evidence for wet conditions during the 4 LGM in East Africa. 1.3.2. Western Indian Ocean/Arabian Sea The majority of the salinity proxy data from the Arabian Sea come from the study of 19 and are based on δ 18 Oseawater reconstructions using paired foraminiferal δ 18 O and Mg/Ca measurements. While translation of δ 18 Oseawater to absolute salinity values during the LGM is complicated by the changing δ 18 Oseawater –salinity relationship20 , the directionality of the proxy is not affected. 19 find substantial evidence for fresher conditions throughout the Arabian Sea, and the extent of this freshening is further corroborated by 21 , who found abnormally light δ 18 O values in foraminifera in the western Indian Ocean. 1.3.3. Northern Australia In agreement with the individual studies used in our synthesis, regional analyses of proxy data in the northern Australian region indicate widespread drying relative to present-day conditions 22, 23 . There is one proxy record of paleo-floods in plunge pools that suggests a more active hydrological cycle 24 and this site is included in our proxy synthesis. There is no indication, however, of a regionally-coherent wettening that would be typical of the “wet-get-drier” (thermodynamic) response. 1.3.4. Western Pacific Fresher conditions during the western Pacific have long been recognized as a general feature of foraminiferal-based investigations of paleoclimate in the region (e.g., thesis of δ 18 Oseawater reconstructions 27 25, 26 ). A recent syn- highlights the fresh conditions that prevailed in the South China Sea, and at least two sites indicate that the freshening extended far into the western/central Pacific 28, 29 . 5 1.4. Proxy search radius The proxy data are not distributed equally in space and some proxy data sites are located geographically close to one another. To avoid over-representing well-sampled regions in the model-proxy comparison, we applied a 150 km search radius to each proxy site and combined proxy information that fell within that distance of each other. The 150 km search radius is based on observational studies showing that the spatial correlation of precipitation intensity falls to an insignificant level beyond a separation distance between 100 and 200 km 30, 31, 32 . Spatial data points that incorporate more than one proxy are designated by triangles in Figs. 2a and 3a of the main text. Triangles in which the proxies agree on the sign of change are plotted in the corresponding color. Triangles in which the combined proxies disagree are plotted as black, are treated as “not robust,” and are not used in the kappa comparisons with the model simulations. Similarly, we mark sites where multiple proxies disagree and/or the interpretation is vague as “not robust” (e.g. see the East Africa section above). 2. Paleoclimate simulations We analyze changes in the climate of the IPWP during the LGM – relative to pre-industrial (PI) conditions – in an ensemble of climate simulations performed with twelve ocean-atmosphere climate models. The simulations were performed by modeling centers from the USA, Japan, UK, Germany, France, China and are coordinated by the Paleoclimate Modeling Intercomparison Project (PMIP) 33 . Five LGM experiments were performed and coordinated by PMIP in its second phase (PMIP2), one simulation was performed by NOAA/GFDL using the CM2.1 model 34 following PMIP2’s experimental protocol, and the remaining six simulations were performed and coordinated by PMIP in its third phase (PMIP3). Table S3 lists the models used, their resolution, and associated references. Further detailed descriptions of the models can be found on the PMIP2 website (http://pmip2.lsce.ipsl.fr/pmip2/) and the Program for Climate Model Diagnosis and Intercomparison (PCMDI) website 6 (http://www-pcmdi.llnl.gov/ipcc/about ipcc.php). All of the LGM (“21k oa” for PMIP2, “lgm” for PMIP3) and PI (“0k oa” for PMIP2, “piControl” for PMIP3) simulations were run with coordinated forcing and boundary conditions as described in the Methods Summary. However, the LGM simulations coordinated by PMIP2 and PMIP3 used slightly different reconstructions of the ice sheet topography and snow cover extent. PMIP2 used the ICE-5G V1 35 reconstruction 33 , while PMIP3 used a blend of the ICE-6G 36 , MOCA 37 , and ANU 38 reconstructions (https://pmip3.lsce.ipsl.fr/wiki/doku.php/pmip3:design :pi:final:icesheet). These differences are important for the simulation of mid-latitude circulation, but we see no evidence that they have a first-order effect on IPWP climate. 2.1. Wet-get-drier response In warmer (colder) climates atmospheric moisture tends to increase (decrease), governed by the Clausius-Clapeyron (C-C) equation. Regions where there is climatological moisture convergence, such as the IPWP, will tend to have increased (decreased) moisture convergence in response to warming (cooling), resulting in an enhanced (weakened) pattern of evaporation minus precipitation (E – P) 39, 40 . The changes in moisture convergence can lead to changes in either precipitation (P) or evaporation (E). This simple balance is relevant over the oceans, where relative humidity is nearly constant, as well as over humid land regions. Over arid or semiarid land surfaces, relative humidity can exhibit large changes and furthermore, the lack of an unlimited water vapor supply constrains moisture convergence to remain small. These are regions where the effect of the thermodynamic drying and associated “wet-get-drier” response are negligible. We use the annual-mean E – P estimated by the ERA40 reanalysis 41 for the 1979-2001 period to compute the changes in E – P expected for the LGM from the effect of thermodynamic drying on moisture convergence. First, we assume that the change in moisture will drive changes in P because E – P is dominated by P over the IPWP. We compute the change in P by multiplying 7 the annual-mean E – P with the C-C scaling (7%/K) and the models’ ensemble-mean tropical mean cooling (-2.5 K). We normalize the P response by the spatial standard deviation over the tropics (25◦ S-25◦ N) in order to define regions with negligible “wet-get-drier” response, such as land regions with limited evapotranspiration. The normalization of E – P leads to wet and dry regions with percentage changes in P that are larger than the 17.5% predicted by C-C for a 2.5K cooling. This allows us to then vary the threshold for defining wet-dry conditions and explore the sensitivity of the model-proxy agreement. We follow a similar procedure to estimate the signature of this mechanism on SSS, using annual-mean SSS values from 1950-2009 climatology 42 . We first remove the tropical-mean SSS, as changes in E – P drive deviations in SSS from the tropical mean (SSS∗ ). We then multiply SSS* by the C-C scaling (7%/K) and the models’ ensemble-mean tropical mean cooling (-2.5K). We acknowledge that SSS is directly driven by E – P but also reflects changes in ocean advection, which our simple calculation neglects. 3. Sunda Shelf mechanism The Sunda shelf – the areas of the Gulf of Thailand, the South China Sea, and the Java Sea that were exposed due to lowered sea level during the LGM – cools severely in the HadCM3 and GFDL-CM2.1 LGM simulations (Fig. S3). This anomalous surface cooling suppresses convection either by exceeding the threshold for convection or by generating anomalous temperature gradients with the surrounding ocean, favoring convection over the ocean. As a result, HadCM3 (Fig. 4a) and GFDL-CM2.1 (Fig. S4a) simulate large anomalous subsidence over the IPWP. However, we note that while HadCM3 and GFDL-CM2.1 simulate reductions in ascending motion over the Sunda Shelf of a similar magnitude, the corresponding changes in precipitation are much larger in HadCM3 than in GFDL-CM2.1 (Fig. 2b). Similarly, CCSM4.0 and GISS-E2-R also simulate strong cooling over the Sunda Shelf (Fig. S3), yet these models do not simulate large reductions in convection. These differences may point towards a role for deep convection parametrizations or land surface feedbacks; indeed, carbon isotopic data 8 suggest that savanna may have replaced forest cover during the LGM in at least the northern half of the Shelf 43 . However, a full diagnosis of these effects requires further simulations and thus is outside the scope of this synthesis paper. The remaining models simulate weaker cooling and convective response over the Sunda Shelf 44 (Fig. S3 and Fig. 4). MIROC3.2 used a present-day land mask in the LGM experiment, thus effectively not representing the effect of lower sea level on the geography of the Maritime Continent. This model does not simulate enhanced cooling over the Sunda Shelf or a reduction in ascending motion over exposed land areas (Fig. S4b), supporting the idea that the weakening of convection over the Sunda Shelf simulated by HadCM3 and GFDL-CM2.1 is a response to the changes in land-sea distribution. Moreover, MIROC3.2 simulates a stronger Pacific Walker circulation (Fig. S4b) consistent with the mechanism weakening the Walker circulation in response to global warming 44 . This strongly suggests that in the absence of land-sea distribution changes over the Maritime continent, the Indo-Pacific Walker circulation responds via the same mechanism to both cooling and warming. In sum, both the exposure of land and its enhanced cooling appear to be required for convection to weaken. This could explain why a previous modeling study that did not specify LGM cooling nor enhanced cooling over land 45 obtained opposite results; i.e. enhanced convection over the Sunda shelf. In the absence of greenhouse gas- or ice sheet-driven cooling, the exposed land warms (due to its reduced thermodynamic damping) favoring convection there. A more comprehensive and systematic explanation of the Sunda shelf mechanism requires a set of controlled simulations exploring the role of different convection schemes, ocean-atmosphere coupling, and land surface parameterizations. 3.1. Effect of the Sahul Shelf The exposure of the land bridge connecting northern Australia with New Guinea, the Sahul shelf, could also have local and remote impacts on hydroclimate. Not all the models simulate 9 drying/reduced convection over the the Sahul shelf (Fig. 2b). However, each model simulates the same type of response over both Sunda and Sahul, reflecting the different sensitivities of tropical convection over exposed land areas simulated by each model. HadCM3 is one of the few models simulating enhanced drying over the Sahul shelf (Fig. 2b), however the changes are not as large as those over the Sunda Shelf. The exposure of the Sahul Shelf could feasibly drive regional responses, for instance the large-scale drying across northern Australia 46 . However, the spatial patterns of mid-tropospheric vertical velocity change (Fig. S5) show a dominant role for the Sunda shelf both reducing convection and associated drying of the IPWP and driving remote responses, e.g. over East Africa. 4. Seasonality Precipitation and salinity proxies do not necessarily record annual-mean conditions. However, the diversity of proxies used in this study – many of which will have a different seasonal bias due to local oceanography and ecology – minimizes the possibility of a systematic seasonality bias in our categorical synthesis maps. Furthermore, to the extent that the mechanisms have a dominant annual-mean response, even a proxy with a seasonal bias will capture the signature of the annual-mean response. We investigate this issue by assessing the proxy-model agreement for changes in precipitation and SSS over different seasons. We find that the changes in precipitation during DJF and JJA simulated by HadCM3 exhibit the same large-scale patterns of a dry IPWP and wet eastern Africa as seen in the annual mean (Fig. S6). IPSL-CM4, in contrast, simulates drying of the IPWP that migrates towards the northern (southern) hemisphere in DJF (JJA) (Fig. S7). The JJA pattern shows statistically significant agreement with the terrestrial proxies (κ = 0.35). However, the JJA drying of the IPWP simulated by IPSL-CM4 does not translate into a pattern of SSS that shows good agreement with the marine proxies because this model does not simulate a salty Bay of Bengal nor fresh western Pacific, neither seasonally nor in the annual mean. All other models do not show statistically significant changes for DJF or JJA. In the present day climate, rainfall and SSS exhibit a marked seasonal con10 trast between the northern and southern hemispheres associated with the monsoons (Fig. S8), yet their seasonality is not as strong in the deep tropics, especially over the equatorial Indian Ocean, Sumatra, and equatorial East Africa. These are the same locations where the Sunda shelf mechanism, as simulated by HadCM3, has the strongest signals. Conversely, an enhancement of the present-day seasonal contrast (Fig. S8) does not appear to explain the patterns of rainfall and SSS change inferred from the proxies, especially over eastern Africa, where proxies show an east-west gradient at the LGM, but present-day seasonality exhibits a north-south contrast. Thus, while it is likely that some of the off-equatorial drying over southeast Asia and northern Australia suggested by the proxies could be due to changes in the monsoons, the Sunda shelf mechanism is the most parsimonious explanation for the changes over the equatorial Indian Ocean and Maritime Continent. In sum, seasonality does not affect our primary result that HadCM3 provides the best match for the proxy data. 5. 5.1. Proxy-Model comparison Representation of model data To best compare the simulations with the information provided by the terrestrial precipitation proxies, we plot each model’s precipitation change as a percentage of the annual mean precipitation in the preindustrial climate (Fig. 1b). For salinity, we use absolute units because unlike precipitation, the small range of global ocean salinities constrains the percentage changes to values of ca. 5% or less. As the salinity proxies are corrected for the global change in mean ocean salinity during the LGM due to the presence of the ice sheets (for SSS transfer functions, 1 psu, and for δ 18 Oseawater , 1h, 47 ), we also remove the global mean salinity change (∼1 psu) from the salinity fields of the model data for comparison. Note, however, that while all PMIP3 LGM simulations included this 1 psu change as an initial boundary condition, among the PMIP2 LGM simulations only CCSM3.0 and GFDL CM2.1 had a 1 psu salinity increase applied. Therefore, we only removed the mean ocean salinity change of 1 psu from 11 those models where it had been applied. Last, to compare the models with the proxies, we placed the model simulation data into the same three-category system, varying the threshold over which the model data is considered wetter/drier or fresher/saltier from 1–40% and 0.1-1 psu, respectively (see Figs. 2c and 3c). 5.2. Weighted Cohen’s κ As described in the Methods, the Cohen’s κ statistic 48 is a measure of categorical data agreement between two “raters” (in our case the proxies vs. each model) who classify N items (the locations with proxy data) into C mutually exclusive categories (the hydroclimate response defined either as drier, unchanged, or wetter; or saltier, unchanged, or fresher). For instance, the categorical data for HadCM3 when the precipitation cut-off is 13% (Figure 1b) is: Model Proxy Drier Unchanged Wetter Drier 22 10 1 Unchanged 3 4 0 Wetter 1 1 5 and for FGOALS1.0g with a 5% cut-off: Model Proxy Drier Unchanged Wetter Drier 20 5 8 Unchanged 1 1 5 Wetter 4 1 2 Cohen’s κ is then defined as the observed fractional agreement (po ) relative to the probability of random agreement (pe ): κ= po − pe 1 − pe 12 where po is the fractional agreement among the raters (the sum of the diagonal elements divided by the number of items, N ) and pe is the probability that the raters agree due to random chance (computed from the observed data as the frequency of occurrence of each category, i.e. the product of the sum of the respective rows normalized by N ). As noted in the Methods Summary, we use a modified version of this Cohen’s κ, the weighted Cohen’s κ 49 , in which multiplying the data matrix by a weight matrix penalizes models for a total miss (e.g., drier when it should be wetter, weight = 0) more than a near miss (e.g., drier when it should be no change, weight = 0.5). 13 References 1. Clark, P. U. et al. The Last Glacial Maximum. Science 16, 710–714 (2009). 2. Lisiecki, L. & Raymo, M. A Pliocene-Pleistocene stack of 57 globally distributed benthic δ 18 O records. Paleoceanography 20, PA1003 (2005). 3. Waelbroeck, C. et al. Constraints on the magnitude and patterns of ocean cooling at the Last Glacial Maximum. Nature Geoscience 2, 127–132 (2009). 4. Hemming, S. Heinrich events: Massive late Pleistocene detritus layers of the North Atlantic and their global climate imprint. Reviews of Geophysics 42, RG1005 (2004). 5. Stager, J., Ryves, D., Chase, B. & Pausata, F. Catastrophic drought in the Afro-Asian monsoon region during Heinrich event 1. Science 331, 1299–1302 (2011). 6. Chiang, J. & Friedman, A. Extratropical Cooling, Interhemispheric Thermal Gradients, and Tropical Climate Change. Annual Review of Earth and Planetary Sciences 40, 383– 412 (2012). 7. Street, F. A. & Grove, A. T. Global maps of lake-level fluctuations since 30,000 yr BP. Quat. Res. 12, 83–118 (1979). 8. Barker, P. A. & Gasse, F. New evidence for a reduced water balance in East Africa during the Last Glacial Maximum: implication for model-data comparison. Quat. Sci. Rev. 22, 823–837 (2003). 9. Verschuren, D. et al. Half-precessional dynamics of monsoon rainfall near the East African equator. Nature 462, 637–641 (2009). 10. Castañeda, I. et al. Continental and sea surface temperature variability in southeast Africa (Zambezi River region) since MIS 3. In AGU Fall Meeting Abstracts, vol. 1, GC51C– 0763 (2010). 11. Schefuß, E., Kuhlmann, H., Mollenhauer, G., Prange, M. & Pätzold, J. Forcing of wet phases in southeast Africa over the past 17,000 years. Nature 480, 509–512 (2011). 12. Garcin, Y., Vincens, A., Williamson, D., Guiot, J. & Buchet, G. Wet phases in tropical southern Africa during the last glacial period. Geophys. Res. Lett. 33, L07703 (2006). 13. Castañeda, I. S., Werne, J. P. & Johnson, T. C. Wet and arid phases in the southeast African tropics since the Last Glacial Maximum. Geology 35, 823–826 (2007). 14. Filippi, M. & Talbot, M. The palaeolimnology of northern Lake Malawi over the last 25 ka based upon the elemental and stable isotopic composition of sedimentary organic matter. Quaternary Science Reviews 24, 1303–1328 (2005). 15. Cohen, A. S. et al. Ecological consequences of early Late Pleistocene megadroughts in tropical Africa. Proc. Nat. Acad. Sci. U.S.A. 104, 16422–16427 (2007). 14 16. Konecky, B. et al. Atmospheric circulation patterns during late Pleistocene climate changes at Lake Malawi, Africa. Earth and Planetary Science Letters 312, 318–326 (2011). 17. Thomas, D., Bailey, R., Shaw, P., Durcan, J. & Singarayer, J. Late Quaternary highstands at Lake Chilwa, Malawi: frequency, timing and possible forcing mechanisms in the last 44 ka. Quaternary Science Reviews 28, 526–539 (2009). 18. Tierney, J. E., Russell, J. M., Sinninghe Damsté, J. S., Huang, Y. & Verschuren, D. Late Quaternary behavior of the East African monsoon and the importance of the Congo Air Boundary. Quat. Sci. Rev. 30, 798–807 (2011). 19. Dahl, K. & Oppo, D. Sea surface temperature pattern reconstructions in the Arabian Sea. Paleoceanography 21, PA1014 (2006). 20. LeGrande, A. & Schmidt, G. Water isotopologues as a quantitative paleosalinity proxy. Paleoceanography 26, PA3225 (2011). 21. Kiefer, T., McCave, I. & Elderfield, H. Antarctic control on tropical Indian Ocean sea surface temperature and hydrography. Geophysical Research Letters 33, L24612 (2006). 22. De Deckker, P., Tapper, N. & Van Der Kaars, S. The status of the Indo-Pacific Warm Pool and adjacent land at the Last Glacial Maximum. Global and Planetary Change 35, 25–35 (2002). 23. Hesse, P., Magee, J. & van der Kaars, S. Late Quaternary climates of the Australian arid zone: a review. Quaternary International 118, 87–102 (2004). 24. Nott, J. & Price, D. Waterfalls, floods and climate change: evidence from tropical Australia. Earth and Planetary Science Letters 171, 267–276 (1999). 25. Lea, D., Pak, D. & Spero, H. Climate impact of late Quaternary equatorial Pacific sea surface temperature variations. Science 289, 1719–1724 (2000). 26. Rosenthal, Y., Oppo, D. & Linsley, B. The amplitude and phasing of climate change during the last deglaciation in the Sulu Sea, western equatorial Pacific. Geophysical Research Letters 30, 1428 (2003). 27. Xu, J., Kuhnt, W., Holbourn, A., Regenberg, M. & Andersen, N. Indo-Pacific Warm Pool variability during the Holocene and Last Glacial Maximum. Paleoceanography 25, PA4230 (2010). 28. Russon, T. et al. The mid-Pleistocene transition in the subtropical southwest Pacific. Paleoceanography 26, PA1211 (2011). 29. Sagawa, T., Yokoyama, Y., Ikehara, M. & Kuwae, M. Shoaling of the western equatorial Pacific thermocline during the last glacial maximum inferred from multispecies temperature reconstruction of planktonic foraminifera. Palaeogeography, Palaeoclimatology, Palaeoecology 346–347, 120–129 (2012). 15 30. Dai, A., Fung, I. Y. & Del Genio, A. D. Surface Observed Global Land Precipitation Variations during 1900–88. Journal of Climate 10, 2943–2962 (1997). 31. New, M., Hulme, M. & Jones, P. Representing Twentieth-Century Space–Time Climate Variability. Part I: Development of a 1961–90 Mean Monthly Terrestrial Climatology. Journal of Climate 12, 829–856 (1999). 32. Moron, V., Robertson, A. W., Ward, M. N. & Camberlin, P. Spatial Coherence of Tropical Rainfall at the Regional Scale. Journal of Climate 20, 5244–5263 (2007). 33. Braconnot, P. et al. Results of PMIP2 coupled simulations of the Mid-Holocene and Last Glacial Maximum, Part 1: experiments and large-scale features. Climate of the Past 3, 261–277 (2007). 34. Delworth, T. et al. GFDL’s CM2 global coupled climate models. Part I: Formulation and simulation characteristics. J. Clim. 19, 643–674 (2006). 35. Peltier, W. R. Global Glacial Isostasy and the Surface of the Ice-Age Earth: The ICE-5G (VM2) Model and Grace. Annu. Rev. Earth Planet. Sci. 32, 111–149 (2004). 36. Toscano, M. A., Peltier, W. R. & Drummond, R. ICE-5G and ICE-6G models of postglacial relative sea-level history applied to the Holocene coral reef record of northeastern St Croix, U.S.V.I.: investigating the influence of rotational feedback on GIA processes at tropical latitudes. Quaternary Science Reviews 30, 3032 – 3042 (2011). 37. Tarasov, L., Dyke, A. S., Neal, R. M. & Peltier, W. A data-calibrated distribution of deglacial chronologies for the North American ice complex from glaciological modeling. Earth and Planetary Science Letters 315–316, 30–40 (2012). 38. Lambeck, K. Glacial Isostasy, 374–380 (Springer, 2009). 39. Chou, C. & Neelin, J. D. Mechanisms of Global Warming Impacts on Regional Tropical Precipitation. Journal of Climate 17, 2688–2701 (2004). 40. Held, I. M. & Soden, B. J. Robust Responses of the Hydrological Cycle to Global Warming. Journal of Climate 19, 5686–5699 (2006). 41. Uppala, S. M. et al. The ERA-40 re-analysis. Q.J.R. Meteorol. Soc. 131, 2961–3012 (2005). 42. Antonov, J. I. et al. World Ocean Atlas 2009, Volume 2: Salinity, 184 (U.S. Government Printing Office, 2010). 43. Wurster, C. et al. Forest contraction in north equatorial Southeast Asia during the Last Glacial Period. Proceedings of the National Academy of Sciences 107, 15508–15511 (2010). 44. DiNezio, P. et al. The response of the Walker circulation to Last Glacial Maximum forcing: Implications for detection in proxies. Paleoceanography 26, PA3217 (2011). 16 45. Bush, A. B. G. & Fairbanks, R. G. Exposing the Sunda shelf: Tropical responses to eustatic sea level change. Journal of Geophysical Research: Atmospheres 108 (2003). 46. Marshall, A. G. & Lynch, A. H. The sensitivity of the Australian summer monsoon to climate forcing during the late Quaternary. Journal of Geophysical Research: Atmospheres 113 (2008). 47. Schrag, D., Hampt, G. & Murray, D. Pore fluid constraints on the temperature and oxygen isotopic composition of the glacial ocean. Science 272, 1930–1932 (1996). 48. Cohen, J. A. A coefficient of agreement for nominal scales. Educational and psychological measurement 20, 37–46 (1960). 49. Cohen, J. Weighted kappa: nominal scale agreement with provision for scaled disagreement or partial credit. Psychological Bulletin 70, 213–220 (1968). 50. Chen, X., Chappell, J. & Murray, A. High (ground) water levels and dune development in central Australia: TL dates from gypsum and quartz dunes around Lake Lewis (Napperby), Northern Territory. Geomorphology 11, 311–322 (1995). 51. English, P., Spooner, N., Chappell, J., Questiaux, D. & Hill, N. Lake Lewis basin, central Australia: environmental evolution and OSL chronology. Quaternary International 83, 81–101 (2001). 52. van der Kaars, S., De Deckker, P. & Gingele, F. A 100,000-year record of annual and seasonal rainfall and temperature for northwestern Australia based on a pollen record obtained offshore. Journal of Quaternary Science 21, 879–889 (2006). 53. Wyrwoll, K. & Miller, G. Initiation of the Australian summer monsoon 14,000 years ago. Quaternary International 83, 119–128 (2001). 54. Fitzsimmons, K., Miller, G., Spooner, N. & Magee, J. Aridity in the monsoon zone as indicated by desert dune formation in the Gregory Lakes basin, northwestern Australia. Australian Journal of Earth Sciences 59, 469–478 (2012). 55. Gasse, F. & Campo, E. V. A 40,000-yr pollen and diatom record from Lake Tritrivakely, Madagascar, in the southern tropics. Quat. Res. 49, 299–311 (1998). 56. Wang, Y. et al. What does leaf wax δD from a mixed C3 /C4 vegetation region tell us? Geochimica et Cosmochimica Acta (2012). 57. Wallis, L. Environmental history of northwest Australia based on phytolith analysis at Carpenter’s Gap 1. Quaternary International 83, 103–117 (2001). 58. Kershaw, A. Climatic change and Aboriginal burning in north-east Australia during the last two glacial/interglacial cycles. Nature 322, 47–49 (1986). 59. Haberle, S. A 23,000-yr pollen record from Lake Euramoo, wet tropics of NE Queensland, Australia. Quaternary Research 64, 343–356 (2005). 17 60. Moss, P. & Kershaw, P. The last glacial cycle from the humid tropics of northeastern Australia: comparison of a terrestrial and a marine record. Palaeogeography, Palaeoclimatology, Palaeoecology 155, 155–176 (2000). 61. Hesse, P. & McTainsh, G. Australian dust deposits: modern processes and the Quaternary record. Quaternary Science Reviews 22, 2007–2035 (2003). 62. Kershaw, A., van der Kaars, S. & Flenley, J. The Quaternary history of far eastern rainforests. In Bush, M. & Flenley, J. (eds.) Tropical Rainforest Responses to Climatic Change, 77–115 (Springer, 2007). 63. Williams, M. et al. Glacial and deglacial climatic patterns in Australia and surrounding regions from 35 000 to 10 000 years ago reconstructed from terrestrial and near-shore proxy data. Quaternary Science Reviews 28, 2398–2419 (2009). 64. Reeves, J. et al. The sedimentary record of palaeoenvironments and sea-level change in the Gulf of Carpentaria, Australia, through the last glacial cycle. Quaternary International 183, 3–22 (2008). 65. Wang, X., van der Kaars, S., Kershaw, P., Bird, M. & Jansen, F. A record of fire, vegetation and climate through the last three glacial cycles from Lombok Ridge core G6-4, eastern Indian Ocean, Indonesia. Palaeogeography, Palaeoclimatology, Palaeoecology 147, 241–256 (1999). 66. Vincens, A., Garcin, Y. & Buchet, G. Influence of rainfall seasonality on African lowland vegetation during the Late Quaternary: pollen evidence from Lake Masoko, Tanzania. J. Biogeogr. 34, 1274–1288 (2007). 67. Stager, J. C. Environmental changes at Lake Cheshi, Zambia since 40,000 years BP. Quat. Res. 29, 54–65 (1988). 68. Mumbi, C. T., Marchant, R., Hooghiemstra, H. & Wooller, M. J. Late Quaternary vegetation reconstruction from the Eastern Arc Mountains, Tanzania. Quat. Res. 69, 326–341 (2008). 69. van der Kaars, W. & Dam, M. A 135,000-year record of vegetational and climatic change from the Bandung area, West-Java, Indonesia. Palaeogeography, Palaeoclimatology, Palaeoecology 117, 55–72 (1995). 70. Stuijts, I., Newsome, J. & Flenley, J. Evidence for Late Quaternary vegetational change in the Sumatran and Javan highlands. Review of Palaeobotany and Palynology 55, 207–216 (1988). 71. van der Kaars, S., Bassinot, F., De Deckker, P. & Guichard, F. Changes in monsoon and ocean circulation and the vegetation cover of southwest Sumatra through the last 83,000 years: The record from marine core BAR94-42. Palaeogeography, Palaeoclimatology, Palaeoecology 296, 52–78 (2010). 18 72. Gasse, F., Ledee, V., Massault, M. & Fontes, J. C. Water-level Fluctuations of Lake Tanganyika in phase with oceanic changes during the last glaciation and deglaciation. Nature 342, 57–59 (1989). 73. Felton, A. A. et al. Paleolimnological evidence for the onset and termination of glacial aridity from Lake Tanganyika, Tropical East Africa. Palaeogeog. Palaeoclimatol. Palaeoecol. 252, 405–423 (2007). 74. Tierney, J. E. et al. Northern hemisphere controls on tropical southeast African climate during the past 60,000 years. Science 322, 252–255 (2008). 75. van der Kaars, S., Wang, X., Kershaw, P., Guichard, F. & Setiabudi, D. A late Quaternary palaeoecological record from the Banda Sea, Indonesia: patterns of vegetation, climate and biomass burning in Indonesia and northern Australia. Palaeogeography, Palaeoclimatology, Palaeoecology 155, 135–153 (2000). 76. Haberle, S. Late Quaternary vegetation change in the Tari Basin, Papua New Guinea. Palaeogeography, Palaeoclimatology, Palaeoecology 137, 1–24 (1998). 77. Street-Perrott, F. A., Marchand, D. S., Roberts, N. & Harrison, S. P. Global Lake-Level Variations from 18,000 to 0 Years Ago: A Paleoclimatic Analysis. U.S. Department of Energy Technical Report 46, Washington, DC 20545 (1989). 78. Casanova, J. & Hillaire-Marcel, C. Chronology and paleohydrology of late Quaternary high lake levels in the Manyara basin (Tanzania) from isotopic data (18 O,13 C,14 C,Th/U) on fossil stromatolites. Quaternary Research 38, 205–226 (1992). 79. COHMAP Members. Oxford Lake Levels Database, NOAA/NGDC Paleoclimatology Program, Boulder CO, USA (1994). 80. Bonnefille, R. & Chalié, F. Pollen-inferred precipitation time-series from equatorial mountains, Africa, the last 40 kyr BP. Global Planet. Change 26, 25–50 (2000). 81. Hope, G. Environmental change in the late Pleistocene and later Holocene at Wanda site, Soroako, South Sulawesi, Indonesia. Palaeogeography, Palaeoclimatology, Palaeoecology 171, 129–145 (2001). 82. Hope, G. & Tulip, J. A long vegetation history from lowland Irian Jaya, Indonesia. Palaeogeography, Palaeoclimatology, Palaeoecology 109, 385–398 (1994). 83. Newsome, J. & Flenley, J. Late Quaternary vegetational history of the Central Highlands of Sumatra. II. Palaeopalynology and vegetational history. Journal of Biogeography 555– 578 (1988). 84. Johnson, T. et al. Late Pleistocene desiccation of Lake Victoria and rapid evolution of cichlid fishes. Science 273, 1091–1093 (1996). 85. Talbot, M. & Laerdal, T. The Late Pleistocene-Holocene palaeolimnology of Lake Victoria, East Africa, based upon elemental and isotopic analyses of sedimentary organic matter. J. Paleolimnol. 23, 141–164 (2000). 19 86. Caratini, C. & Tissot, C. Paleogeographical evolution of the Mahakam delta in Kalimantan, Indonesia during the Quaternary and late Pliocene. Review of Palaeobotany and Palynology 55, 217–228 (1988). 87. Perrott, R. A. & Street-Perrott, F. A. New evidence for a late Pleistocene wet phase in northern intertropical Africa. Palaeoecology of Africa 14, 57–76 (1982). 88. Olago, D., Street-Perrott, F., Perrott, R., Ivanovich, M. & Harkness, D. Late Quaternary glacial-interglacial cycle of climatic and environmental change on Mount Kenya, Kenya. Journal of African Earth Sciences 29, 593–618 (1999). 89. Olago, D. et al. Long-term temporal characteristics of palaeomonsoon dynamics in equatorial Africa. Global and Planetary Change 26, 159–171 (2000). 90. Anshari, G., Peter Kershaw, A. & van der Kaars, S. A late Pleistocene and Holocene pollen and charcoal record from peat swamp forest, Lake Sentarum Wildlife Reserve, West Kalimantan, Indonesia. Palaeogeography, Palaeoclimatology, Palaeoecology 171, 213–228 (2001). 91. Dam, R., Fluin, J., Suparan, P. & van der Kaars, S. Palaeoenvironmental developments in the Lake Tondano area (N. Sulawesi, Indonesia) since 33,000 yr BP. Palaeogeography, Palaeoclimatology, Palaeoecology 171, 147–183 (2001). 92. Taylor, D., Yen, O., Sanderson, P. & Dodson, J. Late Quaternary peat formation and vegetation dynamics in a lowland tropical swamp; Nee Soon, Singapore. Palaeogeography, Palaeoclimatology, Palaeoecology 171, 269–287 (2001). 93. Williams, M. et al. Abrupt return of the summer monsoon 15,000 years ago: new supporting evidence from the lower White Nile valley and Lake Albert. Quaternary Science Reviews 25, 2651–2665 (2006). 94. Maloney, B. & McCormac, F. Thirty thousand years of radiocarbon dated vegetation and climatic change in highland Sumatra. Radiocarbon 37, 181–190 (1995). 95. Wang, X., Sun, X., Wang, P. & Stattegger, K. Vegetation on the Sunda shelf, South China Sea, during the Last Glacial Maximum. Palaeogeography, Palaeoclimatology, Palaeoecology 278, 88–97 (2009). 96. Partin, J. W., Cobb, K. M., Adkins, J. F., Clark, B. & Fernandez, D. P. Millennial-scale trends in west Pacific warm pool hydrology since the Last Glacial Maximum. Nature 449, 452–U3 (2007). 97. Sun, X., Li, X., Luo, Y. & Chen, X. The vegetation and climate at the last glaciation on the emerged continental shelf of the South China Sea. Palaeogeography, Palaeoclimatology, Palaeoecology 160, 301–316 (2000). 98. Bian, Y. et al. A palynological and palaeoclimatological record from the southern Philippines since the Last Glacial Maximum. Chinese Science Bulletin 56, 2359–2365 (2011). 20 99. Van Campo, E. Monsoon fluctuations in two 20,000-yr BP oxygen-isotope/pollen records off southwest India. Quaternary Research 26, 376–388 (1986). 100. Gasse, E. & Street, F. A. Late Quaternary lake-level fluctuations and environments of the northern Rift Valley and Afar region (Ethiopia and Djibouti). Palaeogeogr. Palaeoclimatol. Palaeoecol. 24, 279–325 (1978). 101. Sukumar, R., Ramesh, R., Pant, R. & Rajagopalan, G. A δ 13 C record of late Quaternary climate change from tropical peats in southern India. Nature 364, 703–706 (1993). 102. Rajagopalan, G., Sukumar, R., Ramesh, R., Pant, R. & Rajagopalan, G. Late Quaternary vegetational and climatic changes from tropical peats in southern India- An extended record up to 40,000 years BP. Current Science 73, 60–63 (1997). 103. Shakun, J. D. et al. A high-resolution, absolute-dated deglacial speleothem record of Indian Ocean climate from Socotra Island, Yemen. Earth Planet. Sci. Lett. 259, 442–456 (2007). 104. Penny, D. A 40,000 year palynological record from north-east Thailand; implications for biogeography and palaeo-environmental reconstruction. Palaeogeography, Palaeoclimatology, Palaeoecology 171, 97–128 (2001). 105. Rogers, J. & De Deckker, P. Radiolaria as a reflection of environmental conditions in the eastern and southern sectors of the Indian Ocean: A new statistical approach. Marine Micropaleontology 65, 137–162 (2007). 106. Martinez, J. I., De Deckker, P. & Barrows, T. T. Paleoceanography of the western Pacific warm pool during the Last Glacial Maximum: Long-term climate monitoring of the Maritime Continent. In Kershaw, P., David, B., Tapper, N., Penny, D. & Brown, J. (eds.) Bridging Wallace’s Line: The Environmental and Cultural History and Dynamics of the SE-Asian-Australian Region, vol. 34 of Advances in Geoecology, 147–172 (Catena Verlag, 2002). 107. Mohtadi, M., Steinke, S., Lückge, A., Groeneveld, J. & Hathorne, E. Glacial to Holocene surface hydrography of the tropical eastern Indian Ocean. Earth and Planetary Science Letters 292, 89–97 (2010). 108. Visser, K., Thunell, R. & Stott, L. Magnitude and timing of temperature change in the Indo-Pacific warm pool during deglaciation. Nature 421, 152–155 (2003). 109. Cullen, J. Microfossil evidence for changing salinity patterns in the Bay of Bengal over the last 20,000 years. Palaeogeography, Palaeoclimatology, Palaeoecology 35, 315–356 (1981). 110. de Garidel-Thoron, T. et al. A multiproxy assessment of the western equatorial Pacific hydrography during the last 30 kyr. Paleoceanography 22, PA3204 (2007). 21 111. de Garidel-Thoron, T., Rosenthal, Y., Bassinot, F. & Beaufort, L. Stable sea surface temperatures in the western Pacific warm pool over the past 1.75 million years. Nature 433, 294–298 (2005). 112. Rostek, F. et al. Reconstructing sea surface temperature and salinity using δ 18 O and alkenone records. Nature 364, 319–321 (1993). 113. Ahmad, S., Babu, G., Padmakumari, V. & Raza, W. Surface and deep water changes in the northeast Indian Ocean during the last 60 ka inferred from carbon and oxygen isotopes of planktonic and benthic foraminifera. Palaeogeography, Palaeoclimatology, Palaeoecology 262, 182–188 (2008). 114. Stott, L. D., Poulsen, C., Lund, S. & Thunell, R. C. Super ENSO and global climate oscillations at millennial time scales. Science 297, 222–226 (2002). 115. Steinke, S. et al. On the influence of sea level and monsoon climate on the southern South China Sea freshwater budget over the last 22,000 years. Quaternary Science Reviews 25, 1475–1488 (2006). 116. Anand, P. et al. Coupled sea surface temperature–seawater δ 18 O reconstructions in the Arabian Sea at the millennial scale for the last 35 ka. Paleoceanography 23, PA4207 (2008). 117. Rashid, H., Flower, B., Poore, R. & Quinn, T. A ∼25 ka Indian Ocean monsoon variability record from the Andaman Sea. Quaternary Science Reviews 26, 2586–2597 (2007). 118. Govil, P. & Divakar Naidu, P. Variations of Indian monsoon precipitation during the last 32 kyr reflected in the surface hydrography of the Western Bay of Bengal. Quaternary Science Reviews 30, 3871–3879 (2011). 119. Rashid, H., England, E., Thompson, L. & Polyak, L. Late Glacial to Holocene Indian Summer Monsoon variability based upon sediment records taken from the Bay of Bengal. Terr. Atmos. Ocean. Sci. 22, 215–228 (2011). 120. Banakar, V., Mahesh, B., Burr, G. & Chodankar, A. Climatology of the Eastern Arabian Sea during the last glacial cycle reconstructed from paired measurement of foraminiferal δ 18 O and Mg/Ca. Quaternary Research 73, 535–540 (2010). 121. Spooner, M., Barrows, T., De Deckker, P. & Paterne, M. Palaeoceanography of the Banda Sea, and late Pleistocene initiation of the northwest monsoon. Global and Planetary Change 49, 28–46 (2005). 122. Spooner, M., De Deckker, P., Barrows, T. & Fifield, L. The behaviour of the Leeuwin Current offshore NW Australia during the last five glacial–interglacial cycles. Global and Planetary Change 75, 119–132 (2011). 123. Levi, C. et al. Low-latitude hydrological cycle and rapid climate changes during the last deglaciation. Geochemistry Geophysics Geosystems 8, Q05N12 (2007). 22 124. Yongqiang, Y., Xuehong, Z. & Yufu, G. Global coupled ocean-atmosphere general circulation models in LASG/IAP. Advances in Atmospheric Sciences 21, 444–455 (2004). 125. Marti, O. et al. Key features of the IPSL ocean atmosphere model and its sensitivity to atmospheric resolution. Climate Dynamics 34, 1–26 (2010). 126. Hasumi, H. & Emori, S. K-1 coupled GCM(MIROC) description. K-1 Tech. Rep. 1, 34 (2004). 127. Otto-Bliesner, B. L. et al. Last Glacial Maximum and Holocene Climate in CCSM3. Journal of Climate 19, 2526–2544 (2006). 128. Otto-Bliesner, B. L. et al. Climate Sensitivity of Moderate- and Low-Resolution Versions of CCSM3 to Preindustrial Forcings. Journal of Climate 19, 2567–2583 (2006). 129. Gordon, C. et al. The simulation of SST, sea ice extents and ocean heat transports in a version of the Hadley Centre coupled model without flux adjustments. Climate Dynamics 16, 147–168 (2000). 130. Gent, P. R. et al. The community climate system model version 4. Journal of Climate 24, 4973–4991 (2011). 131. Brady, E. C., Otto-Bliesner, B. L., Kay, J. E. & Rosenbloom, N. Sensitivity to Glacial Forcing in the CCSM4. Journal of Climate (2012). 132. Yukimoto, S. et al. A new global climate model of the meteorological research institute: Mri-cgcm3 mdash;model description and basic performancemdash;. Journal of the Meteorological Society of Japan. Ser. II 90A, 23–64 (2012). 133. Voldoire, A. et al. The CNRM-CM5.1 global climate model: description and basic evaluation. Climate Dynamics 1–31 (2012). 23 Table S1. List of the terrestrial paleoclimate proxy data used in the LGM synthesis, with locations, hydroclimatic change category (W=wetter than present, NC=no change, D=drier than present, NR=not robust) and associated references. Longitude 132.30 113.50 127.45 Latitude -22.80 -22.05 -19.80 Country/Region Central Australia Western Australia Northwestern Australia 46.92 37.87 125.00 145.70 146.63 146.30 118.38 35.50 121.59 131.00 -19.78 -18.24 -17.45 -17.37 -16.83 -16.63 -16.35 -15.50 -13.17 -13.00 Madagascar Zambezi River Delta Northwestern Australia Northeastern Australia Northeastern Australia Northeastern Australia Eastern Indian Ocean Malawi North Australian Basin Northwestern Australia 34.00 -12.00 Malawi Core/Site ID Lake Lewis Fr10/95-GC17 Lake Gregory, Fitzroy Basin Lake Tritrivakely 64PE203-80 Carpenter’s Gap Lynch’s Crater Lake Euramoo ODP Site 820 SO-14-08-05 Lake Chilwa MD98-2167 Lily pond, Wangi falls, Waterfall creek Lake Malawi 140.00 118.07 33.76 -12.00 -10.78 -9.33 Northern Australia Lombok Ridge Malawi/Rungwe Mtns Gulf of Carpentaria G6-4 Lake Masoko 29.75 36.00 -9.08 -8.00 Zambia Tanzania 107.80 107.50 102.42 30.00 -7.20 -7.10 -6.08 -6.00 West Java Java Sumatra East Africa Lake Cheshi Eastern Arc Mountains Bandung Basin Situ Bayongbong BAR94-42 Lake Tanganyika 126.97 143.00 133.45 35.82 29.50 37.70 -5.77 -5.50 -5.00 -3.92 -3.50 -3.32 Banda Sea Papua New Guinea Aru Sea Tanzania Eastern Central Africa Kenya/Tanzania SHI-9014 Tari basin MD98-2175 Lake Manyara Multiple sites Lake Challa 121.30 -2.50 Sulawesi, Indonesia Wanda Site 140.50 100.77 33.00 117.50 36.33 36.17 -2.33 -1.07 -1.00 -0.75 -0.68 -0.42 Irian Jaya Sumatra Tanzania/Uganda/Kenya Kalimantan Kenya Kenya 37.00 37.53 112.10 124.89 0.00 0.05 0.85 1.23 Kenya Kenya Kalimantan Sulawesi, Indonesia Lake Hordorli Danau di Atas Lake Victoria Mahakam Delta Lake Naivasha Lake NakuruElementaita Aberdare Mountains Sacred Lake Lake Sentarum Lake Tondano 103.80 31.00 1.40 1.67 Singapore Albert Nee Soon Lake Albert 24 Proxy high dune activity pollen lake desiccation and dune activity diatoms CBT index, δDwax phytoliths pollen, charcoal pollen pollen, charcoal dust flux shorelines/lake level pollen evidence for paleofloods δ 13 Cwax , vivianite, δDwax , diatoms, OM characterization lake level pollen, charcoal runoff (magnetic susceptibility), pollen lake level pollen Category D D D Reference 50, 51 52 53, 54 D W D D D D D NR D W 55 10, 56 57 58 59 60 61 17 62, 63 24 NR 8, 14, 15, 13, 16 D D W 64 65 12, 66 NC W 67 68 pollen pollen pollen diatoms, trace metals, δDwax pollen, charcoal pollen, charcoal pollen lake level pollen BIT index, δDwax , lake level pollen, site nearly desiccated pollen, charcoal pollen lake desiccation pollen lake level lake level D NC NC D 69 70 71 72, 73, 74 D NC D W D W 75 76 62, 63 7, 77, 78, 79 80 18, 9 D 81 NC NC D D W D 82 70, 83 84, 85 86 7, 77, 79 7, 77, 79 W D D D 87 88, 89 90 91 D D 92 93 pollen δ 13 COC site nearly desiccated pollen, diatoms, lake desiccation pollen lake level Longitude 99.03 107.88 101.70 113.77 114.00 108.57 108.65 36.00 112.21 125.83 117.55 119.45 75.23 41.60 76.75 44.30 54.00 72.50 102.90 Latitude 2.25 2.78 3.22 3.82 4.00 4.15 4.35 5.00 6.16 6.48 8.52 10.47 10.50 11.10 11.25 11.96 12.50 15.50 17.10 Country/Region Sumatra South China Sea Peninsular Malaysia Northern Borneo Borneo (Malaysia) South China Sea South China Sea Kenya South China Sea Davao Gulf, Phillipines Palawan, Phillipines Palawan, Phillipines Pakistan Ethiopia/Djibouti Southern India Gulf of Aden Socotra Pakistan Northeast Thailand Core/Site ID Pea Bullok SO18323 Batu Cave Niah Cave Gunang Buda SO18302 SO18300 Lake Turkana SO17964 MD06-3075 Gangub Cave Makangit Cave MD77-194 Lake Abhe Nilgiri Hills P178-15P Moomi Cave MD76-131 Nong Pa Kho 25 Proxy pollen pollen bat guano δ 13 C bat guano δ 13 C speleothem δ 18 O pollen pollen lake level pollen pollen bat guano δ 13 C bat guano δ 13 C pollen lake level peat δ 13 C δDwax speleothem δ 18 O pollen charcoal Category NC D D NC D NC D D NC D D D D W D D D D D Reference 94 95 43 43 96 95 95 7, 77, 79 97 98 43 43 99 100 101, 102 Tierney et al, in prep 103 99 104 Table S2. List of the marine paleoclimate proxy data used in the LGM synthesis, with locations, salinity change category (F=fresher than present, NC=no change, S=saltier than present, NR=not robust) and associated references. Longitude 111.83 108.51 108.37 166.15 113.50 110.51 115.00 166.28 119.50 120.70 121.03 120.00 121.79 122.42 121.70 120.99 120.67 51.77 129.24 128.64 128.17 103.25 117.90 100.13 90.00 159.36 129.79 146.14 44.78 67.34 90.00 141.77 127.74 73.88 90.08 60.25 90.52 125.83 111.82 113.42 126.50 139.64 121.30 90.03 57.74 51.93 115.31 96.04 90.02 82.00 57.35 Latitude -24.74 -24.46 -23.95 -23.00 -22.13 -19.41 -17.64 -15.79 -15.31 -14.98 -14.01 -13.85 -13.08 -11.52 -11.30 -10.95 -10.83 -10.15 -9.09 -8.79 -8.50 -5.94 -4.69 -1.49 -1.00 0.32 1.13 1.25 1.25 1.90 2.00 2.03 2.68 5.07 5.18 5.39 5.85 6.30 6.37 6.47 6.51 8.02 8.80 9.13 10.50 10.77 10.90 12.46 12.70 14.04 14.32 Country/Region Western Australia Western Australia Western Australia Coral Sea Western Australia Western Australia Western Australia Coral Sea Timor Sea Timor Sea Timor Sea Timor Sea Timor Sea Timor Sea Timor Sea Timor Sea Timor Sea Western Indian Ocean Timor Sea Timor Sea Timor Sea West Sumatra Makassar Strait West Sumatra Bay of Bengal Ontong Java Plateau West Pacific West Pacific Warm Pool Arabian Sea Arabian Sea Bay of Bengal West Pacific Warm Pool Molucca Sea Arabian Sea Bay of Bengal Arabian Sea Bay of Bengal Mindanao South China Sea South China Sea West Pacific Caroline Basin Sulu Sea Bay of Bengal Arabian Sea Arabian Sea South China Sea Andaman Sea Bay of Bengal Bay of Bengal Arabian Sea Core/Site ID Fr10/95-GC20 Fr10/96-GC10 Fr10/95-GC11 MD06-301 Fr10/95-GC17 Fr10/95-GC14 Fr10/95-11 ODP 828A SO18506 SO18500 Fr10/95-5 SO18507 MD01-2378 SO18473 SO18475 SO18476 SO18477 WIND 28K SO18462 SO18460 SO18459 GeoB 10038-4 MD98-2162 GeoB 10029-4 RC14-35 ODP806B MD01-2386 MD97-2138 IOE-143KK SO28-18KL RC14-37 MD97-2140 K12 MD90-0963 SK157-14 SO28-11KL RC14-39 MD98-2181 SO18267 MD01-2390 MD06-3067 3cBX MD97-2141 RC12-339 SO42-87KL NIOP 905 SO17957 RC12-344 RC12-340 SK218/1 SO42-74KL 26 Proxy radiolaria transfer function radiolaria transfer function radiolaria transfer function δ 18 Osw radiolaria transfer function radiolaria transfer function δ 18 O + foram transfer SST δ 18 O + foram transfer SST δ 18 Osw δ 18 Osw δ 18 O + foram transfer SST δ 18 Osw δ 18 Osw δ 18 Osw δ 18 Osw δ 18 Osw δ 18 Osw δ 18 Oruber δ 18 Osw δ 18 Osw δ 18 Osw δ 18 Osw δ 18 Osw δ 18 Osw foram transfer function δ 18 Osw δ 18 Osw δ 18 Osw δ 18 Osw δ 18 Osw foram transfer function δ 18 Osw δ 18 O + foram transfer SST δ 18 Oruber +alkenone SST δ 18 Oruber δ 18 Osw foram transfer function δ 18 Osw δ 18 Osw δ 18 Osw δ 18 Osw δ 18 Osw δ 18 Osw foram transfer function δ 18 Osw δ 18 Osw δ 18 Osw δ 18 Osw foram transfer function δ 18 Osw δ 18 Osw Category NC F F F F F NC S NC S S F F S S S S F NC NC S S NC NC NC F NC F S F NC F NC S S F NC S F F NC F F F F NC S NC NC S F Reference 105 105 105 28 105 105 106 106 27 27 106 27 27 27 27 27 27 21 27 27 27 107 108, 27 107 109 25 27 110 19 19 109 111 106 112 113 19 109 114 27 115 27 29 26 109 19 116 27 117 109 118 19 Longitude 83.58 111.53 90.00 72.97 72.85 68.76 112.90 69.05 119.45 58.80 90.00 126.24 113.48 118.34 36.33 Latitude 14.71 14.76 15.00 15.25 15.48 15.52 16.09 17.08 17.25 17.38 17.50 -8.46 -22.08 -9.65 -20.40 Country/Region Bay of Bengal South China Sea Bay of Bengal Arabian Sea Arabian Sea Arabian Sea South China Sea Arabian Sea South China Sea Arabian Sea Bay of Bengal Banda Sea Western Australia Indonesia Mozambique Channel Core/Site ID VM29-19 SO17954 RC12-343 ORV SK17 SK117-GC8 SO42-26KL SO17950 SO42-36KL SO17927 TN47-6GGC MD77-181 SHI-9016 MD00-2361 MD98-2165 MD79-257 27 Proxy δ 18 Osw δ 18 Osw foram transfer function δ 18 Osw δ 18 Osw δ 18 Osw δ 18 Osw δ 18 Osw δ 18 Osw δ 18 Osw foram transfer function δ 18 O + foram transfer SST δ 18 O + foram transfer SST δ 18 Osw δ 18 O + foram transfer SST Category NC F S S S F F F F NC S S F NC NC Reference 119 27 109 116 120 19 27 19 27 19 109 121 122 123 123 Table S3. List of models used in the LGM and PI simulations, their atmospheric and oceanic resolution, and associated references. References for GISS-E2-R, IPSL-CM5A-LR, MPI-ESM are not yet available. Model Institution, Country Model Resolution Atmosphere (lat. × Ocean (lat. × long.) long.) Reference National Oceanic and Atmospheric Administration Geophysical Fluid Dynamics Laboratory, USA LASG/Institute of Atmospheric Physics, China Institut Pierre Simon Laplace, France Center for Climate System Research (University of Tokyo), National Institute for Environmental Studies, and Frontier Research Center for Global Change (JAMSTEC), Japan National Center for Atmospheric Research, USA UK Met Office Hadley Centre, UKa 2◦ ×2.5◦ L24 1/3◦ –1◦ ×1◦ L50 34 T42 L26 (2.8◦ ×2.8◦ ) 1◦ ×1◦ L33 124 2.5◦ ×3.75◦ L19 T42 L20 (2.8◦ ×2.8◦ ) 1–2◦ ×2◦ L31 0.5–1.4◦ ×1.4◦ L43 125 126 T42 L26 (2.8◦ ×2.8◦ ) 1/3–1◦ ×1◦ L40 127, 128 2.5◦ ×3.75◦ L19 1.25◦ ×1.25◦ L20 129 2◦ ×25◦ L40 1◦ ×1.25◦ L32 2.5◦ ×3.75◦ L19 T63 L47 (1.4◦ ×0.9◦ ) 1.25◦ ×0.9◦ L26 1–2◦ ×2◦ L31 1.4◦ ×1.4◦ L44 1/3–1◦ ×1◦ L60 130, 131 T159 L48 (1.125◦ ×1.125◦ ) T127 L31 (1.4◦ ×1.4◦ ) L26 0.5◦ ×1◦ L50 132 1/3–1◦ ×1◦ L42 133 PMIP2 GFDL-CM2.1 FGOALS-g1.0 IPSL-CM4 MIROC3.2 CCSM3.0 HadCM3 PMIP3 GISS-E2-R IPSL-CM5A-LR MPI-ESM CCSM4.0 MRI-CGCM3 CNRM-CM5 NASA Goddard Institute for Space Studies, USA Institut Pierre Simon Laplace, France Max Planck Institute, Germany National Center for Atmospheric Research, USA Meteorological Research Institute, Japan Centre National de Mètèorologiques, France Recherches 28 20 (b) 30˚N 0.0 Simulated rainfall change 0.1 0.2 0.3 Cohen’s κ 0.4 CNRM−CM5 CCSM4.0 MRI−CGCM3 MPI−ESM−P IPSL−CM5A−LR HadCM3 GISS−E2−R 120˚E CCSM3.0 60˚E MIROC3.2 30˚S IPSL−CM4 10 GFDL−CM2.1 0˚ 30 FGOALS−1.0g 30˚N Proxy−model agreement Wet−get−drier Drier No Change Wetter Not robust (c) ∆P (%) Proxy network drier−wetter threshold (a) 0.5 Wet−get−drier GFDL−CM2.1 FGOALS−1.0g κmax=0.11 (∆P=1 %) κmax=0.19 (∆P=1 %) κmax=0.13 (∆P=5 %) 0˚ 30˚S 30˚N IPSL−CM4 MIROC3.2 CCSM3.0 κmax=0.05 (∆P=13 %) κmax=0.20 (∆P=14 %) κmax=0.22 (∆P=1 %) 0˚ 30˚S 30˚N HadCM3 GISS−E2−R IPSL−CM5A−LR κmax=0.45* (∆P=13 %) κmax=0.10 (∆P=21 %) κmax=0.01 (∆P=47 %) 0˚ 30˚S 30˚N MPI−ESM−P CCSM4.0 MRI−CGCM3 κmax=0.20 (∆P=1 %) κmax=0.02 (∆P=31 %) κmax=0.07 (∆P=22 %) 60˚E 60˚E 0˚ 30˚S CNRM−CM5 30˚N 120˚E 120˚E κmax=0.22 (∆P=21 %) −50 −40 −30 −20 −10 0 0˚ 10 20 30 40 50 ∆P (%) Model−proxy agree Model−proxy do not agree 30˚S 60˚E 120˚E Figure S1. Reconstructed and simulated changes in rainfall during the Last Glacial Maximum (LGM) and Cohen’s κ results, as in Figure 2 of the main text but including only proxy data with one or more radiocarbon, U/Th, or well-constrained OSL dates near to or during the LGM. HadCM3 is still the only model with a significant match. 29 (c) Proxy−model agreement 1.00 ∆SSS (psu) 0.50 (b) Simulated sea−surface salinity change 30˚N 0.0 0.1 0.3 Cohen’s κ 0.4 CNRM−CM5 CCSM4.0 MRI−CGCM3 MPI−ESM−P IPSL−CM5A−LR HadCM3 0.2 GISS−E2−R CCSM3.0 120˚E 60˚E MIROC3.2 30˚S IPSL−CM4 0.25 Wet−get−drier 0˚ 0.75 GFDL−CM2.1 30˚N saltier−fresher threshold Saltier No Change Fresher Not robust FGOALS−1.0g Proxy network (a) 0.5 Wet−get−drier GFDL−CM2.1 FGOALS−1.0g κmax=0.12 (∆SSS=0.10 psu) κmax=0.26 (∆SSS=0.10 psu) κmax=0.10 (∆SSS=1.40 psu) 0˚ 30˚S 30˚N IPSL−CM4 MIROC3.2 CCSM3.0 κmax=0.08 (∆SSS=0.90 psu) κmax=0.26 (∆SSS=0.50 psu) κmax=0.09 (∆SSS=0.40 psu) 0˚ 30˚S 30˚N HadCM3 GISS−E2−R IPSL−CM5A−LR κmax=0.28 (∆SSS=0.90 psu) κmax=0.20 (∆SSS=0.60 psu) κmax=0.21 (∆SSS=0.20 psu) 0˚ 30˚S 30˚N MPI−ESM−P CCSM4.0 MRI−CGCM3 κmax=0.08 (∆SSS=0.80 psu) κmax=0.18 (∆SSS=0.10 psu) κmax=0.08 (∆SSS=0.70 psu) 0˚ 30˚S CNRM−CM5 30˚N 60˚E 120˚E 60˚E 120˚E κmax=0.11 (∆SSS=0.10 psu) −1.0 0˚ −0.5 0.0 0.5 1.0 ∆SSS (psu) Model−proxy agree Model−proxy do not agree 30˚S 60˚E 120˚E Figure S2. Reconstructed and simulated changes in salinity during the Last Glacial Maximum (LGM) and Cohen’s κ results, as in Figure 3 of the main text but including only proxy data with one or more radiocarbon, U/Th, or well-constrained OSL dates near to or during the LGM. The kappa data are similar to the results using the full proxy matrix (Fig. 3 of the main text), but the values are no longer significant at the 95% level due to the depopulation of the proxy matrix to only 25 proxies. 30 30˚N 2416 IPSL−CM4 (−2.67 oC) oC) 16 FGOALS−1.0g (−2.29 2 4 24 24 24 124 GFDL−CM2.1 (−3.51 oC) 6 24 24 24 24 24 24 0˚ 24 24 24 24 24 24 24 24 24 30˚S 16 1 o MIROC3.2 (−2.07 24 C) 2 24 4 24 oC) 24 HadCM3 (−2.55 24 24 24 24 0˚ 6 CCSM3.0 (−2.04 oC) 24 24 24 24 24 16 30˚N 24 24 24 24 24 24 24 30˚S 30˚N 24 1264 GISS−E2−R (−2.54 oC) 4 24 24 126 IPSL−CM5A−LR (−4.24 oC) 2 4 24 24 24 24 24 24 24 24 24 24 24 24 24 o 24 CNRM−CM5 (−1.55 16 24 C) oC) 216 MRI−CGCM3 (−2.75 4 24 24 CCSM4 16 (−2.50 oC) 24 24 24 24 24 24 30˚S 24 24 24 24 30˚N 2416 MPI−ESM−P (−2.35 oC) 24 24 24 0˚ 24 24 24 24 24 24 24 24 60˚E 24 24 24 24 120˚E 60˚E −6 −5 24 24 24 −7 24 24 24 24 24 24 24 30˚S 24 24 0˚ 24 24 24 24 120˚E −4 ∆Ts −3 −2 −1 0 (oC) Figure S3. Change in surface temperature (Ts ) simulated by PMIP models in response to LGM forcing. Values over the ocean correspond to sea surface temperature (SST). Contours show the annual-mean Ts simulated in the PI simulation. The contour interval is 2◦ C. The coastlines correspond to the 120 m isobath of the present day ocean bathymetry. 31 20 -40 30oE 60oE 90oE 150oE 120oE 180oE 150oW 120oW Maritime Continent 90oW S. America Sunda Shelf New Guinea -20 -40 -20 -20 -60 400 -40 -4 0 -40 -20 -20 pressure (hPa) 200 800 1000 20 -20 0 600 20 -60 -40 Africa (b) 20 -40 1000 0 0 -6 800 -4 -40 -60 600 -20 400 0 -2 -20 200 20 0 20 pressure (hPa) (a) -20 30oE 60oE -20 90oE Africa 120oE 180oE 150oE 150oW 120oW S. America Maritime Continent Sunda Shelf New Guinea -25 -15 -10 -5 -2 0 90oW 2 5 10 15 25 Δω (hPa day-1) Figure S4. Changes in vertical velocity (ω) over the equatorial Indo-Pacific simulated by (a) GFDL-CM2.1 and (b) MIROC3.2 in response to LGM forcing (colors). The ω changes are averaged over the 5◦ S–5◦ N latitude band. Contours are mean-annual ω simulated in the preindustrial control experiment. The contour intervals represent 10 hPa day−1 . Note that the colorscale is not linear. 32 20 −2 40 −20 −20 0 −20 −40 40 20 −40 0 −4 ∆ω500 LGM 20 20 0 20 −20 0 −2 20 0 20 30˚S 20 −2 −20 0˚ 40 −2 20 20 −2 −4−020 0 20 30˚N 20 40 −40 20 −2−40 0 −20 20 20 20 −20 −40 −40 −2 0 −40 −20 40 −20 −20 20 20 40 20 20 60˚E 90˚E 120˚E 150˚E 180˚ 150˚W 120˚W 90˚W ∆ω500 (hPa day−1) −20 −15 −10 −8 −4 −2 0 2 4 8 10 15 20 Figure S5. Changes in vertical velocity at the 500 hPa level (ω 500 ) simulated by HadCM3 in response to LGM forcing. Contours are mean-annual ω simulated in the pre-industrial control experiment. The contour intervals represent 10 hPa day−1 . Note that the colorscale is not linear. 33 8 4 8 4 8 8 8 8 4 8 8 8 4 8 8 8 8 4 8 4 8 8 4 8 8 4 8 8 30˚S 4 4 8 8 4 4 8 κmax=0.37* (∆P=9 %) (b) 4 4 0˚ 4 κmax=0.41* (∆P=2 %) (a) 30˚N 4 4 8 8 8 4 4 8 4 4 4 8.0 4.0 2.0 1.0 0.5 0.0 −0.5 −1.0 −2.0 −4.0 −8.0 ∆P (mm day−1) HadCM3 simulated rainfall and salinity change DJF and JJA Model−proxy agree Model−proxy do not agree (c) κ * max=0.29 κmax=0.25 (∆SSS=0.10 psu) (d) (∆SSS=0.40 psu) 34 35 34 35 36 35 34 33 34 32 32 33 35 30˚S 36 34 0˚ 32 35 33 34 35 35 60˚E 2 34 33 3 34 35 120˚E 60˚E 34 0.5 32 34 33 32 34 34 1.0 33 33 33 34 120˚E 34 0.0 34 −0.5 ∆SSS (psu) 30˚N −1.0 Model−proxy agree Model−proxy do not agree Figure S6. Change in precipitation and sea surface salinity during (a,c) DJF and (b,d) JJA simulated by HadCM3 in response to LGM forcing. Contours show the annual-mean precipitation and SSS simulated in the PI simulation. The contour intervals are 2 mm/day and 1 psu respectively. Circles indicate the proxy estimates of precipitation or SSS change (fill color) and proxy-model agreement (border color). Black and red (yellow) borders indicates agreement and disagreement respectively. The coastlines correspond to the 120 m isobath of the present day ocean bathymetry. The maximum Cohens κ and the corresponding optimal threshold for defining drier/wetter or saltier/fresher conditions is shown for each panel. Asterisks indicate statistically significant (p < 0.05) Cohens κ values. 34 30˚N κmax=0.06 (∆P=39 %) 4 (a) 4 κmax=0.35* (∆P=10 %) 8 (b) 4 4 8 4 8 8 8 4 4 4 8 8 8 8 0˚ 4 4 4 4 4 8 4 8 8 8 8 8 4 8 4 8 8 8 4 4 4 30˚S 8.0 4.0 2.0 1.0 0.5 0.0 −0.5 −1.0 −2.0 −4.0 −8.0 ∆P (mm day−1) IPSL−CM4 simulated rainfall and salinity change DJF and JJA Model−proxy agree Model−proxy do not agree (c) κmax=0.07 (∆SSS=0.60 psu) (d) 35 κmax=0.10 (∆SSS=0.60 psu) 35 35 0˚ 36 36 3435 35 34 35 0.5 0.0 −0.5 35 35 36 36 1.0 ∆SSS (psu) 30˚N −1.0 30˚S 36 36 60˚E 120˚E 60˚E 120˚E Figure S7. As in Figure S6, but for IPSL-CM4. 35 Model−proxy agree Model−proxy do not agree (a) JJA minus DJF rainfall 2 30˚N 4 6 4 6 4 4 2 4 6 6 4 2 4 6 6 4 2 2 8 8 8 8 2 2 2 30˚S 4 2 6 6 4 86 8 4 2 0˚ 6 60˚E 120˚E −8.0 −4.0 −2.0 −1.0 −0.5 0.0 0.5 1.0 2.0 4.0 8.0 ∆P (mm/day) (b) SON minus MAM sea−surface salinity 30˚N 35 34.5 0˚ 35 34.5 34 35 35.5 34.5 35 35. 5 30˚S 35.5 120˚E 60˚E −1.0 −0.5 0.0 0.5 1.0 ∆SSS (psu) Figure S8. Observed seasonal contrast in climatological (a) rainfall and (b) sea-surface salinity (SSS). The rainfall and SSS contrasts are the difference between JJA minus DJF and SON minus MAM climatologies respectively. The contour intervals are 2 mm/day and 0.5 psu respectively. 36
© Copyright 2025 Paperzz