Skip to main content Accessibility help
×
Hostname: page-component-68c7f8b79f-fc4h8 Total loading time: 0 Render date: 2026-01-02T05:59:27.414Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  09 September 2025

Ian D. Goodwin
Affiliation:
Macquarie University and ClimaLab
Get access

Information

Type
Chapter
Information
Synoptic Paleoclimatology
The Weather Regime Approach from the Tropics to the Poles
, pp. 595 - 744
Publisher: Cambridge University Press
Print publication year: 2025

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

Book purchase

Temporarily unavailable

References

Primary Sources

Hansen, J. E., et al. (2025). Global warming has accelerated: Are the United Nations and the public well-informed? Environ.: Sci. Policy Sustain. Dev., 67 (1), 6–44. https://doi.org/10.1080/00139157.2025.2434494.Google Scholar
Tierney, J. E., et al. (2020). Past climates inform our future. Science, 370, eaay3701. https://doi.org/10.1126/science.aay3701.CrossRefGoogle ScholarPubMed

Secondary Sources

Alexander, M. A. (2010). Extratropical air-sea interaction, sea surface temperature variability, and the Pacific Decadal Oscillation. In Sun, D.-Z. and Bryan, F. (eds.) Climate Dynamics: Why Does the Climate Vary. Geophys. Monograph Ser. 189, pp. 123–148. American Geophysical Union, Washington, USA. https://doi.org/10.1029/2008GM000794.Google Scholar
Alexander, M. A., et al. (2002). The atmospheric bridge: The influence of ENSO teleconnections on air–sea interaction over the global oceans. J. Clim., 15, 2205–2231.2.0.CO;2>CrossRefGoogle Scholar
Allan, R., et al. (1996). El Niño Southern Oscillation and Climatic Variability. CSIRO, Melbourne, Australia, 405pp.Google Scholar
Ashok, K., et al. (2007). El Niño Modoki and its possible teleconnection. J. Geophys. Res., 112(C11), C11007. https://doi.org/10.1029/2006JC003798.Google Scholar
Ashok, K. and Yamagata, T. (2009). Climate change: The El Niño with a difference. Nature, 461, 481–484.CrossRefGoogle ScholarPubMed
Baldwin, M. P., et al. (2001). The Quasi-Biennial Oscillation. Rev. Geophys., 39(2), 179–229. https://doi.org/10.1029/1999RG000073.CrossRefGoogle Scholar
Bard, E. and Rickaby, R. E. M. (2009). Migration of the subtropical front as a modulator of glacial climate. Nature, 460, 380–383.CrossRefGoogle ScholarPubMed
Barry, R. G. and Carleton, A. M. (2001). Synoptic and Dynamic Climatology. Routledge, London, 620pp.Google Scholar
Behera, S. K. and Yamagata, T. (2001). Subtropical SST dipole events in the southern Indian Ocean. Geophys. Res. Lett., 28, 327–330.CrossRefGoogle Scholar
Bender, M. L. (2013). Paleoclimate. Princeton University Press, Princeton, NJ, 320pp.CrossRefGoogle Scholar
Berger, A. L. (1978). Long-term variations of daily insolation and Quaternary climatic changes. J. Atmos. Sci., 35, 2362–2367.2.0.CO;2>CrossRefGoogle Scholar
Berger, A. L. and Loutre, M. F. (1991). Insolation values for the climate of the last 10 million years. Quat. Sci. Rev., 10, 297–317. https://doi.org/10.1016/0277–3791(91)90033-Q.CrossRefGoogle Scholar
Bergeron, T. (1959). Methods in scientific weather analysis and forecasting: An outline in the history of ideas and hints at a program. In The Atmosphere and Sea in Motion. Rockefeller Institute Press, New York, pp. 440–474.Google Scholar
Berlage, H. P. Jr. (1966). The Southern Oscillation and world weather. Mededelingen en Verhandelingen, No. 88, KNMI, 152pp.Google Scholar
Biastoch, A., et al. (2009). Increase in Agulhas leakage due to poleward shift of Southern Hemisphere westerlies. Nature, 462, 495–498.CrossRefGoogle ScholarPubMed
Biastoch, A., et al. (2015). Atlantic multi-decadal oscillation covaries with Agulhas leakage. Nat. Comms., 6, 10082. https://doi.org/10.1038/ncomms10082.CrossRefGoogle ScholarPubMed
Bigg, G. R. (2003). The Oceans and Climate, 2nd ed. Cambridge University Press. https://doi.org/10.1017/CBO9781139165013.Google Scholar
Bjerknes, J. (1969). Atmospheric teleconnections from the equatorial Pacific. Mon. Weather Rev., 97, 163–172.2.3.CO;2>CrossRefGoogle Scholar
Blunier, T., et al. (1998). Asynchrony of Antarctic and Greenland climate change during the last glacial period. Nature, 394, 739–743.CrossRefGoogle Scholar
Bradley, R. S. (1999). Paleoclimatology: Reconstructing Climates of the Quaternary, 2nd ed. International Geophysics Press, Volume 64, Academic Press, San Diego, 610pp.Google Scholar
Broccoli, A., et al. (2006). Response of the ITCZ to Northern Hemisphere cooling. Geophys. Res. Lett., 33, L01702. https://doi.org/10.1029/2005GL024546.CrossRefGoogle Scholar
Broecker, W. S. (1998). Paleocean circulation during the last deglaciation: A bipolar seesaw? Paleoceanography, 113, 119–121. https://doi.org/10.1029/97PA03707.Google Scholar
Buchan, A. (1868). The mean pressure of the atmosphere over the globe for the months and for the year, Part I. – January, July, and the year. Proc. Roy. Soc. Edinburgh, 6, 303–307.Google Scholar
Buchan, A. (1869). The mean pressure of the atmosphere, and the prevailing winds for the months and for the year. Part II. Proc. Roy. Soc. Edinburgh, 25, 523–524.Google Scholar
Buchan, A. (1890). Report on atmospheric circulation, based on the observations made on Board H.M.S. ‘Challenger’ 1873–76. Proc. Roy. Soc. Edinburgh, 16, 786–791.Google Scholar
Buizert, C., et al. (2018). Abrupt ice-age shifts in southern westerly winds and Antarctic climate forced from the north. Nature, 563, 681–685.CrossRefGoogle ScholarPubMed
Chiang, J. C. H. (2009). The tropics in paleoclimate. Ann. Rev. Earth Planet. Sci., 37, 263–297. https://doi.org/10.1146/annurev.earth.031208.100217.CrossRefGoogle Scholar
Chiang, J. C. H., et al. (2014). South Pacific Split Jet, ITCZ shifts, and atmospheric North–South linkages during abrupt climate changes of the last glacial period. Earth Planet. Sci. Lett., 406, 233–246.CrossRefGoogle Scholar
Clement, A. C., et al. (1999). Orbital controls on the El Niño/Southern Oscillation and the tropical climate. Paleoceanography, 14(4), 441–456.CrossRefGoogle Scholar
Compagnucci, R. H. (2011). Atmospheric circulation over Patagonia from the Jurassic to present: A review through proxy data and climatic modelling scenarios. Bio. J. Linn. Soc., 103, 229–249.CrossRefGoogle Scholar
Cowell, P. J., et al. (2003). The coastal-tract (part 1): A conceptual approach to aggregated modeling of low-order coastal change. J. Coast. Res., 19, 812–827.Google Scholar
Cronin, T. M. (1999). Principles of Paleoclimate (The Critical Moments and Perspectives in Earth History and Paleobiology). Columbia University Press, New York, 592pp.Google Scholar
Crowley, T. J. (1992). North Atlantic deep water cools the Southern Hemisphere. Paleoceanography, 7(4), 489–497.CrossRefGoogle Scholar
Dansgaard, W., et al. (1982). A new Greenland deep ice core. Science, 218(4579), 1273–1277. https://doi.org/10.1126/science.218.4579.1273.CrossRefGoogle ScholarPubMed
Davis, B. A. S. and Brewer, S. (2009). Orbital forcing and role of the latitudinal insolation/temperature gradient. Clim. Dyn., 32, 143–165. https://doi.org/10.1007/s00382-008-0480-9.CrossRefGoogle Scholar
Deser, C. and Phillips, A. (2017). An overview of decadal-scale sea surface temperature variability from the observational record. CLIVAR Exchanges No. 72, Past Global Changes Magazine, Volume 25, No. 1. https://doi.org/10.22498/pages.25.1.2.CrossRefGoogle Scholar
Deutsche Seewarte. (1885). Segelhandbiicher fur den Atlantischen Ozean (mit Atlas von 36 Karten) [Sailing Handbooks for the Atlantic Ocean (with atlas of 36 charts)]. L. Friederichsen and Co., approx. 100pp.Google Scholar
Deutsche Seewarte. (1892). Segelhandbiicher fur den Indischen Ozean (mit Atlas von 35 Karten) [Sailing Handbooks for the Indian Ocean (with atlas of 35 charts)]. L. Friederichsen and Co., approx. 100pp.Google Scholar
Deutsche Seewarte. (1897). Segelhandbiicher fur den Stillen Ozean (mit Atlas von 31 Karten) [Sailing Handbooks for the Pacific Ocean (with atlas of 31 charts)]. L. Friederichsen and Co., approx. 100pp.Google Scholar
Di Lorenzo, E., et al. (2008). North Pacific Gyre Oscillation links ocean climate and ecosystem change. Geophys. Res. Lett., 35, L08607.CrossRefGoogle Scholar
Ekman, V. W. (1905). On the influence of the Earth’s rotation on ocean currents. Arkiv for Matematik, Astronomi och Fysik, Band, 2(11), 1–51.Google Scholar
Fan, T., et al. (2014). Recent Antarctic sea ice trends in the context of Southern Ocean surface climate variations since 1950. Geophys. Res. Lett., 41, 2419–2426. https://doi.org/10.1002/2014GL059239.CrossRefGoogle Scholar
Farnetti, R., et al. (2014). Pacific interdecadal variability driven by tropical–extratropical interactions. Clim. Dyn., 42(11–12), 3337–3355. https://doi.org/10.1007/s00382-013-1906-6.Google Scholar
Fauchereau, N., et al. (2003). Sea-surface temperature co-variability in the Southern Atlantic and Indian oceans and its connections with the atmospheric circulation in the Southern Hemisphere. Int. J. Clim., 23, 663–677. https://doi.org/10.1002/joc.905.CrossRefGoogle Scholar
Feng, M., et al. (2013). La Niña forces unprecedented Leeuwin Current warming in 2011. Sci. Rep., 2, 1277. https://doi.org/10.1038/srep01277.Google Scholar
Folland, C. K., et al. (2002). Relative influences of the interdecadal pacific oscillation and ENSO on the South Pacific Convergence Zone. Geophys. Res. Lett., 29, 13, 21-1–21-4, 1643. https://doi.org/10.1029/2001GL014201.CrossRefGoogle Scholar
Franzke, C. L. E. (2009). Multi-scale analysis of teleconnection indices: Climate noise and nonlinear trend analysis. Nonlinear Process. Geophys., 16, 65–76,CrossRefGoogle Scholar
Franzke, C. L. E., et al. (2020). The structure of climate variability across scales. Rev. Geophys., 58, e2019RG000657. https://doi.org/10.1029/2019RG000657.CrossRefGoogle Scholar
Gil-Alana, L. (2008). Cyclical long-range dependence and the warming effect in a long temperature time series. Int. J. Clim., 28(11), 1435–1443.CrossRefGoogle Scholar
Gill, A. E. (1982). Atmosphere–Ocean Dynamics, Int. Geophys. Ser., Volume 30, Academic Press, London, 662pp.Google Scholar
Gong, D. and Wang, S. (1999). The definition of the Antarctic oscillation index. Geophys. Res. Lett., 26(4), 459–462.CrossRefGoogle Scholar
Gordon, A. L. (1986). Interocean exchange of thermocline water. J. Geophys. Res., 91, 5037–5046.Google Scholar
Graves, T., et al. (2015). Efficient Bayesian inference for natural time series using ARFIMA processes. Nonlinear Process. Geophys., 22(6), 679.CrossRefGoogle Scholar
Gregory, J. W. (1904). The Climate of Australasia: In Reference to Its Control by the Southern Ocean. Whitcombe and Tombs Ltd, Melbourne, Australia.Google Scholar
Gu, D. and Philander, S. G. H. (1997). Interdecadal climate fluctuations that depend on exchanges between the tropics and the extratropics, Science, 275, 805–807.CrossRefGoogle ScholarPubMed
Hanawa, K. and Talley, L. D. (2001). Mode waters. Chapter 5.4. In Siedler, G. and Church, J. (eds.), Ocean Circulation and Climate. Int. Geophys. Ser., Academic Press, San Diego, pp. 373–386.Google Scholar
Hays, J. D., et al. (1976). Variations in the earth’s orbit: Pacemaker of the ice ages. Science, 194, 1121–1132.Google ScholarPubMed
Hebert, R., et al. (2022). Millennial-scale climate variability over land overprinted by ocean temperature fluctuations. Nat. Geo., 15, 899–905. https://doi.org/10.1038/s41561-022-01056-4.CrossRefGoogle ScholarPubMed
Hernandez, A., et al. (2020). Modes of climate variability: Synthesis and review of proxy-based reconstructions through the Holocene. Earth-Sci. Rev., 209, 103286.CrossRefGoogle Scholar
Hildebrandsson, H. H. (1897). Quelque recherches sure les centres d’action de l’atmosphère. K. Sven. Vetenskaps akad. Handl., 29, 1–33 (1.1).Google Scholar
Holte, J., et al. (2017). An Argo mixed layer climatology and database. Geophys. Res. Lett., 44, 5618–5626. https://doi.org/10.1002/2017GL073426.CrossRefGoogle Scholar
Holton, J. R. (1979). An Introduction to Dynamic Meteorology, 2nd ed. Int. Geophys. Ser., Volume 23, Academic Press, New York, 391pp.Google Scholar
Hu, S. and Federov, A. V. (2018). Cross-equatorial winds control El Niño diversity and change. Nat. Clim. Change, 8(9), 798–802. https://doi.org/10.1038/s41558-018-0248-0.CrossRefGoogle Scholar
Imbrie, J. and Imbrie, K. P. (1979). Ice Ages: Solving the Mystery. Harvard University Press, Cambridge, MA, 224pp.CrossRefGoogle Scholar
IPCC. (2021). Annex IV: Modes of variability. (Cassou, C., et al. (eds.)). In Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change (Masson-Delmotte, V., et al. (eds.)). Cambridge University Press, Cambridge, UK and New York, NY, pp. 2153–2192. https://doi.org/10.1017/9781009157896.018.CrossRefGoogle Scholar
Jain, S., et al. (1999). Seasonality and interannual variations of Northern Hemisphere temperature: Equator-to-pole gradient and land-ocean contrast. J. Clim., 12, 1086–1100.2.0.CO;2>CrossRefGoogle Scholar
Johnson, G. C. and Lyman, J. M. (2022). GOSML: A global ocean surface mixed layer statistical monthly climatology: Means, percentiles, skewness, and kurtosis. J. Geophys. Res., 127, e2021JC018219. https://doi.org/10.1029/2021JC018219.CrossRefGoogle Scholar
Karoly, D. J. (1989). Southern Hemisphere circulation features associated with El Niño-Southern Oscillation events. J. Clim., 2, 1239–1252.2.0.CO;2>CrossRefGoogle Scholar
Karoly, D. J. (1990). The role of transient eddies in low-frequency zonal variations of the Southern Hemisphere circulation. Tellus, 42A, 41–50.Google Scholar
Kataoka, T., et al. (2014). On the Ningaloo Niño/ Niña. Clim. Dyn., 43, 1463–1482.CrossRefGoogle Scholar
Kidson, E. (1928). British Antarctic Expedition 1907–1909. Meteorology. Rep. Sci. Investigations. H.J. Green, Melbourne, Australia, 188pp.Google Scholar
Kidson, E. (1946). Australian Antarctic Expedition 1911–14, Meteorology. Discussion of observations at Adelie Land, Queen Mary Land and Macquarie Island. Sci. Rep. Ser., B (VI), 121pp.Google Scholar
Lang, A. L., et al. (2020). Introduction to special collection: ‘Bridging weather and climate: Subseasonal-to-seasonal (S2S) prediction’. J. Geophys. Res. Atmos., 125, e2019JD031833. https://doi.org/10.1029/2019JD031833.CrossRefGoogle Scholar
Latif, M., et al. (2013). Southern Ocean sector centennial climate variability and recent decadal trends. J. Clim., 26, 7767–7782. https://doi.org/10.1175/JCLI-D-12-00281.1.CrossRefGoogle Scholar
Levy, R., et al. (2019). Antarctic ice-sheet sensitivity to obliquity forcing enhanced through ocean connections. Nat. Geo., 12, 132–137. https://doi.org/10.1038/s41561-018-0284-4.CrossRefGoogle Scholar
Lewis, J. M. (1996). Winds over the world sea: Maury and Köppen. Bull. Am. Meterol. Soc., 77(5), 935–952.2.0.CO;2>CrossRefGoogle Scholar
Li, X., et al. (2015). Atlantic-induced pan-tropical climate change over the past three decades. Nat. Clim. Change, 6, 275–280. https://doi.org/10.1038/NCLIMATE2840.Google Scholar
Li, Z., et al. (2021). The origin and fate of Subantarctic Mode Water in the Southern Ocean. J. Phys. Oceanog., 51, 2951–2972.Google Scholar
de Lima, M. I. P. and Lovejoy, S. (2015). Macroweather precipitation variability up to global and centennial scales. Water Resour. Res., 51. https://doi.org/10.1002/2015WR017455.CrossRefGoogle Scholar
Limpasuvan, V. and Hartmann, D. L. (1999). Eddies and the annular modes of climate variability. Geophys. Res. Lett., 26, 3133–3136.CrossRefGoogle Scholar
Limpasuvan, V. and Hartmann, D. L. (2000). Wave-maintained annular modes of climate variability. J. Clim., 13, 4414–4429.2.0.CO;2>CrossRefGoogle Scholar
Liu, Z. and Alexander, M. (2007). Atmospheric bridge, oceanic tunnel, and global climatic teleconnections. Rev. Geophys., 45, RG2005. https://doi.org/10.1029/2005RG000172.CrossRefGoogle Scholar
Lockyer, N. and Lockyer, W. J. S. (1902). On the similarity of the short-period pressure variations over large areas. Proc. Roy. Soc., London, 71, 134–135.Google Scholar
Lockyer, N. and Lockyer, W. J. S. (1904). The behaviour of short-period atmospheric pressure variations over the earth’s surface. Proc. Roy. Soc., London, 73, 457–470.Google Scholar
Lockyer, W. J. S. (1909). A Discussion of Australian Meteorology, being a study of the pressure, rainfall and river changes, both seasonal and from year to year, together with a comparison of the air movements over Australia with those over South Africa and South America. Solar Phys. Committee, 117pp.Google Scholar
Lockyer, W. J. S. (1910). Does the Indian climate change? Nature, 84, 178.CrossRefGoogle Scholar
Lovejoy, S. (2018). Spectra, intermittency, and extremes of weather, macroweather and climate. Sci. Rep., 8(1), 12697. https://doi.org/10.1038/s41598-018-30829-4.CrossRefGoogle ScholarPubMed
Lovejoy, S. (2019). Weather, Macroweather and the Climate: Our Random Yet Predictable Atmosphere. Oxford University Press, New York.CrossRefGoogle Scholar
Lovejoy, S. and Schertzer, D. (2013). The Weather and Climate: Emergent Laws and Multifractal Cascades. Cambridge University Press, Cambridge, 660pp.CrossRefGoogle Scholar
Lovejoy, S. and Varotsos, C. (2016). Scaling regimes and linear/nonlinear responses of last millennium climate to volcanic and solar forcings. Earth Sys. Dyn., 7(1), 133–150.Google Scholar
Lovejoy, S., et al. (2018). Harnessing butterflies: Theory and practice of the stochastic seasonal to interannual prediction system (StocSIPS). In Tsonis, A. A. (eds.), Advances in Nonlinear Geosciences. Springer International Publishing, pp. 305–355. https://doi.org/10.1007/978-3-319-58895-7_29.Google Scholar
Lovejoy, S. and Lambert, F. (2019). Spiky fluctuations and scaling in high-resolution EPICA ice core dust fluxes. Clim. Past, 15, 1999–2017. https://doi.org/10.5194/cp-15-1999-2019.CrossRefGoogle Scholar
Madden, R. and Julian, P. (1994). Observations of the 40–50 day tropical oscillation – A review. Mon. Weather Rev., 122, 814–837.2.0.CO;2>CrossRefGoogle Scholar
Mann, M., et al. (2014). On forced temperature changes, internal variability, and the AMO. Geophys. Res. Lett., 41(9), 3211–3219. https://doi.org/10.1002/2014GL059233.CrossRefGoogle Scholar
Mann, M., et al. (2021). Multidecadal climate oscillations during the past millennium driven by volcanic forcing. Science, 371(6533), 1014–1019. https://doi.org/10.1126/science.abc5810.CrossRefGoogle ScholarPubMed
Mantsis, D. F., et al. (2013). Precessional cycles and their influence on the North Pacific and North Atlantic summer anticyclones. J. Clim., 26, 4596–4611. https://doi.org/10.1175/JCLI-D-12-00343.1.CrossRefGoogle Scholar
Mantua, N. J., et al. (1997). A Pacific interdecadal climate oscillation with impacts on salmon production. Bull. Am. Meterol. Soc., 78, 1069–1079.2.0.CO;2>CrossRefGoogle Scholar
Markle, B. R., et al. (2017). Global atmospheric teleconnections during Dansgaard-Oeschger events. Nat. Geo., 10, 36–40.CrossRefGoogle Scholar
Maury, M. F. (1848–1860). Wind and Current Charts. U.S. Hydrographic Office, Manuscripts Division, Library of Congress, Washington, DC.Google Scholar
Maury, M. F. (1854). Maritime conference held at Brussels for devising a uniform system of meteorological observations at sea, August and September, 1853. In Explanations and Sailing Directions to Accompany the Wind and Weather Current Charts, 6th ed. Maury, M. F. and Biddle, J. Publishing, Philadelphia, pp. 54–96.Google Scholar
Maury, M. F. (1861). The Physical Geography of the Sea and Its Meteorology, 10th ed. Sampson Low and Son, London, UK, 457pp.CrossRefGoogle Scholar
McGregor, S., et al. (2014). Recent Walker circulation strengthening and Pacific cooling amplified by Atlantic warming. Nat. Clim. Change, 4(10), 888–892. https://doi.org/10.1038/NCLIMATE2330.CrossRefGoogle Scholar
Meehl, G. A., et al. (2016). Antarctic sea-ice expansion between 2000 and 2014 driven by tropical Pacific decadal climate variability. Nat. Geosci., 9, 590–595. https://doi.org/10.1038/NGEO2751.CrossRefGoogle Scholar
Mo, K. and Higgins, R. (1998). The Pacific-South American modes and tropical convection during the southern hemisphere winter. Mon. Weather Rev., 126, 1581–1596.2.0.CO;2>CrossRefGoogle Scholar
Mo, K. and Paegle, J. (2001). The Pacific-South American modes and their downstream effects. Int. J. Clim., 21, 1211–1229.CrossRefGoogle Scholar
Morioka, Y., et al. (2014). Role of tropical SST variability on the formation of subtropical dipoles. J. Clim., 27, 4486–4507.CrossRefGoogle Scholar
Mossman, R. C. (1913). Southern Hemisphere seasonal correlations. Symon’s Meteorol. Mag., 48, 1–34.Google Scholar
Mossman, R. C. (1916). The physical conditions of the Weddell Sea. Geogr. J., 48, 479–500.CrossRefGoogle Scholar
Mossman, R. C. (1918). The climate and meteorology of Antarctic and sub–Antarctic regions. J. Scottish Meteorol. Soc., 35, 18–29.Google Scholar
Munk, W. H. (1950). On the wind-driven ocean circulation. J. Atmos. Sci., 7, 80–93.Google Scholar
Neelin, J. D. (2011). Climate Change and Climate Modelling. Cambridge University Press, Cambridge, 282pp.Google Scholar
Neelin, J. D., et al. (1998). ENSO theory. J. Geophys. Res., 103, 14261–14290.Google Scholar
Newman, M., et al. (2016). The Pacific Decadal Oscillation, revisited. J. Clim., 29(12), 4399–4427.CrossRefGoogle Scholar
O’Reilly, C. H., et al. (2016). The signature of low-frequency oceanic forcing in the Atlantic Multidecadal Oscillation. Geophys. Res. Lett., 43(6), 2810–2818. https://doi.org/10.1002/2016GL067925.Google Scholar
Pedro, J. B., et al. (2018). Beyond the bipolar seesaw: Toward a process understanding of interhemispheric coupling. Quat. Sci. Rev., 192, 27–46. https://doi.org/10.1016/j.quascirev.2018.05.005.CrossRefGoogle Scholar
Peixoto, J. P. and Oort, A. H. (1992). Physics of Climate. Springer, New York.CrossRefGoogle Scholar
Philander, S. G. H. (1990). El Niño, La Niña, and the Southern Oscillation. Academic, San Diego, CA.Google Scholar
Power, S., et al. (2021). Decadal climate variability in the tropical Pacific: Characteristics, causes, predictability, and prospects. Science, 374, 48, eaay9165. https://doi.org/10.1126/science.aay9165.CrossRefGoogle ScholarPubMed
Putnam, A. E. (2015). A glacial zephyr. Paleoclimate News and Views. Nat. Geo., 8, 175–176. https://doi.org/10.1038/ngeo2377.CrossRefGoogle Scholar
Ramstein, G., et al. (eds.). (2021). Paleoclimatology. Front. Earth Sci. https://doi.org/10.1007/978-3-030-24982-3.CrossRefGoogle Scholar
Raymo, M. E. and Nisancioglu, K. (2003). The 41 kyr world: Milankovitch’s other unsolved mystery. Paleoceanography, 18(1), 1011. https://doi.org/10.1029/2002PA000791.CrossRefGoogle Scholar
Richter, I., et al. (2010). On the triggering of Benguela Niños: Remote equatorial versus local influences. Geophys. Res. Lett., 37, L20604. https://doi.org/10.1029/2010GL044461.CrossRefGoogle Scholar
Ridgway, K. R. and Dunn, J. R. (2007). Observational evidence for a Southern Hemisphere oceanic supergyre. Geophys. Res. Lett., 34, L13612. https://doi.org/10.1029/2007GL030392.CrossRefGoogle Scholar
Rind, D. (1998). Latitudinal temperature gradients and climate change. J. Geophys. Res., 103, 5943–5971.Google Scholar
Risien, C. M. and Chelton, D. B. (2008). A global climatology of surface wind and Wind stress fields from eight years of QuikSCAT scatterometer data. J. Phys. Oceanogr., 38, 2379–2413.CrossRefGoogle Scholar
Ruddiman, W. F. (2001). Earth’s Climate, Past and Future. W.H. Freeman and Co., New York, 465pp.Google Scholar
Rypdal, M. and Rypdal, K. (2016). Late quaternary temperature variability described as abrupt transitions on a 1∕f noise background. Earth Sys. Dyn., 7(1), 281–293.Google Scholar
Saji, N. H., et al. (1999). A dipole mode in the tropical Indian Ocean. Nature, 401, 360–363.CrossRefGoogle ScholarPubMed
Saltzman, B. (2001). Dynamical Paleoclimatology: Generalized Theory of Global Climate Change. Elsevier Science Publishing Co Inc, US, 354pp.Google Scholar
Schlesinger, M. E. (1994). An oscillation in the global climate system of period 65–70 years. Nature, 367(6465), 723–726. https://doi.org/10.1038/367723a0.CrossRefGoogle Scholar
Schmitz, W. J. (1996a). On the World Ocean Circulation: Volume I: Some global features/North Atlantic circulation. Woods Hole Oceanographic Institution Technical Report, WHOI-96-03, Woods Hole, MA, 141pp.Google Scholar
Schmitz, W. J. (1996b). On the World Ocean Circulation: Volume II: The Pacific and Indian Ocean circulation/Global Update. Woods Hole Oceanographic Institution Technical Report, WHOI-96-08, Woods Hole, MA, 237pp.Google Scholar
Shannon, L. V., et al. (1986). On the existence of an El Niño-type phenomenon in the Benguela system. J. Mar. Res., 44(3), 495–520.CrossRefGoogle Scholar
Shao, Z-G. and Ditlevsen, P. D. (2016). Contrasting scaling properties of interglacial and glacial climates. Nat. Comms., 7, 10951. https://doi.org/10.1038/ncomms10951.CrossRefGoogle ScholarPubMed
Simpson, G. C. (1919a). British Antarctic Expedition 19101913, Meteorology, volume 1, Discussion. Harrison and Sons, London, 326pp.Google Scholar
Simpson, G. C. (1919b). British Antarctic Expedition 19101913, Meteorology, volume 2, Weather Maps and Pressure Curves. Harrison and Sons, London.Google Scholar
Skinner, L., et al. (2020). Southern Ocean convection amplified past Antarctic warming and atmospheric CO2 rise during Heinrich Stadial 4. Comms. Earth Env., 1(1). https://doi.org/10.1038/s43247-020-00024-3.Google Scholar
Stocker, T. F. and Johnsen, S. J. (2003). A minimum thermodynamic model for the bipolar seesaw. Paleoceanography, 18, 1087. https://doi.org/10.1029/2003PA000920.CrossRefGoogle Scholar
Stommel, H. (1948). The westward intensification of wind- driven currents. Trans. Am. Geophys. Union, 29, 202–206.Google Scholar
Sun, D.-Z. and Bryan, F. (eds.). (2010). Climate dynamics: Why does climate vary? Geophys. Monographs, 189, Amer. Geophys. Union, 216pp.CrossRefGoogle Scholar
Sverdrup, H. U. (1947). Wind-driven currents in a baroclinic ocean. Proc. Nat. Ac. Sci. USA, PNAS, 33, 318–326.Google Scholar
Talley, L. D., et al. (2011). Descriptive Physical Oceanography: An Introduction, 6th ed. Academic Press. https://doi.org/10.1016/C2009-0-24322-4.CrossRefGoogle Scholar
Taylor, G. (1928). Climatic relations between Antarctica and Australia. In Joerg, W. L. G. (ed) Problems of Polar Research. Am. Geogr. Soc., Special Publication 7, pp. 285–299.Google Scholar
Teisserenc de Bort, L. (1883). Etude sur l’hiver de 1879–80 et recherches sur l’influence de la position des grands centres d’action de l’atmosphère dans les hivers anormaux. Ann. Soc. Meteor. France, 31, 70–79.Google Scholar
Thompson, D. W. J. and Wallace, J. M. (2000). Annular modes in the extratropical circulation. Part I: Month-to-month variability. J. Clim., 13, 1000–1016.Google Scholar
Thompson, D. W. J. and Solomon, S. (2002). Interpretation of recent Southern Hemisphere climate change. Science, 296, 895–899.CrossRefGoogle ScholarPubMed
Timmermann, A., et al. (2007). The effect of orbital forcing on the mean climate and variability of the Tropical Pacific. J. Clim., 20, 4147–4159. https://doi.org/10.1175/JCLI4240.1.CrossRefGoogle Scholar
Trenberth, K. E. and Shea, D. J. (2006). Atlantic hurricanes and natural variability in 2005. Geophys. Res. Lett., 33, L12704. https://doi.org/10.1029/2006GL026894.CrossRefGoogle Scholar
Troup, A. J. (1965). The Southern Oscillation. Q. J. R. Meteorol. Soc., 91, 490–506.CrossRefGoogle Scholar
Tsonis, A. A. (2018). Insights in climate dynamics from climate networks. In, Tsonis, A. A. (ed) Advances in Nonlinear Geosciences. Springer International Publishing, Switzerland, pp. 631–649. https://doi.org/10.1007/978-3-319-58895-7_29.CrossRefGoogle Scholar
Tsonis, A. A., et al. (2008). On the role of atmospheric teleconnection in climate. J. Clim., 21, 2990–3001.CrossRefGoogle Scholar
Tsubouchi, T, Suga, T and Hanawa, K. (2016). Comparison study of subtropical mode waters in the world ocean. Front. Mar. Sci., 3, 270. https://doi.org/10.3389/fmars.2016.00270.CrossRefGoogle Scholar
Turney, C. S. M. and Jones, R. T. (2010). Does the Agulhas Current amplify global temperatures during super-interglacials? J. Quat. Sci., 25, 839–843. ISSN 0267-8179.CrossRefGoogle Scholar
Vallis, G. K. (2017). Atmospheric and Oceanic Fluid Dynamics: Fundamentals and Large-scale Circulation. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Vimeux, F., et al. (1999). Glacial-interglacial changes in ocean surface conditions in the Southern Hemisphere. Nature, 398, 410–413.CrossRefGoogle Scholar
Walker, G. T. (1923). Correlation in seasonal variations of weather VIII. Mem. India Meteorol. Dept., 24, Part IV, 75–131.Google Scholar
Walker, G. T. (1924). Correlations in seasonal variations of weather. IX. Mem. India Meteorol. Dept., 24(4), 275–332.Google Scholar
Walker, G. T. (1928). World weather. Q. J. Meteorol. Soc., 54(226), 79–87.CrossRefGoogle Scholar
Walker, G. T. and Bliss, E. W. (1932). World weather V. Mem. India Meteorol. Dept., 4(36), 53–84.Google Scholar
Walker, G. T. and Bliss, E. W. (1936). World weather VI. Mem. India Meteorol. Dept., 4(39), 119–139.Google Scholar
Wallace, J. M. (2025). Gilbert T Walker’s Enduring Studies of Climate Variability. IISc Centenary Series, Vol. 6. IISc Press and World Scientific Publishing, Singapore, 385pp, https://doi.org/10.1142/13888.CrossRefGoogle Scholar
Wallace, J. M. and Gutzler, D. S. (1981). Teleconnections in the geopotential height field during the Northern Hemisphere winter. Mon. Weather Rev., 109, 784–812.2.0.CO;2>CrossRefGoogle Scholar
Weijer, W., et al. (2019). Stability of the Atlantic Meridional overturning circulation: A review and synthesis. J. Geophys. Res. Oceans, 124, 5336–5375. https://doi.org/10.1029/2019JC015083.CrossRefGoogle Scholar
Williams, P. D., et al. (2017). A census of atmospheric variability from seconds to decades. Geophys. Res. Lett., 44, 11201–11211. https://doi.org/10.1002/2017GL075483.CrossRefGoogle Scholar
Wunsch, C. (2003). The spectral description of climate change including the 100 ky energy. Clim. Dyn., 20(4), 353–363.CrossRefGoogle Scholar
Wyrtki, K. and Meyers, G. (1975a). The trade wind field over the Pacific Ocean Part I: The mean field and the mean annual variation. Univ. Hawaii, Tech. Rept. HIG-75-1, 26pp.Google Scholar
Wyrtki, K. and Meyers, G. (1975b). The trade wind field over the Pacific Ocean Part II: Bimonthly fields of wind stress: 1950 to 1972. Univ. Hawaii, Tech. Rept. HIG-75-2, 16pp.Google Scholar
Yamagata, T., et al. (2016). Old and new faces of climate variations. Chapter 1. In Behera, S. K. and Yamagata, T. (eds.), Indo-Pacific Climate Variability and Predictability. World Scientific Publishing Co., Singapore, pp. 1–23.Google Scholar
Yeh, S. W., et al. (2009). El Niño in a changing climate. Nature, 461, 511–514. https://doi.org/10.1038/nature08316.CrossRefGoogle Scholar
Zebiak, S. E. (1993). Air-Sea interaction in the equatorial Atlantic region. J. Clim., 6, 1567–1586.2.0.CO;2>CrossRefGoogle Scholar
Zebiak, S. E. and Cane, M. A. (1987). A model El Niño–Southern Oscillation. Mon. Weather Rev., 115, 2262–2278.2.0.CO;2>CrossRefGoogle Scholar
Amaya, D. J. (2019). The Pacific Meridional Mode and ENSO: A review. Curr. Clim. Change Rep., 5, 296–307. https://doi.org/10.1007/s40641-019-00142-x.CrossRefGoogle Scholar
An, S.-I. and Kim, J.-W. (2017). Role of nonlinear ocean dynamic response to wind on the asymmetrical transition of El Niño and La Niña. Geophys. Res. Lett., 44(1), 393–400. https://doi.org/10.1002/2016GL071971.CrossRefGoogle Scholar
An, Z., et al. (2015). Global monsoon dynamics and climate change. Annu. Rev. Earth Planet. Sci., 43, 29–77. https://doi.org/10.1146/annurev-earth-060313-054623.Google Scholar
An, Z., et al. (2021). ENSO Irregularity and Asymmetry. Chapter 7. In McPhaden, M. J., et al. (eds.), El Niño Southern Oscillation in a Changing Climate. Geophysical Monograph 253. American Geophysical Union and John Wiley & Sons, New York, pp. 153–172.Google Scholar
Ashok, K., et al. (2007). El Niño Modoki and its possible teleconnection. J. Geophys. Res. Oceans, 112, C11007.CrossRefGoogle Scholar
Ashok, K. and Yamagata, T. (2009). The El Niño with a difference. Nature, 461, 481–484.CrossRefGoogle ScholarPubMed
Ball, F. K. (1956).The theory of strong katabatic winds. Aust. J. Phys., 9, 373–386.CrossRefGoogle Scholar
Ball, F. K. (1960). Winds on the ice slope of Antarctica. Antarctic Meteorology, Proceedings of the Symposium of Melbourne, Pergamon, pp. 9–16.Google Scholar
Barry, R. G. and Carleton, A. M. (2001). Synoptic and Dynamic Climatology. Routledge, London, UK, 620pp.Google Scholar
Barry, R. G. and Chorley, R. J. (2003). Atmosphere, Weather and Climate, 8th ed. Routledge, London, 421pp.Google Scholar
Behera, S. K. and Yamagata, T. (2001). Subtropical SST dipole events in the southern Indian Ocean. Geophys. Res. Lett., 28, 327–330.CrossRefGoogle Scholar
Boers, N., et al. (2014). The South American rainfall dipole: A complex network analysis of extreme events. Geophys. Res. Lett., 41(20), 7397–7405. https://doi.org/10.1002/2014GL061829.CrossRefGoogle Scholar
Bromwich, D. H., et al. (1993). Spatial and temporal characteristics of the intense katabatic winds at Terra Nova Bay, Antarctica. In Bromwich, D. H. and Stearns, C. R. (eds.), Antarctic Meteorology and Climatology: Studies Based on Automatic Weather Stations, Antarctic Research Series 61, American Geophysical Union, Washington, DC, pp. 47–68.CrossRefGoogle Scholar
Bromwich, D. H., et al. (2011). Climatological aspects of cyclogenesis near Adélie Land Antarctica. Tellus, 63A, 921–938. https://doi.org/10.1111/j.1600-0870.2011.00537.x.Google Scholar
Cai, W., et al. (2011a). Influence of global-scale variability on the Subtropical Ridge over Southeast Australia. J. Clim., 24, 6035–6053.CrossRefGoogle Scholar
Cai, W., et al. (2011b). Teleconnection pathways of ENSO and the IOD and the mechanisms for impacts on Australian rainfall. J. Clim., 24, 3910–3923.CrossRefGoogle Scholar
Capotondi, A., et al. (2021). ENSO Diversity. Chapter 4. In McPhaden, M. J., et al. (eds.), El Niño Southern Oscillation in a Changing Climate. Geophysical Monograph 253, American Geophysical Union and John Wiley & Sons, New York, pp. 65–86.Google Scholar
Carvalho, L. M. V de and Cavalcanti, I. F. A. (2016). The South American Monsoon System (SAMS). Chapter 6. In Carvalho, L. M. V. de and Jones, C. (eds.), The Monsoons and Climate Change. Springer Climate, pp. 121–148. https://doi.org/10.1007/978-3-319-21650-8_6.CrossRefGoogle Scholar
Caton-Harrison, T., et al. (2022). Reanalysis representation of low-level winds in the Antarctic near-coastal region. Weather Clim. Dynam., 3, 1415–1437.CrossRefGoogle Scholar
Chemke, R. (2021). The future poleward shift of Southern Hemisphere summer mid-latitude storm tracks stems from ocean coupling. Nat. Commun., 13, 1730. https://doi.org/10.1038/s41467-022-29392-4.Google Scholar
Chiang, J. C. H. (2009). The tropics in paleoclimate. Annu. Rev. Earth Planet. Sci., 37, 263–297. https://doi.org/10.1146/annurev.earth.031208.100217.CrossRefGoogle Scholar
Claud, C., et al. (2009). Southern Hemisphere winter cold-air mesocyclones: Climatic environments and associations with teleconnections. Clim. Dyn., 33, 383–408. https://doi.org/10.1007/s00382-008-0468-5.CrossRefGoogle Scholar
Clement, A. C., et al. (1999). Orbital controls on the El Niño-Southern Oscillation and the tropical climate. Paleoceanogr., 14, 441–456.CrossRefGoogle Scholar
Clement, A. C., et al. (2000). Suppression of El Niño during the mid-Holocene by changes in the Earth’s orbit. Paleoceanography, 15, 731–737.CrossRefGoogle Scholar
Clement, A. C., et al. (2004). The importance of precessional signals in the tropical climate. Clim. Dyn., 22, 327–341.CrossRefGoogle Scholar
Dai, A. and Wigley, T. M. L. (2000). Global patterns of ENSO-induced precipitation. Geophys. Res. Lett., 27, 1283–1286.CrossRefGoogle Scholar
de Szoeke, R. A. and Levine, M. D. (1981). The advective flux of heat by mean geostrophic motions in the Southern Ocean. Deep-Sea Res., 28, 1057–1085.CrossRefGoogle Scholar
Deser, C., et al. (2010). Sea surface temperature variability: Patterns and mechanisms. Annu. Rev. Mar. Sci., 2, 115–143. https://doi.org/10.1146/annurev-marine-120408-151453.CrossRefGoogle ScholarPubMed
Diaz, H. F., et al. (2001). ENSO variability, teleconnections and climate change. Int. J. Climatol., 21, 1845–1862. https://doi.org/10.1002/joc.631.CrossRefGoogle Scholar
Dommenget, D., et al. (2013). Analysis of the non-linearity in the pattern and time evolution of El Niño southern oscillation. Clim. Dyn., 40, 2825–2847. https://doi.org/10.1007/s00382-012-1475-0.CrossRefGoogle Scholar
Dong, X., et al. (2020). Robustness of the recent global atmospheric reanalyses for Antarctic near-surface wind speed climatology. J. Climate, 33, 4027–4043.CrossRefGoogle Scholar
England, M. H., et al. (2014). Recent intensification of wind-driven circulation in the Pacific and the ongoing warming hiatus. Nat. Clim. Change. https://doi.org/10.1038/NCLIMATE2106.CrossRefGoogle Scholar
Fedorov, A. V., et al. (2021). ENSO low-frequency modulation and mean state interactions. Chapter 8. In McPhaden, M. J., et al. (eds.), El Niño Southern Oscillation in a Changing Climate. Geophysical Monograph 253, American Geophysical Union and John Wiley & Sons, New York, pp. 173–200.Google Scholar
Fogt, R. L. and Bromwich, D. H. (2006). Decadal variability of the ENSO teleconnection to the high-latitude South Pacific governed by coupling with the southern annular mode. J. Clim., 19(6), 979–997.CrossRefGoogle Scholar
Fogt, R. L., et al. (2009). Historical SAM variability. Part II: Twentieth-century variability and trends from reconstructions, observations, and the IPCC AR4 models. J. Clim., 22, 5346–5365. https://doi.org/10.1175/2009JCLI2786.1.CrossRefGoogle Scholar
Fogt, R. L. and Marshall, G. J. (2020). The Southern Annular Mode: Variability, trends, and climate impacts across the Southern Hemisphere. WIREs Clim Change, 11, e652. https://doi.org/10.1002/wcc.652.CrossRefGoogle Scholar
Funk, S., et al. (2016). The East African monsoon system: Seasonal climatologies and recent variations. Chapter 8. In Carvalho, L. M. V. de and Jones, C. (eds.), The Monsoons and Climate Change. Springer Climate, Switzerland, pp. 163–185. https://doi.org/10.1007/978-3-319-21650-8_2.Google Scholar
Garreaud, R. D. and Wallace, J. M. (1998). Summertime incursions of mid-latitude air into tropical and subtropical South America. Mon. Wea. Rev., 126, 2713–2733.2.0.CO;2>CrossRefGoogle Scholar
Garreaud, R. D., et al. (2009). Present-day South American climate. Palaeogeogr. Palaeoclim. Palaeoecol., 281, 180–195.CrossRefGoogle Scholar
Geller, M. A., Zhou, T., and Yuan, W. (2016). The QBO, gravity waves forced by tropical convection, and ENSO. J. Geophys. Res: Atmos., 121, 8886–8895. https://doi.org/10.1002/2015JD024125.CrossRefGoogle Scholar
Glickman, T. S. (2000). Glossary of Meteorology, 2nd ed. American. Meteorological Society, Boston, 850pp.Google Scholar
Goodwin, I. D., et al. (2004). Mid latitude winter climate variability in the South Indian and southwest Pacific regions since 1300 AD. Clim. Dyn., 22, 783–794.CrossRefGoogle Scholar
Goodwin, I. D., et al. (2014). A reconstruction of extratropical Indo-Pacific sea-level pressure patterns during the Medieval Climate Anomaly. Clim. Dyn., 43(5–6), 1197–1219. https://doi.org/10.1007/s00382-013-1899-1.CrossRefGoogle Scholar
Grist, J. P. and Nicholson, S. E. (2001). A study of the dynamic factors influencing the rainfall variability in the West African Sahel. J. Clim., 14, 1337–1359.2.0.CO;2>CrossRefGoogle Scholar
Grodsky, S. A. and Carton, J. A. (2003). The Intertropical Convergence Zone in the South Atlantic and the Equatorial Cold Tongue. J. Clim., 16, 723–733.2.0.CO;2>CrossRefGoogle Scholar
Grodsky, S. A., et al. (2003). Near surface westerly wind jet in the Atlantic ITCZ. Geophys. Res. Lett., 30(19), 2009. https://doi.org/10.1029/2003GL017867.CrossRefGoogle Scholar
Hall, A. and Visbeck, M. (2002). Synchronous variability in the Southern Hemisphere atmosphere, sea ice, and ocean resulting from the annular mode. J. Clim., 15, 3043–3057.2.0.CO;2>CrossRefGoogle Scholar
Hallberg, R. and Gnanadesikan, A. (2006). The role of eddies in determining the structure and response of the wind-driven Southern Hemisphere overturning: Results from the Modeling Eddies in the Southern Ocean (MESO) project. J. Phys. Oceanogr., 36, 2232–2252. https://doi.org/10.1175/JPO2980.1.CrossRefGoogle Scholar
Hastenrath, S. (1991). Climate Dynamics of the Tropics. Kluwer, Amsterdam, 488pp.CrossRefGoogle Scholar
Hastenrath, S. and Heller, L. (1977). Dynamics of climatic hazards in Northeast Brazil. Q. J. R. Meteorol. Soc., 103, 77–92.CrossRefGoogle Scholar
Hazel, J. E. and Stewart, A. L. (2019). Are the near-Antarctic easterly winds weakening in response to enhancement of the Southern Annular Mode? J. Climate, 32, 1895–1918.CrossRefGoogle Scholar
Held, I. M. and Hou, A. Y. (1980). Nonlinear axially symmetric circulations in a nearly inviscid atmosphere. J. Atmos. Sci., 37, 515–533.2.0.CO;2>CrossRefGoogle Scholar
Henley, B. J., et al. (2015). A tripole index for the Interdecadal Pacific Oscillation. Clim. Dyn., 45, 3077–3090. https://doi.org/10.1007/s00382-015-2525-1.CrossRefGoogle Scholar
Hobbs, W. R. and Raphael, M. N. (2010). Characterizing the zonally asymmetric component of the SH circulation. Clim. Dyn., 35, 859–873. https://doi.org/10.1007/s00382-009-0663-z.CrossRefGoogle Scholar
Hsu, P. C. (2016). Global monsoon in a changing climate. Chapter 2. In Carvalho, L. M. V. de and Jones, C. (eds.), The Monsoons and Climate Change. Springer Climate, Switzerland, pp. 7–24. https://doi.org/10.1007/978-3-319-21650-8_2.Google Scholar
Hurrell, J. W. and van Loon, H. (1994). A modulation of the atmospheric annual cycle in the Southern Hemisphere. Tellus, 46A, 325–338.Google Scholar
Hurrell, J. W., van Loon, H. and Shea, D. J. (1998). The mean state of the troposphere. Chapter 1. In Karoly, D. J. and Vincent, D. G. (eds.) Meteorology of the Southern Hemisphere. American Meteorological Society, Boston, MA, pp. 1–46.Google Scholar
Huth, R. and Beranová, R. (2021). How to recognize a true mode of atmospheric circulation variability. Earth Space Sci., 8, e2020EA001275. https://doi.org/10.1029/2020EA001275.CrossRefGoogle Scholar
Jiang, X., et al. (2020). Fifty years of research on the Madden-Julian Oscillation: Recent progress, challenges, and perspectives. J. Geophys. Res. Atmospheres, 125, e2019JD030911. https://doi.org/10.1029/2019JD030911.CrossRefGoogle Scholar
Jones, J. M., et al. (2009). Historical SAM Variability. Part I: Century-length seasonal reconstructions. J. Clim., 22, 5319–5345. https://doi.org/10.1175/2009JCLI2785.1.CrossRefGoogle Scholar
Karnauskas, K. B. and Ummenhofer, C. C. (2014). On the dynamics of the Hadley circulation and subtropical drying. Clim. Dyn., 42, 2259–2269. https://doi.org/10.1007/s00382-014-2129-1.CrossRefGoogle Scholar
Karoly, D. J. (1998). General circulation. Chapter 2. In Karoly, D. J. and Vincent, D. G. (eds.), Meteorology of the Southern Hemisphere. American Meteorological Society, Boston, MA, pp. 47–87.CrossRefGoogle Scholar
Kiladis, G. N. and Mo, K. C. (1998). Interannual and intraseasonal variability in the Southern Hemisphere. In Karoly, D. J. and Vincent, D. G. (eds), Meteorology of the Southern Hemisphere. American Meteorological Society, Boston, p. 410.Google Scholar
King, J. C. and Turner, J. (1997). Antarctic Meteorology and Climatology. Cambridge University Press, Cambridge, U.K., 409pp. https://doi.org/10.1017/CBO9780511524967.CrossRefGoogle Scholar
Kug, J. S., et al. (2021). ENSO remote forcing: Influence of climate variability outside the tropical Pacific. Chapter 11. In McPhaden, M. J. et al. (eds.), El Niño Southern Oscillation in a Changing Climate. Geophysical Monograph 253, American Geophysical Union and John Wiley & Sons, New York, pp. 249–266.Google Scholar
Langlais, C. E., et al. (2015). Sensitivity of Antarctic Circumpolar Current transport and eddy activity to wind patterns in the Southern Ocean. J. Phys. Oceanogr., 45, 1051–1067.CrossRefGoogle Scholar
Larkin, N. K. and Harrison, D. E. (2002). ENSO warm (El Niño) and cold (La Niña) event life cycles: Ocean surface anomaly patterns, their symmetries, asymmetries, and implications. J. Clim., 15, 1118–1140.2.0.CO;2>CrossRefGoogle Scholar
Levitus, S., et al. (2005). Warming of the world ocean, 1955–2003. Geophys. Res. Lett., 32, L02604. https://doi.org/10.1029/2004GL021592.CrossRefGoogle Scholar
Lisonbee, J., et al. (2020). Defining the north Australian monsoon onset: A systematic review. Prog. Phys. Geogr., 44, 3, 398–418.CrossRefGoogle Scholar
Loewe, F. (1956). Etudes de glaciologie en Terre Adelie 1951–52, Expeditions Polaires Francaises, 1948- Travaux,; no.9, Actualites scientifiques et industrielles; no.1247. Paris: Hermann, 1956 159p.Google Scholar
Lovenduski, N. S. and Gruber, N. (2005). Impact of the Southern Annular Mode on Southern Ocean circulation and biology. Geophys. Res. Lett., 32, L11603. https://doi.org/10.1029/2005GL022727.CrossRefGoogle Scholar
Madden, R. A. and Julian, P. R. (1971). Detection of a 40–50 day oscillation in zonal wind in tropical Pacific. J. Atmos. Sci., 28, 702.2.0.CO;2>CrossRefGoogle Scholar
Madden, R. A., & Julian, P. R. (1972). Description of global-scale circulation cells in tropics with a 40–50 day period. J. Atmos. Sci., 29, 1109.2.0.CO;2>CrossRefGoogle Scholar
Mantua, N. J., et al. (1997). A Pacific Interdecadal Climate Oscillation with impacts on salmon production. Bull. Am. Meteorol. Soc., 78(6), 1069–1079. http://doi.org/10.1175/1520-0477(1997)078<1069:apicow>2.0.co;2.2.0.CO;2>CrossRefGoogle Scholar
Marshall, G. J. (2003). Trends in the Southern Annular Mode from observations and reanalyses. J. Clim., 16, 4134–4143.2.0.CO;2>CrossRefGoogle Scholar
Marshall, G. J. (2009). On the annual and semi-annual cycles of precipitation across Antarctica. Int. J. Climatol., 29, 2298–2308.CrossRefGoogle Scholar
Marshall, J., et al. (2014). The ocean’s role in setting the mean position of the Inter-Tropical Convergence Zone. Clim. Dyn., 42, 1967–1979.CrossRefGoogle Scholar
Mather, K. B. and Miller, G. S. (1966). Wind drainage off the high plateau of eastern Antarctica. Nature, 209, 281–284.CrossRefGoogle Scholar
Mather, K. B. and Miller, G. S. (1967). The problem of the katabatic wind on the coast of Terre Adélie. Polar Rec., 13, 425–432.CrossRefGoogle Scholar
McAlpine, J. R., et al. (1983). Climate of Papua New Guinea. Commonwealth Scientific and Industrial Research Organisation – Australian National University Press, Canberra, 200pp.Google Scholar
McBride, J. (1998). Indonesia, Papua New Guinea, and Tropical Australia: The Southern Hemisphere Monsoon. Chapter 3. In Karoly, D. J. and Vincent, D. G. (eds.), Meteorology of the Southern Hemisphere. American Meteorological Society, Boston, MA, pp. 89–99.Google Scholar
McGregor, G. R. and Nieuwolt, S. (1982). Tropical Climatology, 2nd ed. John Wiley & Sons, Chichester, 339pp.Google Scholar
McGregor, S., et al. (2021). The effect of strong volcanic eruptions on ENSO. Chapter 12. In McPhaden, M. J., et al. (eds.), El Niño Southern Oscillation in a Changing Climate. Geophysical Monograph 253, American Geophysical Union and John Wiley & Sons, New York, pp. 267–287, 506pp.Google Scholar
McPhaden, M. J., et al. (eds.). (2021). El Niño Southern Oscillation in a Changing Climate. Geophysical Monograph 253, American Geophysical Union and John Wiley & Sons, New York, 506pp.Google Scholar
Meehl, G. A. (1991). A reexamination of the mechanism of the semi- annual cycle in the Southern Hemisphere. J. Clim., 4, 911–926.2.0.CO;2>CrossRefGoogle Scholar
Meehl, G. A., et al. (1998). A modulation of the mechanism of the semiannual oscillation in the Southern Hemisphere. Tellus, 50A, 442–450.Google Scholar
Meredith, M. P. and Hogg, A. M. (2006). Circumpolar response of Southern Ocean eddy activity to a change in the Southern Annular Mode. Geophys. Res. Lett., 33, L16608. https://doi.org/10.1029/2006GL026499.CrossRefGoogle Scholar
Miller, A. J., et al. (1994). The 1976–77 climate shift of the Pacific Ocean. Oceanography, 7, 1–6.CrossRefGoogle Scholar
Mo, K. and Higgins, R. (1998). The Pacific-South American modes and tropical convection during the southern hemisphere winter. Mon. Weather Rev., 126, 1581–1596.2.0.CO;2>CrossRefGoogle Scholar
Mo, K. and Paegle, J. (2001). The Pacific-South American modes and their downstream effects. Int. J. Climatol., 21, 1211–1229.CrossRefGoogle Scholar
Molnar, P., et al. (2010). Orographic controls on climate and paleoclimate of Asia: Thermal and mechanical roles for the Tibetan Plateau. Annu. Rev. Earth Planet. Sci., 38, 77–102.CrossRefGoogle Scholar
Morioka, Y., et al. (2013). How is the Indian Ocean Subtropical Dipole excited? Clim. Dyn., 41, 1955–1968. https://doi.org/10.1007/s00382-012-1584-9.CrossRefGoogle Scholar
Neelin, J. D. (2011). Climate Change and Climate Modelling. Cambridge University Press, Cambridge, 282pp.Google Scholar
Newman, M., et al. (2016). The Pacific Decadal Oscillation, revisited. J. Clim., 29(12), 4399–4427.CrossRefGoogle Scholar
Nicholls, N., McBride, J. L. and Ormerod, R. J. (1982). On predicting the onset of the Australian west season at Darwin. Mon. Wea. Rev., 110, 14–17.2.0.CO;2>CrossRefGoogle Scholar
Nicholson, S. E. (2016). The Turkana low-level jet: Mean climatology and association with regional aridity. Int. J. Climatol., 36, 2598–2614. https://doi.org/10.1002/joc.4515.CrossRefGoogle Scholar
Nicholson, S. E. (2017). Climate and climatic variability of rainfall over eastern Africa, Rev. Geophys., 55, 590–635. https://doi.org/10.1002/2016RG000544.CrossRefGoogle Scholar
Nicholson, S. E. (2018). The ITCZ and the seasonal cycle over Equatorial Africa. Bull. Am. Meteorol. Soc., 337–348. https://doi.org/10.1175/BAMS-D-16-0287.1.CrossRefGoogle Scholar
Norris, J. R. (2005). Multidecadal changes in near-global cloud cover and estimated cloud cover radiative forcing. J. Geophys. Res., 110, D08206. https://doi.org/10.1029/2004JD005600.Google Scholar
Okumura, Y. M. (2019). ENSO diversity from an atmospheric perspective. Curr. Clim. Change Rep., 5, 245–257. https://doi.org/10.1007/s40641-019-00138-7.CrossRefGoogle Scholar
Oliver, J. E. and Hidore, J. J. (2002). Climatology: An Atmospheric Science, 2nd ed. Pearson, New Jersey, 410pp.Google Scholar
Parish, T. R. and Bromwich, D. H. (1987). The surface windfield over the Antarctic ice sheets. Nature, 328, 51–54.CrossRefGoogle Scholar
Parish, T. R. and Cassano, J. J. (2003). The role of katabatic winds on the Antarctic surface wind regime. Mon. Weather Rev., 131, 317–333.2.0.CO;2>CrossRefGoogle Scholar
Parish, T. R. and Bromwich, D. H. (2007). Reexamination of the near-surface airflow over the Antarctic Continent and implications on atmospheric circulations at high southern latitudes. Mon. Weather Rev., 135, 1961–1973.CrossRefGoogle Scholar
Pepler, A., et al. (2019). A global climatology of surface anticyclones, their variability, associated drivers and long-term trends. Clim. Dyn., 52, 5397–5412.CrossRefGoogle Scholar
Philander, S. G. H., et al. (1996). Why the ITCZ is mostly north of the equator. J. Clim., 9, 2958–2972.2.0.CO;2>CrossRefGoogle Scholar
Power, S., et al. (2021). Decadal climate variability in the tropical Pacific: Characteristics, causes, predictability, and prospects. Science, 374, 48, eaay9165. https:doi.org/10.1126/science.aay9165.CrossRefGoogle ScholarPubMed
Ramage, C. S. (1971). Monsoon Meteorology. International Geophysical Series, Volume 15, Academic Press, London, 296pp.Google Scholar
Rintoul, S. R. (2018). The global influence of localized dynamics in the Southern Ocean. Nature, 558, 209–218. https://doi.org/10.1038/s41586-018-0182-3.CrossRefGoogle ScholarPubMed
Saji, H. H., et al. (1999). A dipole mode in the tropical Indian Ocean. Nature, 401, 360–363.CrossRefGoogle ScholarPubMed
Sallée, J. B., et al. (2008). An estimate of Lagrangian eddy statistics and diffusion in the mixed layer of the Southern Ocean. J. Marine Res., 66, 441–463.CrossRefGoogle Scholar
Sallée, J. B., et al. (2010). Zonally asymmetric response of the Southern Ocean mixed-layer depth to the Southern Annular Mode. Nat. Geosci., 3, 273–279.CrossRefGoogle Scholar
Schmidt, D. F. and Grise, K. M. (2019). Impacts of subtropical highs on summertime precipitation in North America. J. Geophys. Res. Atmospheres, 124(11), 188–204. https://doi.org/10.1029/2019JD031282.CrossRefGoogle Scholar
Schneider, T. and Sobel, A. H. (2007). The Global Circulation of the Atmosphere. Princeton University Press, New Jersey, 400pp.Google Scholar
Schneider, T., et al. (2014). Migrations and dynamics of the intertropical convergence zone. Nature, 513, 45–53. https://doi.org/10.1038/nature13636.CrossRefGoogle ScholarPubMed
Sen Gupta, A. and England, M. H. (2006). Coupled ocean–atmosphere–ice response to variations in the southern annular mode. J. Clim., 19, 4457–4486. https://doi.org/10.1175/JCLI3843.1.CrossRefGoogle Scholar
Senapati, B., et al. (2014). Global wave number-4 pattern in the southern subtropical sea surface temperature. Sci. Rep., 11, 142. https://doi.org/10.1038/s41598-020-80492-x.Google Scholar
Siedler, G., et al. (2013). Ocean Circulation and Climate: A 21st century Perspective. International Geophysics Series, Volume 103, Academic Press, London, 715pp.Google Scholar
Silva, V. B. S. and Kousky, V. E. (2012). The South American monsoon system: Climatology and variability. Chapter 5. In Wang, S.-Y. (ed), Modern Climatology. InTech, Croatia, pp. 123–152, 398pp.Google Scholar
Simmonds, I. and Jones, D. A. (1998). The mean structure and temporal variability of the semiannual oscillation in the southern extra- tropics. Int. J. Climatol., 18, 473–504.3.0.CO;2-0>CrossRefGoogle Scholar
Spensberger, C., et al. (2020). The connection between the Southern Annular Mode and a feature-based perspective on Southern Hemisphere midlatitude winter variability. J. Clim., 33, 115–129. https://doi.org/10.1175/JCLI-D-19-9224.1.CrossRefGoogle Scholar
Stearns, C. R., et al. (1993). Monthly mean climatic data for Antarctic automatic weather stations. In Bromwich, D. H. and Stearns, C. R. (eds.) Antarctic Meteorology and Climatology: Studies based on automatic weather stations, Antarctic Research Series 61, 1–22. American Geophysical Union, Washington, DC.Google Scholar
Sturman, A. and Tapper, N. (2006). The Weather and Climate of Australia and New Zealand, 2nd ed. Oxford University Press, Oxford, 541pp.Google Scholar
Sun, W., et al. (2022). Pacific multidecadal (50–70 year) variability instigated by volcanic forcing during the Little Ice Age (1250–1850). Clim. Dyn. https://doi.org/10.1007/s00382-021-06127-7.CrossRefGoogle Scholar
Suppiah, R. (1992). The Australian summer monsoon: A review. Prog. Physi. Geogr., 16, 283–318.Google Scholar
Thompson, D. W. J. and Wallace, J. M. (2000). Annular modes in the extratropical circulation. Part I: Month-to-month variability. J. Clim., 13, 1000–1016.Google Scholar
Timmermann, A., et al. (2018). El Niño–Southern Oscillation complexity. Nature, 559, 535–545. https://doi.org/10.1038/s41586-018-0252-6.CrossRefGoogle ScholarPubMed
Toggweiler, J. R. and Russell, J. (2008). Ocean circulation in a warming climate. Nature, 45. https://doi.org/10.1038/nature06590.Google Scholar
Toma, V. E. and Webster, P. J. (2010). Oscillations of the intertropical convergence zone and the genesis of easterly waves. Part I: Diagnostics and theory. Clim. Dyn., 34, 587–604. https://doi.org/10.1007/s00382-009-0584-x.Google Scholar
Trenberth, K. E. (1997). The definition of El Niño. Bull. Am. Meteorol. Soc., 78(12), 2271–2777.2.0.CO;2>CrossRefGoogle Scholar
Trenberth, K. E., et al. (2000). The global monsoon as seen through the divergent atmospheric circulation. J. Clim., 13, 3969–3993.2.0.CO;2>CrossRefGoogle Scholar
Turner, J. and Marshall, G. J. (2011). Climate Change in the Polar Regions. Cambridge University Press, Cambridge, 434pp.CrossRefGoogle Scholar
Tyson, P. D. and Preston-Whyte, R. A. (2000). The Weather and Climate of Southern Africa. Oxford University Press, Cape Town, 396pp.Google Scholar
Van den Broeke, M. R. (1998). The semiannual oscillation and Antarctic climate, part 1: Influence on near-surface temperatures (1957–1979). Ant. Sci., 10, 175–183.CrossRefGoogle Scholar
Van den Broeke, M. R. (2000). The semi-annual oscillation and Antarctic climate. Part 3: The role of near-surface wind speed and cloudiness. Int. J. Climatol., 20, 117–130.3.0.CO;2-B>CrossRefGoogle Scholar
Van den Broeke, M. R. and van Lipzig, N. (2003). Factors controlling the near- surface wind field in Antarctica. Mon. Weather Rev., 131, 733–743.2.0.CO;2>CrossRefGoogle Scholar
van Loon, H. (1967). The half-yearly oscillations in middle and high Southern latitudes and the coreless winter. J. Atmos. Sci., 24, 472–486.2.0.CO;2>CrossRefGoogle Scholar
van Loon, H. (1972). Temperature, pressure and wind in the Southern Hemisphere. In Newton, C. W. (ed), Meteorology of the Southern Hemisphere, Vol. 18, No. 35, Meteorological Monographs. American Meteorological Society, Boston, pp. 25–100.CrossRefGoogle Scholar
Van Loon, H. and Jenne, R. L. (1972). The zonal harmonic standing waves in the Southern Hemisphere. J. Geophys. Res., 77, 992–1003.Google Scholar
Van Loon, H. and Rogers, J. S. C. (1984a). The yearly wave in pressure and zonal geostrophic wind at sea level on the Southern Hemisphere and its interannual variability. Tellus, 36A, 348–354.Google Scholar
Van Loon, H. and Rogers, J. S. C. (1984b). Interannual variations in the half- yearly cycle of pressure gradients and zonal wind at sea level on the Southern Hemisphere. Tellus, 36A, 76–86.Google Scholar
Vignon, É., et al. (2020). Gravity wave excitation during the coastal transition of an extreme katabatic flow in Antarctica. J. Atmos. Sci., 77, 1295–1312.CrossRefGoogle Scholar
Visbeck, M. and Hall, A. (2004). Reply. J. Clim., 17, 2255–2258.2.0.CO;2>CrossRefGoogle Scholar
Vizy, E. K. and Cook, K. H. (2003). Connections between the summer east African and Indian rainfall regimes. J. Geophys. Res., 108(16), 4510. https://doi.org/10.1029/2003JD003452.Google Scholar
Walland, D. and Simmonds, I. (1999). Baroclinicity meridional temperature gradients and the Southern Semiannual Oscillation. J. Clim. 12, 3376–3382.2.0.CO;2>CrossRefGoogle Scholar
Wang, C., et al. (2017). El Niño and Southern Oscillation (ENSO). In Glynn, P. W., et al. (eds.), Coral Reefs of the Eastern Tropical Pacific, Coral Reefs of the World, 8, pp. 85–106. https://doi.org/10.1007/978-94-017-7499-4_4.Google Scholar
Wang, F. (2010). Subtropical dipole mode in the Southern Hemisphere: A global view. Geophys. Res. Lett., 37, 1–4.CrossRefGoogle Scholar
Webster, P. J. (1987). The elementary monsoon. In Fein, J. S. and Stephens, P. L. (eds.), Monsoons. John Wiley & Sons, New York, pp. 3–32.Google Scholar
Weisse, R. and von Storch, H. (2010). Marine Climate and Climate Change: Storms, Wind Waves and Storm Surges. Praxis Publishing Ltd, Chichester, 419pp.CrossRefGoogle Scholar
Wells, N. (1999). The Atmosphere and Ocean: A Physical Introduction, 2nd ed. John Wiley & Sons, Chichester, 394pp.Google Scholar
Wendler, G., et al. (1993). Katabatic winds in Adelie coast. In Bromwich, D. H. and Stearns, C. R. (eds.), Antarctic Meteorology and Climatology: Studies Based on Automatic Weather Stations. Antarctic Research Series 61, American Geophysical Union, Washington, DC, pp. 23–46.Google Scholar
Wheeler, M. C. and McBride, J. L. (2012). Australasian monsoon. In Lau, W. K. M. and Waliser, D. E. (eds.), Intraseasonal Variability in the Atmosphere–Ocean Climate System, 2nd ed. Springer-Verlag, Berlin Heidelberg, pp. 147–197.Google Scholar
Widlansky, M. J., et al. (2011). On the location and orientation of the South Pacific convergence zone. Clim. Dyn., 36, 561–578. https://doi.org/10.1007/s00382-010-0871-6.CrossRefGoogle Scholar
Wunsch, C. (1998). Work done by the wind on the oceanic general circulation. J. Phys. Oceanogr., 28, 2332–2340.2.0.CO;2>CrossRefGoogle Scholar
Xie, S.-P., et al. (2018). Eastern Pacific ITCZ dipole and ENSO diversity. J. Clim., 31, 4449–4462. https://doi.org/10.1175/JCLI-D-17-0905.1.CrossRefGoogle Scholar
Zebiak, S. E. and Cane, M. A. (1987). A model El Niño-Southern Oscillation. Mon. Weather Rev., 115, 2262–2278.2.0.CO;2>CrossRefGoogle Scholar
Zhang, H. and Moise, A. (2016). The Australian Summer Monsoon in current and future climate. Chapter 5. In Carvalho, L. M. V. de and Jones, C. (eds.), The Monsoons and Climate Change. Springer Climate, Switzerland, pp. 67–120. https://doi.org/10.1007/978-3-319-21650-8_5.Google Scholar
Arblaster, J. M. and Meehl, G. A. (2006). Contributions of external forcings to the Southern Annular Mode trends. J. Clim., 19, 2896–2905.CrossRefGoogle Scholar
Ashcroft, L. C., et al. (2009). Cold events over Southern Australia: Synoptic climatology and hemispheric structure. J. Clim., 22, 6679–6698.CrossRefGoogle Scholar
Ashok, K., et al. (2007). El Niño Modoki and its possible teleconnection. J. Geophys. Res. Oceans, 112, C11007. https://doi.org/10.1029/2006JC003798.CrossRefGoogle Scholar
Ashok, K., et al. (2009). ENSO Modoki impact on the Southern Hemisphere storm track activity during extended austral winter. Geophys. Res. Lett., 36(12), L12705. https://doi.org/10.1029/2009GL038847.CrossRefGoogle Scholar
Ashok, K. and Yamagata, T. (2009). The El Niño with a difference. Nature, 461, 481–484.CrossRefGoogle ScholarPubMed
Austin, J. F. (1980). The blocking of middle latitude westerly winds by planetary waves. Q. J. R. Meteor. Soc., 106, 327–350.CrossRefGoogle Scholar
Babian, S., et al. (2018). A new index for the wintertime southern hemispheric split jet. Atmos. Chem. Phys., 18, 6749–6760. https://doi.org/10.5194/acp-18-6749-2018.CrossRefGoogle Scholar
Baiman, R., et al. (2023). Synoptic drivers of atmospheric river induced precipitation near Dronning Maud Land, Antarctica. J. Geophys. Res. Atmos., 128, e2022JD037859. https://doi.org/10.1029/2022JD037859.CrossRefGoogle Scholar
Bals-Elsholz, T. M., et al. (2001). The wintertime Southern Hemisphere split jet: Structure, variability, and evolution. J. Clim., 14(21), 4191–4215.2.0.CO;2>CrossRefGoogle Scholar
Barry, R. G. and Carleton, A. M. (2001). Synoptic and Dynamic Climatology. Routledge, London, UK, 620pp.Google Scholar
Bergeron, T. (1928). Uber die dreidimensional veknupfende Wetteranalyse, Part 1: Prinzipelle Einfuhurung in das Problem der Luftmassen-und Frontenbildung. Geofys. Publ. (Oslo), 5, 1–111.Google Scholar
Berry, G., et al. (2011). A global climatology of atmospheric fronts. Geophys. Res. Lett., 38, L04809. https://doi.org/10.1029/2010GL046451.CrossRefGoogle Scholar
Bjerknes, J. and Solberg, H. (1922). Life cycles of cyclones and the polar front theory of atmospheric circulation. Geofys. Publ., 3, 1–18.Google Scholar
Black, A., et al. (2021). Australian northwest cloudbands and their relationship to atmospheric rivers and precipitation. Mon. Weather Rev., 149, 1125–1139.CrossRefGoogle Scholar
Blamey, R. C., et al. (2018). The influence of atmospheric rivers over the South Atlantic on winter rainfall in South Africa. J. Hydroclimatol., 19, 127–142. https://doi.org/10.1175/JHM-D-17-0111.1.Google Scholar
Blanchard-Wrigglesworth, E., et al. (2023). The largest ever recorded heatwave – Characteristics and attribution of the Antarctic heatwave of March 2022. Geophys. Res. Lett., 50, e2023GL104910. https://doi.org/10.1029/2023GL104910.CrossRefGoogle Scholar
Bozkurt, D., et al. (2018). Foehn event triggered by an atmospheric river underlies record-setting temperature along continental Antarctica. J. Geophys. Res. Atmospheres, 123, 3871–3892. https://doi.org/10.1002/2017JD027796.CrossRefGoogle Scholar
Bromwich, D. H. (2013). Central West Antarctica among the most rapidly warming regions on Earth. Nat. Geosci., 6, 139–145. https://doi.org/10.1038/NGEO1671.CrossRefGoogle Scholar
Brown, J. C., et al. (2009). An investigation of the links between ENSO flavors and rainfall processes in Southeastern Australia. Mon. Weather Rev., 137, 3786–3795.CrossRefGoogle Scholar
Browning, S. A. and Goodwin, I. D. (2013). Large-scale influences on the evolution of winter subtropical maritime cyclones affecting Australia’s east coast. Mon. Weather Rev., 141, 2416–2431. https://doi.org/10.1175/MWR-D-12-00312.1.CrossRefGoogle Scholar
Carleton, A. M. (1989). Antarctic sea ice relationships with indices of the atmospheric circulation of the southern hemisphere. Clim. Dyn., 2, 207–220.Google Scholar
Catto, J. L., et al. (2019). The future of midlatitude cyclones. Curr. Clim. Change Rep., 5, 407–420. https://doi.org/10.1007/s40641-019-00149-4.CrossRefGoogle Scholar
Cavicchia, L., et al. (2020). Future changes in the occurrence of hybrid cyclones: The added value of cyclone classification for the east Australian low-pressure systems. Geophys. Res. Lett., 47, e2019GL085751.CrossRefGoogle Scholar
Chand, S. S., et al. (2019). Review of tropical cyclones in the Australian region: Climatology, variability, predictability, and trends. WIREs Clim. Change, 10, e602. https://doi.org/10.1002/wcc.602.CrossRefGoogle Scholar
Chand, S. S., et al. (2021). Corrigendum. Review of tropical cyclones in the Australian region: Climatology, variability, predictability, and trends. WIREs Clim. Change, 12, e686. https://doi.org/10.1002/wcc.686.Google Scholar
Chang, E. K. M., et al. (2002). Storm-track dynamics. J. Clim., 15, 2163–2183. https://doi.org/10.1175/1520-0442(2002)015,02163:STD.2.0.CO;2.2.0.CO;2>CrossRefGoogle Scholar
Charabi, Y. and Al-Hatruski, S. (2010). Indian Ocean Tropical Cyclones and Climate Change. Springer, Dordrecht, 355pp.CrossRefGoogle Scholar
Chemke, R. (2021). The future poleward shift of Southern Hemisphere summer mid-latitude storm tracks stems from ocean coupling. Nat. Commun., 13, 1730. https://doi.org/10.1038/s41467-022-29392-4.Google Scholar
Chemke, R. and Ming, Y. (2020). Large atmospheric waves will get stronger, while small waves will get weaker by the end of the 21st century. Geophys. Res. Lett., 47, e2020GL090441.CrossRefGoogle Scholar
Chen, B., et al. (1996). Evolution of the tropospheric split jet over the South Pacific Ocean during the 1986–89 ENSO cycle, Mon. Weather Rev., 124, 1711–1731.2.0.CO;2>CrossRefGoogle Scholar
Chiang, J. C. H., et al. (2014). South Pacific Split Jet, ITCZ shifts, and atmospheric North–South linkages during abrupt climate changes of the last glacial period. Earth Planet. Sci. Lett., 406, 233–246.CrossRefGoogle Scholar
Chiang, J. C. H., et al. (2018). Contrasting impacts of the South Pacific Split Jet and the Southern Annular Mode modulation on Southern Ocean circulation and biogeochemistry. Paleoceanogr. Palaeoclimatol., 33(1), 2–20.Google Scholar
Coughlan, M. J. (1983). A comparative climatology of blocking action in the two hemispheres. Aust. Meteor. Mag., 31, 3–13.Google Scholar
Cullen, N. J., et al. (2019). The influence of weather systems in controlling mass balance in the Southern Alps of New Zealand. J. Geophys. Res. Atmos., 124, 4514–4529. https://doi.org/10.1029/2018JD030052.CrossRefGoogle Scholar
Da Rocha, R. P., et al. (2019). Subtropical cyclones over the oceanic basins: A review. Ann. N.Y. Acad. Sci., 1436, 138–156. https://doi.org/10.1111/nyas.13927.CrossRefGoogle ScholarPubMed
De Kock, W. M., et al. (2021). Large summer rainfall events and their importance in mitigating droughts over the South Western Cape, South Africa. J. Hydrometeorol., 22, 587–599.CrossRefGoogle Scholar
Dettinger, M. D. (2020). Effects of atmospheric rivers. Chapter 5. In Ralph, F. M., et al. (eds.), Atmospheric Rivers. Springer, Switzerland, pp. 141–178, 252pp.CrossRefGoogle Scholar
Diamond, H. J., et al. (2012). Development of an enhanced tropical cyclone tracks database for the Southwest Pacific from 1840–2010. Int. J. Climatol., 32, 2240–2250.CrossRefGoogle Scholar
Ding, Q., et al. (2011). Winter warming in West Antarctica caused by central tropical Pacific warming. Nat. Geosci., 4, 398–403.CrossRefGoogle Scholar
Ding, Q., et al. (2012). Influence of the tropics on the Southern Annular Mode. J. Clim., 25, 6330–6348.CrossRefGoogle Scholar
Dowdy, A. J. (2014). Long-term changes in Australian tropical cyclone numbers. Atmos. Sci. Lett., 15, 292–298.CrossRefGoogle Scholar
Dowdy, A. and Kuleshov, Y. (2012). An analysis of tropical cyclone occurrence in the Southern Hemisphere derived from a new satellite-era data set. Int. J. Remote Sens., 33, 7382–7397.CrossRefGoogle Scholar
Dowdy, A. J., et al. (2014). Fewer large waves projected for eastern Australia due to decreasing storminess Nat. Clim. Chang., 4, 283–6.CrossRefGoogle Scholar
Dowdy, A. J., et al. (2019). Review of Australian east coast low pressure systems and associated extremes. Clim. Dyn., 53, 4887–4910. https://doi.org/10.1007/s00382-019-04836-8.CrossRefGoogle Scholar
Eichler, T. P. and Gottschalck, J. (2013). A comparison of Southern Hemisphere cyclone track climatology and interannual variability in coarse-gridded reanalysis datasets. Adv. Meteorol., 2013, 1–16. https://doi.org/10.1155/2013/891260.Google Scholar
Espinoza, V., et al. (2018). Global analysis of climate change projection effects on atmospheric rivers. Geophys. Res. Lett., 45, 4299–4308. https://doi.org/10.1029/2017GL076968.CrossRefGoogle Scholar
Evans, J. L. and Braun, A. (2012). A climatology of subtropical cyclones in the South Atlantic. J. Clim., 25, 7328–7340.CrossRefGoogle Scholar
Fauchereau, N., et al. (2003). Sea-surface temperature co-variability in the Southern Atlantic and Indian oceans and its connections with the atmospheric circulation in the Southern Hemisphere. Int. J. Climatol., 23, 663–677. https://doi.org/10.1002/joc.905.CrossRefGoogle Scholar
Favre, A., et al. (2012). Relationships between cut-off lows and the semiannual and southern oscillations. Clim. Dyn., 38, 1473–1487. https://doi.org/10.1007/s00382-011-1030-4.CrossRefGoogle Scholar
Favre, A., et al. (2013). Cut-off Lows in the South Africa region and their contribution to precipitation. Clim. Dyn., 41, 2331–2351. https://doi.org/10.1007/s00382-012-1579-6.CrossRefGoogle Scholar
Fleming, Z. L., et al. (2012). Review: Untangling the influence of air-mass history in interpreting observed atmospheric composition. Atmos. Res., 104–105, 1–39.Google Scholar
Forgarasi, S. and Strome, M. (1978). Computer Program for Calculating Atmospheric Planetary Waves. Inland Waters Directorate, Environmental Canada, Canada, 8pp.Google Scholar
Fredericksen, J. S. (1982). A unified three-dimensional instability theory of the onset of blocking and cyclogenesis. J. Atmos. Sc., 39, 969–987.Google Scholar
Fuenzalida, H. A., et al. (2005). Climatology of cut-off lows in the Southern Hemisphere. J. Geophys. Res., 110, 1–10. https://doi.org/10.1029/2005JD005934.Google Scholar
Fyfe, J. C. (2003). Extratropical Southern Hemisphere cyclones: Harbingers of climate change? J. Clim., 16, 2802–2805.2.0.CO;2>CrossRefGoogle Scholar
Garreaud, R. D. (1999). Cold air incursions over Subtropical and Tropical South America: A numerical case study. Mon. Wea. Rev., 127, 2823–2853.2.0.CO;2>CrossRefGoogle Scholar
Garreaud, R. D. (2000). Cold air incursions over subtropical South America: Mean structure and dynamics. Mon. Wea. Rev., 128, 2544–2559.2.0.CO;2>CrossRefGoogle Scholar
Garreaud, R. D. (2001). Subtropical cold surges: Regional aspects and global distribution. Int. J. Climatol., 21, 1181–1197. https://doi.org/10.1002/joc.687.CrossRefGoogle Scholar
Garreaud, R. D., et al. (2009). Present-day South American climate. Palaeogeogr. Palaeoclimatol. Palaeoecol., 281, 180–195.CrossRefGoogle Scholar
Gehring, J., et al. (2022). Orographic flow influence on precipitation during an atmospheric river event at Davis, Antarctica. J. Geophys. Res. Atmospheres, 127, e2021JD035210. https://doi.org/10.1029/2021JD035210.CrossRefGoogle Scholar
Gimeno, L., et al. (2014). Atmospheric rivers: A mini-review. Front. Earth Sci., 2(2), 1–5. https://doi.org/10.3389/feart.2014.00002.CrossRefGoogle Scholar
Gimeno, L., et al. (2016). Major mechanisms of atmospheric moisture transport and their role in extreme precipitation events. Ann. Rev. Environ. Resour., 41, 117–141.CrossRefGoogle Scholar
Goebbert, K. H. and Leslie, L. M. (2010). Interannual variability of northwest Australian tropical cyclones. J. Clim., 23, 4538–4555.CrossRefGoogle Scholar
Goodwin, I. D. (1990). Snow accumulation and surface topography in the katabatic zone of eastern Wilkes Land, Antarctica. Ant. Sci., 2(3), 235–242.CrossRefGoogle Scholar
Goodwin, I. D., et al. (2014). A reconstruction of extratropical Indo-Pacific sea-level pressure patterns during the Medieval Climate Anomaly. Clim. Dyn., 43(5–6), 1197–1219. https://doi.org/10.1007/s00382-013-1899-1.CrossRefGoogle Scholar
Goodwin, I. D., et al. (2016). Tropical and extratropical-origin storm wave types and their influence on the East Australian longshore sand transport system under a changing climate. J. Geophys. Res. Oceans, 121, 4833–4853.CrossRefGoogle Scholar
Gorodetskaya, I. V., et al. (2014). The role of atmospheric rivers in anomalous snow accumulation in East Antarctic. Geophys. Res. Lett., 41, 6199–6206. https://doi.org/10.1002/2014GL060881.CrossRefGoogle Scholar
Gorodetskaya, I. V., et al. (2023). Record-high Antarctic Peninsula temperatures and surface melt in February 2022: A compound event with an intense atmospheric river. npj Climate Atmos. Sc., 6(202), 1–18. https://doi.org/10.1038/s41612-023-00529-6.Google Scholar
Gozzo, L. F., et al. (2014). Subtropical cyclones over the southwestern South Atlantic: Climatological aspects and case study. J. Clim., 27, 8543–8562.CrossRefGoogle Scholar
Gramcianinov, C. B., et al. (2019). The properties and genesis environments of South Atlantic cyclones. Clim. Dyn., 53, 4115–4140.CrossRefGoogle Scholar
Griffiths, M., et al. (1998). Observations of a cut-off low over southern Australia. Q.J.R. Meteorol. Soc., 124, 1109–1132.Google Scholar
Guishard, M. P., et al. (2009). Atlantic subtropical storms. Part II: Climatology. J. Clim., 22, 3574–3594.CrossRefGoogle Scholar
Harangozo, S. A. and Harrison, M. S. J. (1983). On the use of synoptic data in indicating the presence of cloud bands over southern Africa. S. Afr. J. Sci., 79(10), 413–414.Google Scholar
Harrison, M. S. J. (1984). A generalized classification of South African rain-bearing synoptic systems. J. Climatol., 4, 547–560, https://doi.org/10.1002/joc.3370040510.CrossRefGoogle Scholar
Hart, N. C. G., et al. (2010). Tropical–extratropical interactions over southern Africa: Three cases of heavy summer season rainfall. Mon. Wea. Rev., 138, 2608–2623. https://doi.org/10.1175/2010MWR3070.1.CrossRefGoogle Scholar
Hart, N. C. G., et al. (2013). Cloud bands over southern Africa: Seasonality, contribution to rainfall variability and modulation by the MJO. Clim. Dyn., 41, 1199–1212. https://doi.org/10.1007/s00382-012-1589-4.CrossRefGoogle Scholar
Hart, R. E. (2003). A cyclone phase space derived from thermal wind and thermal asymmetry. Mon. Weather Rev., 131, 585–616.2.0.CO;2>CrossRefGoogle Scholar
Hobbs, W. R. and Raphael, M. N. (2007). A representative time-series for the Southern Hemisphere zonal wave 1. Geophys. Res. Lett., 34, L05702. https://doi.org/10.1029/2006GL028740.CrossRefGoogle Scholar
Hobbs, W. R. and Raphael, M. N. (2010). Characterizing the zonally asymmetric component of the SH circulation. Clim. Dyn., 35, 859–873. https://doi.org/10.1007/s00382-009-0663-z.CrossRefGoogle Scholar
Hodges, K. I., et al. (2011). A comparison of extratropical cyclones in recent reanalyses ERA-Interim, NASA MERRA, NCEP CFSR, and JRA-25. J. Clim., 24, 4888–4906.CrossRefGoogle Scholar
Holland, G. J. (1993). 1993: Ready reckoner. Global Guide to Tropical Cyclone Forecasting WMO/TD-560, G. J. Holland, Ed., World Meteorological Organization, 9.1–9.26.Google Scholar
Holland, G., et al. (2018). Tropical cyclones and the global energy budget: Their role and implications. Abstract Southern Hemisphere Meteorological Conference, 2018.Google Scholar
Hopkins, L. C. and Holland, G. J. (1997). Australian heavy-rain days and associated east coast cyclones: 1958–92. J. Clim., 10, 621–634.2.0.CO;2>CrossRefGoogle Scholar
Hoskins, B. J. and Hodges, K. I. (2005). A new perspective on Southern Hemisphere storm tracks. J. Clim., 18, 4108–4129. https://doi.org/10.1175/JCLI3570.1.CrossRefGoogle Scholar
Howard, E. R., et al. (2019). Tropical lows in southern Africa: Tracks, rainfall contributions and the role of ENSO. J. Geophys Res. Atmos., 124(21), 11009–11032. https://doi.org/10.1029/2019JD030803.CrossRefGoogle Scholar
Inatsu, M. and Hoskins, B. J. (2006). The zonal asymmetry of the Southern Hemisphere winter storm track. J. Clim., 17, 4882–4892.Google Scholar
Irving, D., et al. (2010). Mesoscale cyclone activity over the ice-free Southern Ocean: 1999–2008. J. Clim., 23, 5404–5420.CrossRefGoogle Scholar
Irving, D. and Simmonds, I. (2015). A novel approach to diagnosing Southern Hemisphere planetary wave activity and its influence on regional climate variability. J. Clim., 28(23), 9041–9057.CrossRefGoogle Scholar
James, P. E. (1939). Air masses and fronts in South America. Geogr. Rev., 29(1) (Jan., 1939), 132–134.CrossRefGoogle Scholar
Jones, D. A. and Simmonds, I. (1993). A climatology of Southern Hemisphere extratropical cyclones. Clim. Dyn., 9, 131–145.CrossRefGoogle Scholar
Karoly, D. J. (1989). Southern Hemisphere circulation features associated with El Niño -Southern Oscillation events. J. Clim., 2, 1239–1252.2.0.CO;2>CrossRefGoogle Scholar
Karoly, D. J. (1990). The role of transient eddies in low frequency zonal variations of the Southern Hemisphere circulation. Tellus, 42A, 41–50.Google Scholar
Kidson, J. W. and Sinclair, M. R. (1995). The influence of persistent anomalies on Southern Hemisphere storm tracks. J. Climate, 8, 1938–1950.2.0.CO;2>CrossRefGoogle Scholar
Kiem, A. S., et al. (2016). Links between East Coast Lows and the spatial and temporal variability of rainfall along the eastern seaboard of Australia. J. South. Hemisph. Earth. Syst. Sci., 66(2), 162–176.CrossRefGoogle Scholar
Kiladis, G. N. and Diaz, H. F. (1989). Global climate anomalies associated with extremes in the Southern Oscillation. J. Clim., 2, 1069–1090.2.0.CO;2>CrossRefGoogle Scholar
Kiladis, G. N. and Mo, K. C. (1998). Interannual and intraseasonal variability in the Southern Hemisphere. In Karoly, D. J. and Vincent, D. G. (eds.), Meteorology of the Southern Hemisphere. American Meteorological Society, Boston, 410pp.Google Scholar
Kingston, D. G., et al. (2016). Floods in the Southern Alps of New Zealand: The importance of atmospheric rivers. Hydrol. Processes, 30, 5063–5070. https://doi.org/10.1002/hyp.10982.CrossRefGoogle Scholar
Klotzbach, P. J. and Landsea, C. W. (2015). Extremely intense hurricanes: Revisiting Webster et al. (2005) after 10 years. J. Clim., 28, 7621–7629. https://doi.org/10.1175/JCLI-D-15-0188.1.CrossRefGoogle Scholar
Knapp, K. R., et al. (2010). The international best track archive for climate stewardship (IBTrACS). Bull. Am. Meteorol. Soc., 91, 363–376.CrossRefGoogle Scholar
Knutson, T. R., et al. (2010). Tropical cyclones and climate change. Nat. Geosci., 3, 157–163. https://doi.org/10.1038/ngeo779.CrossRefGoogle Scholar
Köppen, W. (1931). Grundriss der Klimakunde. Walter de Gruyter, Berlin.CrossRefGoogle Scholar
Köppen, W. (1936). Das geographische System der Klimate. In Köppen, W. and Geiger, R. (eds.), Handbuch der Klimatologie. Gebrüder Borntraeger, Berlin, pp. 1−44.Google Scholar
Lamy, F., et al. (2010). Holocene changes in the position and intensity of the southern westerly wind belt. Nat. Geosc., 3, 695–699.CrossRefGoogle Scholar
Lamy, F., et al. (2019). Precession modulation of the South Pacific westerly wind belt over the past million years. Proc. Natl. Acad. Sci. USA, 116(47), 23455–23460. https://doi.org/10.1073/pnas.1905847116.CrossRefGoogle ScholarPubMed
Langhamer, L., et al. (2018). Lagrangian detection of moisture sources for the Southern Patagonia Icefield (1979–2017). Front. Earth Sci., 6, 219.CrossRefGoogle Scholar
Limpasuvan, V. and Hartmann, D. L. (1999). Eddies and the annular modes of climate variability, Geophys. Res. Lett., 26, 3133–3136. https://doi.org/10.1029/1999gl010478.CrossRefGoogle Scholar
Lin, I.-I., et al. (2021). ENSO and tropical cyclones. Chapter 17. In McPhaden, M. J., Santoso, A. and Cai, W. (eds.), El Niño Southern Oscillation in a Changing Climate. Geophysical Monograph 253. American Geophysical Union and John Wiley & Sons, New York, pp. 377–408, 506pp.Google Scholar
Lin, Z. (2019). The South Atlantic-South Indian ocean pattern: A zonally oriented teleconnection along the Southern Hemisphere westerly jet in austral summer. Atmosphere, 10, 259.CrossRefGoogle Scholar
Little, K., et al. (2019). The role of atmospheric rivers for extreme ablation and snowfall events in the Southern Alps of New Zealand. Geophys. Res. Lett., 46, 2761–2771. https://doi.org/10.1029/2018GL081669.CrossRefGoogle Scholar
Lupo, A. R., et al. (2019). Changes in global blocking character in recent decades. Atmosphere, 10, 92. https://doi.org/10.3390/atmos10020092.CrossRefGoogle Scholar
Maclennan, M. L., et al. (2022). Contribution of atmospheric rivers to Antarctic precipitation. Geophys. Res. Lett., 49, e2022GL100585. https://doi.org/10.1029/2022GL100585.CrossRefGoogle ScholarPubMed
Macron, C., et al. (2014). How do tropical temperature troughs form and develop over southern Africa? J. Clim., 27, 1633–1647. https://doi.org/10.1175/JCLI-D-13-00175.1.CrossRefGoogle Scholar
Manhique, A. J., et al. (2015). Extreme rainfall and floods in southern Africa in January 2013 and associated circulation patterns. Nat. Hazards, 77, 679–691. https://doi.org/10.1007/s11069-015-1616-y.CrossRefGoogle Scholar
Mansfield, A. W. and Glassey, S. D. (1957). Notes on weather analysis in the Falkland Islands Dependencies, Antarctica. Falkland Islands Dependencies Survey Science Reports, No. 16. London: Her Majesty’s Stationery Office.Google Scholar
McAlpine, J. R., et al. (1983). Climate of Papua New Guinea. Commonwealth Scientific and Industrial Research Organisation – Australian National University Press, Canberra, 200pp.Google Scholar
McGregor, G. R. and Nieuwolt, S. (1998). Tropical Climatology, 2nd ed. John Wiley & Sons, Chichester, 339pp.Google Scholar
Mendes, D., et al. (2010). Climatology of extratropical cyclones over the South American–southern oceans sector. Theor. Appl. Climatol., 100, 239–250. https://doi.org/10.1007/s00704-009-0161-6.CrossRefGoogle Scholar
Mo, K. C., et al. (1987). A GCM Study on the maintenance of the June 1982 blocking in the Southern Hemisphere. J. Atmos. Sci., 44, 1123–1142. https://doi.org/10.1175/1520-0469(1987)044<1123:AGSOTM>2.0.CO;2.2.0.CO;2>CrossRefGoogle Scholar
Munoz, C., et al. (2020). A midlatitude climatology and interannual variability of 200- and 500-hPa Cut-Off Lows. J. Clim., 33, 2201–2222. https://doi.org/10.1175/JCLI-D-19-0497.1.CrossRefGoogle Scholar
Murray, R. J. and Simmonds, I. (1991a). A numerical scheme for tracking cyclone centres from digital data. Part I: Development and operation of the scheme. Aust. Meteorol. Mag., 39, 155–166.Google Scholar
Murray, R. J. and Simmonds, I. (1991b). A numerical scheme for tracking cyclone centres from digital data. Part II: Application to January and July general circulation model simulations. Aust. Meteorol. Mag., 39, 167–180.Google Scholar
Neu, U., et al. (2013). IMILAST: A community effort to intercompare extratropical cyclone detection and tracking algorithms. Bull. Am. Meteorol. Soc., 94, 529–547. https://doi.org/10.1175/BAMS-D-11-00154.1.CrossRefGoogle Scholar
Newell, R. E., et al. (1992). Tropospheric rivers? – A pilot study. Geophys. Res. Lett., 12, 2401–2404.Google Scholar
Nicholson, S. E., et al. (2018). Rainfall over the African continent from the 19th through the 21st century. Glob. Planet. Change, 165, 114–127.CrossRefGoogle Scholar
Oliveira, F. N. M., et al. (2014). A new climatology for Southern Hemisphere blockings in the winter and the combined effect of ENSO and SAM phases. Int. J. Climatol., 34, 1676–1692.CrossRefGoogle Scholar
Oliver, J. E. (1970). A genetic approach to climate classification, Assoc. Am. Geogr. Ann., 60, 615–637.CrossRefGoogle Scholar
Otkin, J. A. and Martin, J. E. (2004). A synoptic climatology of the subtropical Kona storm. Mon. Weather Rev., 132, 1502–1517.2.0.CO;2>CrossRefGoogle Scholar
Papritz, L., et al. (2015). A climatology of cold air outbreaks and their impact on air–sea heat fluxes in the high-latitude South Pacific. J. Clim., 28, 342–364.CrossRefGoogle Scholar
Pepler, A. S., et al. (2016). The influence of local sea surface temperatures on Australian east coast cyclones. J. Geophys. Res., 121(22), 13352–13363.CrossRefGoogle Scholar
Pepler, A., et al. (2019). A global climatology of surface anticyclones, their variability, associated drivers and long- term trends. Clim. Dyn., 52(9–10), 5397–5412. https://doi.org/10.1007/s00382-018-4451-5.CrossRefGoogle Scholar
Pepler, A. and Dowdy, A. (2021). Fewer deep cyclones projected for the midlatitudes in a warming climate, but with more intense rainfall. Environ. Res. Lett., 16(5), 54044. https://doi.org/10.1088/1748-9326/abf528.CrossRefGoogle Scholar
Petoukhov, V., et al. (2013). Quasiresonant amplification of planetary waves and recent Northern Hemisphere weather extremes. Proc. Natl. Acad. Sci. USA, 110, 5336–5341. https://doi.org/10.1073/pnas.1222000110.CrossRefGoogle ScholarPubMed
Pezza, A. B. and Ambrizzi, T. (2005). Dynamical conditions and synoptic tracks associated with different types of cold surge over tropical South America. Int. J. Climatol., 25, 215–241. https://doi.org/10.1002/joc.1080.CrossRefGoogle Scholar
Pezza, A. B. and Simmonds, I. (2005). The first South Atlantic hurricane: Unprecedented blocking, low shear and climate change. Geophys. Res. Lett., 32(15), L15712.CrossRefGoogle Scholar
Pezza, A. B., et al. (2007). Southern Hemisphere cyclones and anticyclones: Recent trends and links with decadal variability in the Pacific Ocean. Int. J. Climatol., 27, 1403–1419.CrossRefGoogle Scholar
Pillay, M. T. and Fitchett, J. M. (2019). Tropical cyclone landfalls south of the Tropic of Capricorn, southwest Indian Ocean. Clim. Res., 79(1), 23–37.CrossRefGoogle Scholar
Pillay, M. T. and Fitchett, J. M. (2020). Southern Hemisphere tropical cyclones: A critical analysis of regional characteristics. Int. J. Climatol., 41(1), 1–16. https://doi.org/10.1002/joc.6613.Google Scholar
Pillay, M. T. and Fitchett, J. M. (2021). On the conditions of formation of Southern Hemisphere tropical cyclones. Weather Clim. Extremes, 34, 100376. https://doi.org/10.1016/j.wace.2021.100376.Google Scholar
Pinheiro, H. R., et al. (2017). A new perspective of the climatological features of upper-level cut-off lows in the Southern Hemisphere. Clim. Dyn., 48, 541–559. https://doi.org/10.1007/s00382-016-3093-8.CrossRefGoogle Scholar
Pinheiro, M. C., et al. (2019). Atmospheric blocking and intercomparison of objective detection methods: Flow field characteristics. Clim. Dyn., 53, 4189–4216. https://doi.org/10.1007/s00382-019-04782-5.CrossRefGoogle ScholarPubMed
Pittock, A. B. (1980). Patterns of climatic variation in Argentina and Chile – I. Precipitation, 1931–1960. Mon. Weather Rev., 108, 1347–1361.2.0.CO;2>CrossRefGoogle Scholar
Pook, M., and Gibson, T. (1999). Atmospheric blocking and storm tracks during SOP-1 of the FROST project. Aust. Met. Mag., 48, 51–60.Google Scholar
Pook, M. J., et al. (2006). The synoptic decomposition of cool-season rainfall in the southeastern Australian cropping region. J. Appl. Meteor. Climatol., 45, 1156–1170.CrossRefGoogle Scholar
Pook, M. J. (2009). The autumn break for cropping in south-east Australia: Trends, synoptic influences and impacts on wheat yield. Int. J. Climatol., 29, 2012–2026.CrossRefGoogle Scholar
Pook, M. J., et al. (2012). The synoptic climatology of cool-season rainfall in the Central Wheatbelt of Western Australia. Mon. Wea. Rev., 140, 28–43.CrossRefGoogle Scholar
Pook, M. J., et al. (2013). The seasonal cycle of blocking and associated physical mechanisms in the Australian region and relationship with rainfall. Mon. Weather Rev., 141(12), 4534–4553.CrossRefGoogle Scholar
Pook, M. J., et al. (2014). A comparative synoptic climatology of cool-season rainfall in major grain-growing regions of southern Australia. Theor. Appl. Climatol., 117(3–4), 521–533.CrossRefGoogle Scholar
Previdi, M. and Polvani, L. M. (2014). Climate system response to stratospheric ozone depletion and recovery. Q.J.R. Meteorol. Soc., 140(685), 2401–2419. https://doi.org/10.1002/qj.2330.CrossRefGoogle Scholar
Prince, H. D., et al. (2021). A climatology of atmospheric rivers in New Zealand. J. Clim., 34, 4383–4402. https://doi.org/10.1175/JCLI-D-20-0664.1.CrossRefGoogle Scholar
Qi, L., et al. (1999). Cut-off low pressure systems over southern Australia: Climatology and case study. Int. J. Climatol., 19, 1633–1649.3.0.CO;2-0>CrossRefGoogle Scholar
Ralph, F. M., et al. (2004). Satellite and CALJET aircraft observations of atmospheric rivers over the eastern North-Pacific Ocean during the El Niño winter of 1997/98. Mon. Weather Rev., 132, 1721–1745.2.0.CO;2>CrossRefGoogle Scholar
Ralph, F. M., et al. (eds.), (2020). Atmospheric Rivers. Springer, Switzerland, 252pp.CrossRefGoogle Scholar
Ramage, C. S. (1962). The subtropical cyclone. J. Geophys. Res., 67, 1401–1411.Google Scholar
Ramsay, H. A., et al. (2008). Interannual variability of tropical cyclones in the Australian region: Role of large-scale environment. J. Clim., 21, 1083–1103.CrossRefGoogle Scholar
Raphael, M. N. (2004). A zonal wave 3 index for the Southern Hemisphere. Geophys. Res. Lett., 31, L23212. https://doi.org/10.1029/2004GL020365.CrossRefGoogle Scholar
Reason, C. J. C. and Smart, S. (2015). Tropical south east Atlantic warm events and associated rainfall anomalies over southern Africa. Front. Environ. Sci., 3, 24. https://doi.org/10.3389/fenvs.2015.00024.CrossRefGoogle Scholar
Reboita, M. S. (2010). Climatological features of cutoff low systems in the Southern Hemisphere. J. Geophys. Res. Atmos., 115, D17104. https://doi.org/10.1029/2009JD013251.CrossRefGoogle Scholar
Reboita, M. S., et al. (2015). Trend and teleconnection patterns in the climatology of extratropical cyclones over the Southern Hemisphere. Clim. Dyn., 45, 1929–1944.CrossRefGoogle Scholar
Reboita, M. S., et al. (2019a). The South Atlantic Subtropical Anticyclone: Present and future climate. Front. Earth Sci., 7, 8. https://doi.org/10.3389/feart.2019.00008.CrossRefGoogle Scholar
Reboita, M. S., et al. (2019b). Key features and adverse weather of the named Subtropical cyclones over the Southwestern South Atlantic Ocean. Atmosphere, 10, 6. https://doi.org/10.3390/atmos10010006.Google Scholar
Reeder, M. J. and Smith, R. K. (1998). Mesoscale meteorology. Meteorology of the Southern Hemisphere, Meteor Monogr No. 49, American Meteorological Society, pp. 201–241.CrossRefGoogle Scholar
Reid, K. J., et al. (2019). The Australian Northwest Cloudband: Climatology, mechanisms and association with precipitation. J. Clim., 32(20), 6665–6684. https://doi.org/10.1175/JCLI-D-19-0031.1.CrossRefGoogle Scholar
Reid, K. J., et al. (2022). Tropical, subtropical, and extratropical atmospheric rivers in the Australian region. J. Clim., 35, 2697–2708. https://doi.org/10.1175/JCLI-D-21-0606.1.CrossRefGoogle Scholar
Renwick, J. A. (2005). Persistent positive anomalies in the Southern Hemisphere circulation. Mon. Wea. Rev., 133, 977–988. https://doi.org/10.1175/MWR2900.1.CrossRefGoogle Scholar
Reyers, M. and Shao, Y. (2019). Cutoff lows off the coast of the Atacama Desert under present day conditions and in the Last Glacial Maximum. Glob. Planet. Change, 181, 102983. https://doi.org/10.1016/j.gloplacha.2019.102983.CrossRefGoogle Scholar
Riehl, H. and Malkus, J. (1958). On the heat balance in the equatorial trough zone. Geophysica, 6, 503–538.Google Scholar
Risbey, J. S., et al. (2009). On the remote drivers of rainfall variability in Australia. Mon. Weather Rev., 137, 3233–3253.CrossRefGoogle Scholar
Rossby, C.-G., et al. (1939). Relation between variations in the intensity of the zonal circulation of the atmosphere and the displacements of the semi-permanent centres of action. J. Mar. Res., 2(1), 39–55.CrossRefGoogle Scholar
Rutz, J. J., et al. (2020). Global and regional perspectives. Chapter 4. In Ralph, F. M., et al. (eds.), Atmospheric Rivers. Springer, pp. 89–140, 252pp.Google Scholar
Saavedra, F., et al. (2020). Atmospheric rivers contribution to the snow accumulation over the Southern Andes (26.5° S–37.5° S). Front. Earth Sci., 8, 261. https://doi.org/10.3389/feart.2020.00261.CrossRefGoogle Scholar
Salby, M. L. (1984). Survey of planetary-scale travelling waves: The state of theory and observations. Rev. Geophys. Space Phys., 22, 209–236.CrossRefGoogle Scholar
Schneider, C., et al. (2003). Weather observations across the southern Andes at 53 S. Physical Geography, 24(2), 97–119.CrossRefGoogle Scholar
Schwartz, M. D. and Corcoran, W. T. (2005). Airmass climatology. In Oliver, J. E. (eds.), Encyclopedia of World Climatology. Encyclopedia of Earth Sciences Series. Springer, Dordrecht. https://doi.org/10.1007/1-4020-3266-8_6.Google Scholar
Screen, J. A. and Simmonds, I. (2014). Amplified mid-latitude planetary waves favour particular regional weather extremes. Nat. Clim. Change, 4, 704–709.CrossRefGoogle Scholar
Senapati, B., et al. (2021). Global wave number-4 pattern in the southern subtropical sea surface temperature. Sci. Rep., 11, 142.CrossRefGoogle ScholarPubMed
Shapiro, L. J. and Goldenberg, S. B. (1998). Atlantic sea surface temperatures and tropical cyclone formation. J. Clim., 11, 578–590. https://doi.org/10.1175/1520-0442(1998)011!0578:ASSTATO2.0.CO;2.2.0.CO;2>CrossRefGoogle Scholar
Shutts, G. J. (1983). The propagation of eddies in diffluent jetstreams: Eddy vorticity forcing of ‘blocking’ flow fields. Q. J. R. Meteorol. Soc., 109, 737–761.Google Scholar
Simmonds, I. R. and Murray, J. (1999). Southern extratropical cyclone behavior in ECMWF analyses during the FROST special observing periods. Weather Forecast, 14, 878–891.2.0.CO;2>CrossRefGoogle Scholar
Simmonds, I., et al. (1999). A refinement of cyclone tracking methods with data from FROST. Aust. Meteor. Mag., 28, 617–622.Google Scholar
Simmonds, I. R. and Keay, K. (2000a). Mean Southern Hemisphere extratropical cyclone behavior in the 40-year NCEP, NCAR reanalysis, J. Clim., 13(5), 873–885.2.0.CO;2>CrossRefGoogle Scholar
Simmonds, I. and Keay, K. (2000b). Variability of Southern Hemisphere extratropical cyclone behaviour, 1958–97. J. Clim., 13, 550–561.2.0.CO;2>CrossRefGoogle Scholar
Sinclair, M. R. (1994). An objective cyclone climatology for the Southern Hemisphere. Mon. Weather Rev., 122, 2239–2256.2.0.CO;2>CrossRefGoogle Scholar
Sinclair, M. R. (1996). A climatology of anticyclones and blocking for the Southern Hemisphere. Mon. Weather Review, 124, 245–263.2.0.CO;2>CrossRefGoogle Scholar
Sinclair, M. R. (1997). Objective identification of cyclones and their circulation intensity and climatology. Weather Forecast, 12, 595–612.2.0.CO;2>CrossRefGoogle Scholar
Sinclair, M. R. (2002). Extratropical transition of southwest Pacific tropical cyclones. Part I: Climatology and mean structure changes. Mon. Weather Rev., 130, 590–609.2.0.CO;2>CrossRefGoogle Scholar
Sinclair, M. R. (2004). Extratropical transition of Southwest Pacific tropical cyclones. Part II: Midlatitude circulation characteristics. Mon. Weather Rev., 132, 2145–2168.2.0.CO;2>CrossRefGoogle Scholar
Singleton, A. T. and Reason, C. J. C. (2006). A numerical model study of an intense cutoff low pressure system over South Africa. Mon. Weather Rev., 135, 1128–1150.Google Scholar
Singleton, A. T. and Reason, C. J. C. (2007). Variability in the characteristics of cut-off low pressure systems over subtropical Southern Africa. Int. J. Climatol., 27, 295–310.CrossRefGoogle Scholar
Sodermann, H., et al. (2020). Structure, process and mechanism. Chapter 2. In Ralph, F. M., et al. (eds.), Atmospheric Rivers. Springer, Switzerland, pp. 15–44, 252pp.Google Scholar
Solman, S. A. and Menéndez, C. G. (2002). ENSO-related variability of the Southern Hemisphere winter storm track over the eastern Pacific–Atlantic sector. J. Atmos. Sci., 59, 2128–2140.2.0.CO;2>CrossRefGoogle Scholar
Speer, M. S., Leslie, L. M., and Hartigan, J. (2022). Jet stream changes over Southeast Australia during the early cool season in response to accelerated Global Warming. Climate, 10, 84. https://doi.org/10.3390/cli10060084.CrossRefGoogle Scholar
Spensberger, C., et al. (2020). The connection between the Southern Annular Mode and a feature-based perspective on Southern Hemisphere midlatitude winter variability. J. Clim., 33, 115–129. https://doi.org/10.1175/JCLI-D-19-9224.1.CrossRefGoogle Scholar
Sturman, A. and Tapper, N. (2006). The Weather and Climate of Australia and New Zealand, 2nd ed. Oxford University Press, Oxford, 541pp.å.Google Scholar
Taljaard, J. J. (1969). Airmasses of the Southern Hemisphere, Notos, 79–104.Google Scholar
Taljaard, J. J. (1972). Synoptic Meteorology of the Southern Hemisphere, 13 ed. Meteorol. Monogr., Boston.CrossRefGoogle Scholar
Taljaard, J. J. (1985). Cut-off lows in the South African region. South African Weather Bureau Technical paper 14.Google Scholar
Tamarin-Brodsky, T. and Kaspi, Y. (2017a). Enhanced poleward propagation of storms under climate change. Nat. Geosci., 10, 908–913. https://doi.org/10.1038/s41561-017-0001-8.CrossRefGoogle Scholar
Tamarin-Brodsky, T. and Kaspi, Y. (2017b). Mechanisms controlling the downstream poleward deflection of midlatitude storm tracks. J. Atmos. Sci., 74, 553–572.Google Scholar
Terry, J. P. (2007). Tropical Cyclones: Climatology and Impacts in the South Pacific. Springer, New York, 219pp.Google Scholar
Thomas, C. M. and Schultz, D. M. (2019). Global climatologies of fronts, airmass boundaries, and airstream boundaries: Why the definition of ‘front’ matters. Mon. Weather Rev., 147, 691–717.CrossRefGoogle Scholar
Thompson, D. W. J. and Wallace, J. M. (2000). Annular modes in the extratropical circulation. Part I: Month-to-month variability. J. Clim., 13, 1000–1016.Google Scholar
Tibaldi, S., et al. (1994). Northern and Southern Hemisphere seasonal variability of blocking frequency and predictability. Mon. Weather Rev., 122, 1971–2003.2.0.CO;2>CrossRefGoogle Scholar
Timbal, B. and Drosdowsky, W. (2012). The relationship between the decline of Southern Australian rainfall and the strengthening of the subtropical ridge. Int. J. Climatol., 33(4), 1021–1034. https://doi.org/10.1002/joc.3492.Google Scholar
Toggweiler, J. R. (2009). Climate change. Shifting westerlies. Science, 323(5920), 1434–1435. https://doi.org/10.1126/science.1169823.CrossRefGoogle ScholarPubMed
Toggweiler, J. R. and Russell, J. (2008). Ocean circulation in a warming climate. Nature, 451(7176), 286–288. https://doi.org/10.1038/nature06590.CrossRefGoogle Scholar
Trenberth, K. E. (1979). Interannual variability of the 500 mb zonal mean flow in the Southern Hemisphere. Mon. Weather Rev., 107, 523–534.2.0.CO;2>CrossRefGoogle Scholar
Trenberth, K. E. (1980a). Atmospheric quasi-biennial oscillations. Mon. Weather Rev., 108, 1370–1377.2.0.CO;2>CrossRefGoogle Scholar
Trenberth, K. E. (1980b). Planetary waves at 500 mb in the Southern Hemisphere. Mon. Weather Rev., 108, 1378–1389.2.0.CO;2>CrossRefGoogle Scholar
Trenberth, K. E. (1985). Persistence of daily geopotential heights over the Southern Hemisphere. Mon. Weather Rev., 113(1), 38–53.Google Scholar
Trenberth, K. E. and Mo, K. C. (1985). Blocking in the Southern Hemisphere. Mon. Weather Rev., 113, 3–21. https://doi.org/10.1175/1520-0493(1985)113,0003:BITSH.2.0.CO;2.2.0.CO;2>CrossRefGoogle Scholar
Tung, K. K. and Lindzen, R. S. (1979). A theory of stationary long waves. Part I: A simple theory of blocking. Mon. Weather Rev., 107(6), 714–734.Google Scholar
Turner, J., et al. (2016). Variability and trends in the Southern Hemisphere high latitude, quasi-stationary planetary waves. Int. J. Climatol., 37, 2325–2336. https://doi.org/10.1002/joc.4848.Google Scholar
Valenzuela, R. A. and Garreaud, R. D. (2019). Extreme daily rainfall in central- southern Chile and its relationship with low-level horizontal water vapor fluxes. J. Hydrometeorol., 20, 1829–1850. https://doi.org/10.1175/JHM-D-19-0036.1.CrossRefGoogle Scholar
van Loon, H. (1956). Blocking action in the Southern Hemisphere. Notos, 5, 171–177.Google Scholar
van Loon, H. (1972). Temperature, pressure and wind in the Southern Hemisphere. In Newton, C. W. (ed), Meteorology of the Southern Hemisphere, Vol. 18, No. 35, Meteorological Monographs. American Meteorological Society, Boston, pp. 25–100.CrossRefGoogle Scholar
van Loon, H. (1979). The association between latitudinal temperature gradient and eddy transport. Part I: Transport of sensible heat in winter. Mon. Weather Rev., 107, 525–534.2.0.CO;2>CrossRefGoogle Scholar
van Loon, H. and Jenne, R. L. (1972). The zonal harmonic standing waves in the Southern Hemisphere. J. Geophys. Res., 77, 992–1003.Google Scholar
van Loon, H. and Rogers, J. C. (1984). Interannual variations in the half- yearly cycle of pressure gradients and zonal wind at sea level on the Southern Hemisphere. Tellus, 36A, 76–86.Google Scholar
Viale, M. and Nuñez, M. N. (2011). Climatology of winter orographic precipitation over the subtropical central Andes and associated synoptic and regional characteristics. J. Hydrometeor., 12, 481–507. https://doi.org/10.1175/2010JHM1284.1.CrossRefGoogle Scholar
Viale, M., et al. (2013). Upstream orographic enhancement of a narrow cold-frontal rainband approaching the Andes. Mon. Weather Rev., 141(5), 1708–1730.CrossRefGoogle Scholar
Viale, M., et al. (2018). Impacts of atmospheric rivers on precipitation in Southern South America. J. Hydrometeorol., 19, 1671–1686. https://doi.org/10.1175/JHM-D-18-0006.1.CrossRefGoogle Scholar
Vincent, D. G. (1994). The South Pacific convergence zone (SPCZ): A review. Mon. Weather Rev., 122, 1949–1970. https://doi.org/10.1175/1520-0493.2.0.CO;2>CrossRefGoogle Scholar
Vuille, M. and Ammann, C. (1997). Regional snowfall patterns in the high, arid Andes. Clim. Change, 36, 413–423. https://doi.org/10.1023/A:1005330802974.CrossRefGoogle Scholar
Waliser, D. and Guan, B. (2017). Extreme winds and precipitation during landfall of atmospheric rivers. Nat Geosci., 10, 179–183.CrossRefGoogle Scholar
Webster, P. J. (1982). Seasonality in the local and remote atmospheric response to sea surface temperature anomalies. J. Atmos. Sci., 39, 41–52.2.0.CO;2>CrossRefGoogle Scholar
Webster, P. J., et al. (2005). Changes in tropical cyclone number, duration, and intensity in a warming environment. Science, 309, 1844–1846. https://doi.org/10.1126/science.1116448.CrossRefGoogle Scholar
Wendland, W. M. and McDonald, N. S. (1986). Southern Hemisphere airstream climatology. Mon. Weather Rev., 114, 88–94.2.0.CO;2>CrossRefGoogle Scholar
Wiedenmann, J. M., et al. (2002). The climatology of blocking anticyclones for the Northern and Southern Hemisphere block intensity as a diagnostic. J. Clim., 15, 3459–3473.2.0.CO;2>CrossRefGoogle Scholar
Wille, J. D., et al. (2021). Antarctic atmospheric river climatology and precipitation impacts. J. Geophys. Res. Atmos., 126, e2020JD033788. https://doi.org/10.1029/2020JD03378.CrossRefGoogle Scholar
Wille, J. D., et al. (2024a). The extraordinary March 2022 East Antarctica ‘heat’ wave. Part I: Observations and meteorological drivers. J. Clim., 37, 757–778. https://doi.org/10.1175/JCLI-D-23-0175.1.Google Scholar
Wille, J. D., et al. (2024b). The extraordinary March 2022 East Antarctica ‘heat’ wave. Part II: Impacts on the Antarctic ice sheet. J. Clim., 37, 779–799. https://doi.org/10.1175/JCLI-D-23-0176.1.Google Scholar
Woollings, T., et al. (2018). Blocking and its response to climate change. Curr. Clim. Change Rep., 4, 287–300. https://doi.org/10.1007/s40641-018-0108-z.CrossRefGoogle ScholarPubMed
Wright, A. D. (1974). Blocking action in the Australian Region. Department of Science Bureau of Meteorology (Technical Report 10), 29pp.Google Scholar
Wright, W. J. (1997). Tropical-extratropical cloudbands and Australian rainfall: I climatology. Int. J. Climatol., 17, 807–829.3.0.CO;2-J>CrossRefGoogle Scholar
Xia, L., et al. (2016). A study of quasi-millennial extratropical winter cyclone activity over the Southern Hemisphere. Clim. Dyn., 47, 2121–2138. https://doi.org/10.1007/s00382-015-2954-x.CrossRefGoogle Scholar
Yanase, W., et al. (2014). Parameter spaces of environmental fields responsible for cyclone development from tropics to extratropics. J. Clim., 27, 652–671.CrossRefGoogle Scholar
Yang, X. and Chang, E. K. M. (2006). Variability of the Southern Hemisphere winter split flow – A case of two-way reinforcement between mean flow and eddy anomalies, J. Atmos. Sci., 63, 634–650. https://doi.org/10.1175/JAS3643.1.CrossRefGoogle Scholar
Yasunari, T. (1977). Stationary waves in the Southern Hemisphere mi-latitude zone revealed from average brightness charts. J. Meteorol. Soc. Japan, Ser. II, 55(3), 274–285.Google Scholar
Zhu, Y. and Newell, R. E. (1994). Atmospheric rivers and bombs. Geophys. Res. Lett., 21(8), 1999–2002.CrossRefGoogle Scholar
Zhu, Y. and Newell, R. E. (1998). A proposed algorithm for moisture fluxes from atmospheric rivers. Mon. Weather Rev., 126, 725–735.2.0.CO;2>CrossRefGoogle Scholar
Aguirre, C., et al. (2019). Role of synoptic activity on projected changes in upwelling- favourable winds at the ocean’s eastern boundaries. Npj Clim. Atmos. Sci., 2, 44. https://doi.org/10.1038/s41612-019-0101-9.CrossRefGoogle Scholar
Allan, R., et al. (2011). The international Atmospheric Circulation Reconstructions over the Earth (ACRE) initiative. Bull. Am. Meteorol. Soc., 92, 1421–1425. https://doi.org/10.1175/2011BAMS3218.1.CrossRefGoogle Scholar
Allan, R., et al. (2016). Toward integrated historical climate research: The example of Atmospheric Circulation Reconstructions over the Earth. Wiley Interdiscip. Rev.: Clim. Change, 7, 164–174.Google Scholar
Arblaster, J. M. and Meehl, G. A., (2006). Contributions of external forcings to the Southern Annular Mode trends. J. Clim., 19, 2896–2905.CrossRefGoogle Scholar
Barber, N. F. and Ursell, F. (1948). The generation and propagation of ocean waves and swell; I. Wave periods and velocities. Phil. Trans. Roy. Soc. A, 240, 527–560.Google Scholar
Bárdossy, A, et al. (2015). Circulation patterns identified by spatial rainfall and ocean wave fields in Southern Africa. Front. Environ. Sci., 3, 31. https://doi.org/10.3389/fenvs.2015.00031.CrossRefGoogle Scholar
Barrett, H. G., et al. (2018). Reconstructing El Niño Southern Oscillation using data from ships’ logbooks, 1815–1854. Part I: Methodology and evaluation. Clim. Dyn., 50, 845–862. https://doi.org/10.1007/s00382-017-3644-7.Google Scholar
Beal, L. M., et al. (2011). On the role of the Agulhas system in ocean circulation and climate. Nature, 472, 429–436.CrossRefGoogle ScholarPubMed
Beal, L. M. and Elipot, S. (2016). Broadening not strengthening of the Agulhas Current since the early 1990s. Nature, 540, 570–573. https://doi.org/10.1038/nature19853.CrossRefGoogle Scholar
Beggs, P. J., et al. (2004). Identification of von Karman vortices in the surface winds of Heard Island. Bound.-Layer Meteorol., 113, 287–297.CrossRefGoogle Scholar
Bromirski, P. D., et al. (2011). Dynamical suppression of sea level rise along the Pacific coast of North America: Indications for imminent acceleration. J. Geophys. Res., 116, C07005. https://doi.org/10.1029/2010JC006759.Google Scholar
Bromirski, P. D., et al. (2017). Tsunami and infragravity waves impacting Antarctic ice shelves. J. Geophys. Res. Oceans, 122, 5786–5801. https://doi.org/10.1002/2017JC012913.CrossRefGoogle Scholar
Bromwich, D. H., et al. (2011). Climatological aspects of cyclogenesis near Adélie Land Antarctica. Tellus A: Dyn. Meteorol. Oceanogr., 63, 5, 921–938. https://doi.org/10.1111/j.1600-0870.2011.00537.x.CrossRefGoogle Scholar
Bronnimann, S., et al. (2018). A roadmap to climate data services. Geosci. Data J., 5, 28–39. https://doi.org/10.1002/gdj3.56.CrossRefGoogle Scholar
Bronselaer, B., et al. (2018). Change in future climate due to Antarctic meltwater. Nature, 564(7734), 53–58. https://doi.org/10.1038/s41586-018-0712-z.CrossRefGoogle ScholarPubMed
Brunet, M. and Jones, P. (2011). Data rescue initiatives: Bringing historical climate data into the 21st century. Clim. Res., 47, 29–40. https://doi.org/10.3354/cr00960.CrossRefGoogle Scholar
Burrows, C. J. (1976). Icebergs in the Southern Ocean. New Zealand Geographer, 32, 127–138.CrossRefGoogle Scholar
Caldeira, R. M. A., et al. (2002). Sea-surface signatures of the island mass effect phenomena around Madeira Island, Northeast Atlantic. Remote Sensing Environ., 80, 336–360.CrossRefGoogle Scholar
Caldeira, R. M. A., et al. (2005). Island wakes in the Southern California Bight. J. Geophys. Res., 110, C11012, https://doi.org/10.1029/2004JC002675.Google Scholar
Caldeira, R. M. A. and Sangra, P. (2012). Complex geophysical wake flows Madeira Archipelago case study. Ocean Dyn., 62, 683–700. https://doi.org/10.1007/s10236-012-0528-6.CrossRefGoogle Scholar
Cardone, V. J., et al. (1990). On trends in historical marine wind data. J. Clim., 3, 113–127.2.0.CO;2>CrossRefGoogle Scholar
Cazenave, A., et al. (2018). Global sea-level budget 1993-present. Earth Syst. Sci. Data, 10, 1551–1590 https://doi.org/10.5194/essd-10-1551-2018.Google Scholar
Cetina-Heredia, P., et al. (2014). Long-term trends in the East Australian Current separation latitude and eddy driven transport. J. Geophys. Res. Oceans, 119. https://doi.org/10.1002/2014JC010071.CrossRefGoogle Scholar
Chafik, L., et al. (2018). North Atlantic ocean circulation and decadal sea level change during the altimetry era. Sci. Rep., 9, 1041. https://doi.org/10.1038/s41598-018-37603-6.Google Scholar
Chamorro, A. (2018). Coastal winds dynamics in the Peruvian upwelling system under warming conditions: Impact of El Niño and regional climate change. Sorbonne Université École doctorale 129: Sciences de l’Environnement, Laboratoire d’Océanographie et du Climat: Expérimentations et Approches Numériques, 146pp.Google Scholar
Chappell, P. R. and Lorrey, A. M. (2014). Identifying New Zealand, Southeast Australia, and Southwest Pacific historical weather data sources using Ian Nicholson’s Log of Logs. Geosci. Data J., 1(1), 49–60.CrossRefGoogle Scholar
Church, J. A., et al. (1991). A model of sea level rise caused by ocean thermal expansion. J. Clim., 4, 438–456.2.0.CO;2>CrossRefGoogle Scholar
Church, J. A., et al. (2005). Significant decadal-scale impact of volcanic eruptions on sea level and ocean heart content. Nature, 438, 74–77. https://doi.org/10.1038/nature04237.CrossRefGoogle Scholar
Church, J. A. and White, N. J. (2011). Sea-level rise from the late 19th to the early 21st century. Surv. Geophys., 32, 4–5, 585–602. https://doi.org/10.1007/s10712-011-9119-1.CrossRefGoogle Scholar
Church, J. A., et al. (2013). Sea-Level Change. Cambridge University Press, Cambridge.Google Scholar
Clark, J. A. and Lingle, C. S. (1979). Predicted relative sea-level changes (18,000 years B.P. to present) caused by late-glacial retreat of the Antarctic ice sheet. Quat. Res., 11, 279–298.CrossRefGoogle Scholar
Clark, P. U., et al. (2020). Oceanic forcing of penultimate deglacial and last interglacial sea-level rise. Nature, 577, 660–664.CrossRefGoogle ScholarPubMed
Collard, F., et al. (2009). Monitoring and analysis of ocean swell fields from space: New methods for routine observations. J. Geophys. Res., 114, C07023. https://doi.org/10.1029/2008JC005215.Google Scholar
Cresswell, G. R. (2000). Currents of the continental shelf and upper slope of Tasmania. Pap. Proc. R. Soc. Tasmania, 133, 21e30.Google Scholar
de Mesquita, A. R. (2003). Sea-level variations along the Brazilian Coast: A short review. J. Coast. Res., Special Issue No. 35., 21–31.Google Scholar
Dendy, S., et al. (2017). Sensitivity of last interglacial sea-level high stands to ice sheet configuration during Marine Isotope Stage 6. Quat. Sci. Rev., 171, 234–244.CrossRefGoogle Scholar
Deutsche Seewarte. (1885). Segelhandbiicher fur den Atlantischen Ozean. [Sailing Handbooks for the Atlantic Ocean]. L. Friederichsen and Co., approx. 100pp.,Google Scholar
Deutsche Seewarte. (1892). Segelhandbiicher fur den Indischen Ozean [Sailing Handbooks for the Indian Ocean]. L. Friederichsen and Co., approx. 100pp.Google Scholar
Deutsche Seewarte. (1897). Segelhandbiicher fur den Stillen Ozean [Sailing Handbooks for the Pacific Ocean]. L. Friederichsen and Co., approx. 100pp.Google Scholar
Downes, C. R. (1977). History of the British ocean weather ships. The Marine Observer, 48, 179–186.Google Scholar
Durack, P. J. and Wijffels, S. E. (2010). Fifty-year trends in global ocean salinities and their relationship to broad-scale warming. J. Clim., 23, 16, 4342–4362. https://doi.org/10.1175/2010jcli3377.1.CrossRefGoogle Scholar
Durrant, T. and Greenslade, D. (2011). Evaluation and implementation of AUSWAVE. CAWCR Technical Report No. 041. www.cawcr.gov.au, 52pp.Google Scholar
Durrant, T., et al. (2014). A Global Wave Hindcast focussed on the Central and South Pacific. CAWCR Technical Report No. 070. www.cawcr.gov.au, 45pp.Google Scholar
England, M. H., et al. (2014). Recent intensification of wind-driven circulation in the Pacific and the ongoing warming hiatus. Nat. Clim. Change. https://doi.org/10.1038/NCLIMATE2106.CrossRefGoogle Scholar
Fasullo, J. T., et al. (2016). Is the detection of accelerated sea level rise imminent? Sci. Rep., 6, 31245.Google Scholar
Feng, M., et al. (2003). Annual and interannual variations of the Leeuwin Current at 32°S. J. Geophys. Res., 108, C11, 3355. https://doi.org/10.1029/2002JC001763.Google Scholar
Feng, M. and Meyers, G. (2004). Multidecadal variations of Fremantle sea level: Footprint of climate variability in the tropical Pacific. Geophys. Res. Lett., 31, L16302. https://doi.org/10.1029/2004GL019947.CrossRefGoogle Scholar
Feng, M., et al. (2013). La Niña forces unprecedented Leeuwin Current warming in 2011. Sci. Rep., 2, 1277. https://doi.org/10.1038/srep01277.Google Scholar
Fox-Kemper, B., et al. (2021). Ocean, Cryosphere and Sea Level Change. In Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change [Masson-Delmotte, V., et al. (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 1211–1362. https://doi.org/10.1017/9781009157896.011.CrossRefGoogle Scholar
Francis, D., et al. (2019). Polar cyclones at the origin of the reoccurrence of the Maud Rise Polynya in austral winter 2017. J. Geophys. Res.: Atmos., 124, 5251–5267. https://doi.org/10.1029/2019JD030618.CrossRefGoogle Scholar
Frankcombe, L. M., et al. (2013). Sea level changes forced by Southern Ocean winds. Geophys. Res. Lett., 40(21), 5710–5715.CrossRefGoogle Scholar
Freeman, E., et al. (2017). ICOADS Release 3.0: A major update to the historical marine climate record. Int. J. Climatol., 37, 2211–2232.CrossRefGoogle Scholar
Gallego, D., et al. (2017). The steady enhancement of the Australian Summer Monsoon in the last 200 years. Sci. Rep., 7, 16166. https://doi.org/10.1038/s41598-017-16414-1.CrossRefGoogle ScholarPubMed
Gallet, B. and Young, W. R. (2014). Refraction of swell by surface currents. J. Mar. Res., 72, 105–126.CrossRefGoogle Scholar
Garcia-Herrera, R., et al. (2005a). CLIWOC: A climatological database for the world’s oceans 1750–1854. Clim. Change, 73, 1–12.CrossRefGoogle Scholar
García-Herrera, R., et al. (2005b). Description and general background to ships’ logbooks as a source of climatic data. Clim. Change, 73, 13–36.CrossRefGoogle Scholar
García-Herrera, R., et al. (2018). Understanding weather and climate of the last 300 years from ships’ logbooks. Wiley Interdisciplinary Reviews: Clim. Change, 9, e544. https://doi.org/10.1002/wcc.544.Google Scholar
Garreaud, R. D. and Munoz, R. C. (2005). The low-level jet off the west coast of subtropical South America: Structure and variability. Mon. Weather Rev., 133, 2246–2261.CrossRefGoogle Scholar
Golledge, N. R., et al. (2019). Global environmental consequences of twenty-first-century ice-sheet melt. Nature, 566, 7742, 65. https://doi.org/10.1038/s41586-019-0889-9.CrossRefGoogle ScholarPubMed
Gonzalez Rodriguez, M., et al. (2016). Brazilian coastal processes: Wind, wave climate and sea level. Chapter 2. In Short, A. D. and Klein, A. H. da F. (eds.), Brazilian Beach Systems. Coastal Research Library 17. Springer, Switzerland, pp. 37–66. 611pp. https://doi.org/10.1007/978-3-319-30394-9_1.Google Scholar
Goodwin, I. D. (2003). Unraveling climate influences on late Holocene sea-level and coastal evolution. In Mackay, A., et al. (eds.), Global Change in the Holocene. London: Edward Arnold, pp. 406–421.Google Scholar
Goodwin, I. D. (2005). A mid-shelf wave direction climatology for south-eastern Australia, and its relationship to the El Nino – Southern Oscillation, since 1877 AD. Int. J. Climatol., 25, 1715–1729.Google Scholar
Goodwin, I. D. and Grossman, E. E. (2003). Middle to late Holocene coastal evolution along the south coast of Upolu Island, Samoa. Mar. Geol., 202, 1–16.CrossRefGoogle Scholar
Goodwin, I. D., et al. (2016). Tropical and extratropical-origin storm wave types and their influence on the East Australian longshore sand transport system under a changing climate. J. Geophys. Res. Oceans, 121, 4833–4853.CrossRefGoogle Scholar
Goodwin, I. D., et al. (2023). Robbins Island: The index site for regional Last Interglacial sea level, wave climate and the subtropical ridge around Bass Strait, Australia. Quat. Sc. Rev., 305, 107996. https://doi.org/10.1016/j.quascirev.2023.107996.CrossRefGoogle Scholar
Goyal, R., et al. (2021). Historical and projected changes in the Southern Hemisphere surface westerlies. Geophys. Res. Lett., 48, e2020GL090849. https://doi.org/10.1029/2020GL090849.CrossRefGoogle Scholar
Gregory, J. M. (2019). Concepts and terminology for sea level: Mean, variability and change, both local and global. Surv. Geophys., 40, 1251–1289. https://doi.org/10.1007/s10712-019-09525-.Google Scholar
Grossman, E. E., et al. (1998). The Holocene sea-level highstand in the equatorial Pacific: Analysis of the insular paleosea-level database. Coral Reefs, 17, 309–327.CrossRefGoogle Scholar
Haigh, I. D., et al. (2019). The tides they are a-changin’: A comprehensive review of past and future non-astronomical changes in tides, their driving mechanisms and future implications. Rev. Geophys., 57. https://doi.org/10.1029/2018rg000636.Google Scholar
Hamlington, B. D., et al. (2013). Contribution of the Pacific Decadal Oscillation to global mean sea level trends. Geophys. Res. Lett., 40, 19, 5171–5175. https://doi.org/10.1002/grl.50950.CrossRefGoogle Scholar
Hamlington, B. D., et al. (2017). Separating decadal global water cycle variability from sea level rise. Sci. Rep., 7, 995. https://doi.org/10.1038/s41598-017-00875-5.CrossRefGoogle ScholarPubMed
Hamlington, B. D., et al. (2019). Uncovering the pattern of forced sea level rise in the satellite altimeter record. Geophys. Res. Lett., 46, 4844–4853. https://doi.org/10.1029/2018gl081386.CrossRefGoogle Scholar
Hamlington, B. D., et al. (2020). Understanding of contemporary regional sea-level change and the implications for the future. Rev. Geophys., 58, e2019RG000672. https://doi.org/10.1029/2019RG000672.CrossRefGoogle ScholarPubMed
Han, W., et al. (2017). Spatial patterns of sea level variability associated with natural internal climate modes. Surv. Geophys., 38, 1, 217–250. https://doi.org/10.1007/s10712-016-9386-y.CrossRefGoogle ScholarPubMed
Hannaford, M., et al. (2015). Early-nineteenth-century southern African precipitation reconstructions from ships’ logbooks. The Holocene, 25, 2, 379–390. https://doi.org/10.1177/0959683614557573.CrossRefGoogle Scholar
Hardman-Mountford, N. J., et al. (2003). Ocean climate of the South East Atlantic observed from satellite data and wind models. Prog. Oceanogr., 59, 181–221.CrossRefGoogle Scholar
Hasselmann, K., et al. (1973). Measurements of wind-wave growth and swell decay during the Joint North Sea Wave Project (JONSWAP). Deutsch. Hydrogr. Z., Suppl., A8, 12, 95pp.Google Scholar
Hay, C., et al. (2014). The sea-level fingerprints of ice-sheet collapse during interglacial periods. Quat. Sci. Rev., 87, 60–69.CrossRefGoogle Scholar
Hay, C. C., et al. (2017). On the robustness of Bayesian fingerprinting estimates of global sea level change. J. Clim., 30, 3025–3038.CrossRefGoogle Scholar
Headland, R. K., et al. (2023). Historical occurrence of Antarctic icebergs within mercantile shipping routes and the exceptional events of the 1890s. J. Glaciol., 1–13. https://doi.org/10.1017/jog.2023.80.CrossRefGoogle Scholar
Heidemann, H. and Ribbe, J. (2019). Marine heat waves and the influence of El Niño off Southeast Queensland, Australia. Front. Mar. Sci., 6, 56. https://doi.org/10.3389/fmars.2019.00056.CrossRefGoogle Scholar
Heimbach, P. and Hasselmann, K. (2000). Development and application of satellite retrievals of ocean wave spectra. In Halpern, D. (ed.), Satellites, Oceanography and Society. Elsevier, Amsterdam, pp. 5–33.Google Scholar
Hell, M. C., et al. (2009). Identifying ocean swell generation events from Ross Ice Shelf seismic data. J. Atmos. Ocean. Technol., 36, 2171–2189. https://doi.org/10.1175/JTECH-D-19-0093.1.Google Scholar
Hemer, M. A., et al. (2008). A classification of wave generation characteristics during large wave events on the Southern Australian margin. Cont. Shelf Res., 28(2008), 634–652.CrossRefGoogle Scholar
Hemer, M. A., et al. (2013). Projected changes in wave climate from a multi-model ensemble. Nat. Clim. Change, 3. https://doi.org/10.1038/NCLIMATE1791.CrossRefGoogle Scholar
Hinton, A. (1998). Tidal changes. Prog. Phys. Geogr., 22, 2, 282–294.CrossRefGoogle Scholar
Holbrook, N. J., et al. (2011). ENSO to multi-decadal time scale changes in East Australian Current transports and Fort Denison sea level: Oceanic Rossby waves as the connecting mechanism. Deep-Sea Res. Part 2. https://doi.org/10.1016/j.dsr2.2010.06.007.CrossRefGoogle Scholar
Holbrook, N. J., et al. (2019). A global assessment of marine heatwaves and their drivers. Nat. Commun., 10, 2624. https://doi.org/10.1038/s41467-019-10206-z.CrossRefGoogle ScholarPubMed
Holthuijsen, L. H. (2007). Waves in Oceanic and Coastal Waters. Cambridge University Press, 379pp.CrossRefGoogle Scholar
Hughes, C. W., et al. (2019). Sea level and the role of coastal-trapped waves in mediating the interaction between the coast and open ocean. Surv. Geophys. https://doi.org/10.1007/s10712-019-09535-x.CrossRefGoogle Scholar
Jacobs, S. S. (1991). On the nature and significance of the Antarctic Slope Front. Mar. Chem., 35, 9–24.CrossRefGoogle Scholar
Jevrejeva, S., et al. (2016). Coastal sea-level rise with warming above 2 degree. Proc. Natl. Acad. Sci. USA, 113, 47, 13342–13347. https://doi.org/10.1073/pnas.1605312113.CrossRefGoogle Scholar
Jevrejeva, S., et al. (2019). Probabilistic sea level projections at the coast by 2100. Surv. Geophys., 40, 1673–1696. https://doi.org/10.1007/s10712-019-09550-y.CrossRefGoogle Scholar
Kalnay, E., et al. (1996). The NCEP/NCAR 40-year reanalysis project. Bull. Am. Meteorol. Soc., 77, 437–470.2.0.CO;2>CrossRefGoogle Scholar
Kämpf, J. (2024). On the wind-driven formation of plankton patches in island wakes. J. Mar. Sci. Eng., 12, 193. https://doi.org/10.3390/jmse12010193.CrossRefGoogle Scholar
Kämpf, J. and Chapman, P. (2016). The Benguela current upwelling system. Chapter 7. In Kämpf, J. and Chapman, P., (eds.), Upwelling Systems of the World, Springer, Switzerland, pp. 251–314, 433pp. https://doi.org/10.1007/978-3-319-42524-5_7.CrossRefGoogle Scholar
Kataoka, T., et al. (2014). On the Ningaloo Niño/ Niña. Clim. Dyn., 43, 1463–1482.CrossRefGoogle Scholar
Kepert, J. D. and Smith, R. K. (1992). A simple model of the Australian West Coast Trough. Mon. Weather Rev., 120, 9, 2042–2055.2.0.CO;2>CrossRefGoogle Scholar
Können, G. P. and Koek, F. B. (2005). Description of the CLIWOC database. Clim. Change, 73(1–2), 117–130.CrossRefGoogle Scholar
Kopp, R. E., et al. (2015). Geographic variability of sea-level change. Curr. Clim. Change Rep., 1(3), 192–204. https://doi.org/10.1007/s40641-015-0015-5.CrossRefGoogle Scholar
Köppen, V. (1874). Uber die Abhangigkeit des klimatischen Charakters der Winde von ihrem Ursprunge (Dependence of climatological characteristics on the wind’s trajectory). Reportorium filr Meteor., 4, 15pp.Google Scholar
Kuttel, M., et al. (2010). The importance of ship log data: Reconstructing North Atlantic, European and Mediterranean sea level pressure fields back to 1750. Clim. Dyn., 34, 115–1128.CrossRefGoogle Scholar
Lee, T. and McPhaden, M. J. (2008). Decadal phase change in large-scale sea level and winds in the Indo- Pacific region at the end of the 20th century. Geophys. Res. Lett., 35, L01605. https://doi.org/10.1029/2007GL032419.CrossRefGoogle Scholar
Levitus, S., et al. (2000). Warming of the world ocean. Science, 287, 2225–2229. https://doi.org/10.1126/science.287.5461.2225.CrossRefGoogle Scholar
Levitus, S., et al. (2009). Global ocean heat content 1955–2008 in light of recently revealed instrumentation problems. Geophys. Res. Lett., 36, L07608. https://doi.org/10.1029/2008GL037155.CrossRefGoogle Scholar
Levitus, S., et al. (2012). World ocean heat content and thermosteric sea level change (0–2000 m), 1955–2010. Geophys. Res. Lett., 39, L10603. https://doi.org/10.1029/2012GL051106.CrossRefGoogle Scholar
Lewis, J. M. (1996). Winds over the world sea: Maury and Köppen. Bull. Am. Meteorol. Soc., 77, 5, 935–952.2.0.CO;2>CrossRefGoogle Scholar
Li, X., et al. (2015). Atlantic-induced pan-tropical climate change over the past three decades. Nat. Clim. Change, 6, 275–280. https://doi.org/10.1038/NCLIMATE2840.Google Scholar
Li, X.-M. (2016). A new insight from space into swell propagation and crossing in the global oceans. Geophys. Res. Lett., 43, 5202–5209. https://doi.org/10.1002/2016GL068702.CrossRefGoogle Scholar
Lima, D. C. A., et al. (2018). A global view of coastal low-level wind jets using an ensemble of reanalyses. J. Clim., 31, 1525–1546.CrossRefGoogle Scholar
Lima, D. C. A., et al. (2019). A climatological analysis of the Benguela coastal low-level jet. J. Geophys, Res. Atmos., 124, 3960–3978. https://doi.org/10.1029/2018JD028944.Google Scholar
Lisiecki, L. E. and Raymo, M. E. (2005). A Pliocene-Pleistocene stack of 57 globally distributed benthic δ18O records. Paleooceanogr., 20, PA1003.Google Scholar
Lorrey, A. M., et al. (2022). Meteorological data rescue: Citizen science lessons learned from Southern Weather Discovery. Patterns, 3, 100495. https://doi.org/10.1016/j.patter.2022.100495.CrossRefGoogle ScholarPubMed
Lucas, C., Timbal, B. and Nguyen, H. (2014). The expanding tropics: A critical assessment of the observational and modeling studies. WIREs Clim. Change, 5, 89–112. https://doi.org/10.1002/wcc.251.CrossRefGoogle Scholar
MacAyeal, D. R., et al. (2006). Transoceanic wave propagation links iceberg calving margins of Antarctica with storms in tropics and Northern Hemisphere. Geophys. Res. Lett., 33, L17502. https://doi.org/10.1029/2006gl027235.CrossRefGoogle Scholar
Mackintosh, N. A. and Herdman, H. F. P. (1940). Distribution of the pack-ice in the Southern Ocean. Discovery Rep., 19, 285–296.Google Scholar
Maher, N., et al. (2018). ENSO change in climate projections: Forced response or internal variability? Geophys. Res. Lett., 45, 11390–11398. https://doi.org/10.1029/2018GL079764.CrossRefGoogle Scholar
Mantua, N. J., et al. (1997). A Pacific interdecadal climate oscillation with impacts on salmon production. Bull. Am. Meteorol. Soc., 78, 1069–1079.2.0.CO;2>CrossRefGoogle Scholar
Martin, S., et al. (2022). Comparison of Antarctic iceberg observations by Cook in 1772–75, Halley in 1700, Bouvet in 1739 and Riou in 1789 with modern data. J. Glaciol., 1–8. https://doi.org/10.1017/jog.2022.111.Google Scholar
Massel, S. R. (1996). Ocean Surface Waves: Their physics and prediction. Advanced Series on Ocean Engineering, Volume 11, 491pp, World Scientific Publishing, Singapore.Google Scholar
Massom, R. A., et al. (2018). Antarctic ice shelf disintegration triggered by sea ice loss and ocean swell. Nature, 558, 383–389. https://doi.org/10.1038/s41586-018-0212-1.CrossRefGoogle ScholarPubMed
Maury, M. F. (1848–1860). Wind and Current Charts. [U.S. Hydrographic Office], Manuscripts Division, Library of Congress, Washington, DC.Google Scholar
Maury, M. F. (ed). (1854). Maritime conference held at Brussels for devising a uniform system of meteorological observations at sea, August and September, 1853. In Explanations and Sailing Directions to Accompany the Wind and Weather Current Charts, 6th ed. EC and J Biddle: Philadelphia, pp. 54–96.Google Scholar
McGregor, S., et al. (2014). Recent Walker circulation strengthening and Pacific cooling amplified by Atlantic warming. Nat. Clim. Change, 4, 10, 888–892. https://doi.org/10.1038/NCLIMATE2330.CrossRefGoogle Scholar
McGregor, H. V., et al. (2015). Robust global ocean cooling trend for the pre-industrial Common Era. Nat. Geosci., 8, 9, 671–677. https://doi.org/10.1038/NGEO2510.CrossRefGoogle Scholar
McKay, N. P., et al. (2011). The role of ocean thermal expansion in Last Interglacial sea level rise. Geophys. Res. Lett., 38, L14605. https://doi.org/10.1029/2011GL04828.Google Scholar
Menne, M. J., et al. (2018). The global historical climatology network monthly temperature dataset, version 4. J. Clim., 31, 9835–9854. https://doi.org/10.1175/JCLI-D-18-0094.1.CrossRefGoogle Scholar
Merrifield, M. A. and Maltrud, M. E. (2011). Regional sea level trends due to a Pacific trade wind intensification. Geophys. Res. Lett., 38, L21605.CrossRefGoogle Scholar
Meucci, A., et al. (2020). Comparison of wind speed and wave height trends from twentieth-century models and satellite altimeters. J. Clim., 33, 2, 611–624. https://doi.org/10.1175/JCLI-D-19-0540.1.CrossRefGoogle Scholar
Milne, G. A. and Mitrovica, J. X. (1998). Postglacial sea-level change on a rotating Earth. Geophys. J. Int., 133, 1, 1–19. https://doi.org/10.1046/j.1365-246x.1998.1331455.x.CrossRefGoogle Scholar
Milne, G. A., et al. (1999). Near-field hydro-isostasy: The implementation of a revised sea-level equation. Geophys. J. Int., 139, 464–482.CrossRefGoogle Scholar
Mitrovica, J. X. and Peltier, W. R. (1991). On postglacial geoid subsidence over the equatorial oceans. J. Geophys. Res., 96, 53–71.Google Scholar
Mitrovica, J. X., et al. (2001). Glacial isostatic adjustment on a rotating earth. Geophys. J. Int., 147, 3, 562–578. https://doi.org/10.1046/j.1365-246x.2001.01550.x.Google Scholar
Mitrovica, J. X., et al. (2001). Recent mass balance of polar ice sheets inferred from patterns of global sea-level change. Nature, 409, 1026–1029.CrossRefGoogle ScholarPubMed
Mitrovica, J. X. and Milne, G. A. (2002). On the origin of late Holocene sea-level highstands within equatorial ocean basins. Quat. Sci. Rev., 21(20–22), 2179–2190. https://doi.org/10.1016/S0277-3791(02)00080-X.CrossRefGoogle Scholar
Mitrovica, J. X., et al. (2018). Quantifying the sensitivity of sea level change in coastal localities to the geometry of polar ice mass flux. J. Clim., 31, 3701–3709. https://doi.org/10.1175/JCLI-D-17-0465.1.CrossRefGoogle Scholar
Mörner, N.-A. (1976). Eustasy and geoid changes. J. Geol., 84, 123–151.CrossRefGoogle Scholar
Mortlock, T. R. and Goodwin, I. D. (2015). Directional wave climate and power variability in the Tasman Sea. Cont. Shelf Res., 98, 36–53.CrossRefGoogle Scholar
Mortlock, T. R. and Goodwin, I. D. (2016). Impacts of enhanced central Pacific ENSO on wave climate and headland-bay beach morphology. Cont. Shelf Res., 120, 14–25.CrossRefGoogle Scholar
Mortlock, T. R., et al. (2020). Influence of the subtropical ridge on directional wave power in the Southeast Indian Ocean. Int. J. Climatol., 40, 12, 5352–5367. https://doi.org/10.1002/joc.6522.CrossRefGoogle Scholar
Munk, W. H. (1947). Tracking storms by forerunners of swell. J. Met., 4, 2, 45–57.2.0.CO;2>CrossRefGoogle Scholar
Munk, W. H. (1950). Origin and generation of waves, Proc. 1st Conf. Coastal Engineering (Long Beach), New York, ASCE, pp. 1–4.Google Scholar
Munk, W. H. and Traylor, M. A. (1947). Refraction of ocean waves: A process linking under- water topography to beach erosion. J. Geol., LV, 1, 1–26.Google Scholar
Munk, W. H. and Snodgrass, F. E. (1957). Measurements of southern swell at Guadalupe Island. Deep-Sea Res., 4, 272–286. https://doi.org/10.1016/0146-6313(56)90061-2.Google Scholar
Munk, W. H., et al. (1963a). Directional recording developed wind seas based on the similarity theory of S. A. Kitaigorodskii. J. Geophys. Res., 69, 5181–5190.Google Scholar
Munk, W. H., et al. (1963b). Directional recording of swell from distant storms. Philos. Trans. R. Soc. London, Ser. A, 255, 505–584.Google Scholar
Munk, W. H., et al. (2013). Directional recording of swell from distant storms. Phil. Trans. R. Soc. A, 371, 20130039. http://dx.doi.org/10.1098/rsta.2013.0039.CrossRefGoogle ScholarPubMed
Nakada, M. and Lambeck, K. (1989). Late Pleistocene and Holocene sea-level variations in Australian region and mantle rheology. Geophys. J. Int., 96, 497–517.CrossRefGoogle Scholar
Nerem, R. S., et al. (2018). Climate-change–driven accelerated sea-level rise detected in the altimeter era. Proc. Natl. Acad. Sci. USA, 115, 9, 2022–2025. https://doi.org/10.1073/pnas.1717312115.CrossRefGoogle ScholarPubMed
Newman, M., et al. (2016). The Pacific Decadal Oscillation, revisited. J. Clim., 29, 12, 4399–4427. https://doi.org/10.1175/JCLI-D-15-0508.1.CrossRefGoogle Scholar
Nhantumbo, B. J., et al. (2020). The relationship between coastal sea level variability in South Africa and the Agulhas Current. J. Mar. Syst., 211(2020), 103422.CrossRefGoogle Scholar
Nhantumbo, B. J., et al. (2021). Atmospheric and climatic drivers of tide gauge sea level variability along the East and South Coast of South Africa. J. Mar. Sci. Eng., 9, 924. https://doi.org/10.3390/jmse9090924.CrossRefGoogle Scholar
Nicholson, I. (1990). Log of Logs Volume 1. Roebuck Society, Canberra.Google Scholar
Nicholson, I. (1993). Log of Logs Volume 2. Roebuck Society, Canberra.Google Scholar
Nicholson, I. (1999). Log of Logs Volume 3. Roebuck Society, Canberra.Google Scholar
Nicholson, S. E. (2010). A low-level jet along the Benguela coast, an integral part of the Benguela current ecosystem. Clim. Change, 99, 613–624.CrossRefGoogle Scholar
Norrgard, S. (2017). Royal Navy logbooks as secondary sources and their use in climatic investigations: Introducing the log-board. Int. J. Climatol., 37, 2027–2036. https://doi.org/10.1002/joc.4832.CrossRefGoogle Scholar
Odériz, I., et al. (2021). Natural variability and warming signals in global ocean wave climates. Geophys. Res. Lett., 48, e2021GL093622. https://doi.org/10.1029/2021GL093622.CrossRefGoogle Scholar
Parish, T. R. (1983). On the influence of the Antarctic Peninsula on the wind field over the western Weddell Sea. J. Geophys. Res., 88, C4, 2684–2692.Google Scholar
Parish, T. R. (1992). On the role of katabatic winds in forcing large-scale tropospheric motions. J. Atmos. Sci., 49(15), 1374–1385.Google Scholar
Pariwono, J. I., et al. (1986). Long-period variations of sea level in Australia. Geophys. J. R. Astron. Soc., 87, 43–54.CrossRefGoogle Scholar
Peltier, W. R. (1998). Postglacial variations in the level of the sea: Implications for climate dynamics and solid-earth geophysics. Rev. Geophys., 36(4), 603–689.CrossRefGoogle Scholar
Peltier, W. R. (2001). Global isostatic adjustment and modern instrumental records of relative sea level history. In Douglas, B. C., et al. (eds.), Sea-level Rise History and Consequences. Academic Press, San Diego, pp. 65–95.Google Scholar
Peltier, W. R., et al. (2015). Space geodesy constrains ice-age terminal deglaciation: The global ICE-6G_C (VM5a) model. J. Geophys. Res., 120, 450–487.CrossRefGoogle Scholar
Pepler, A. S. and Hope, P. (2018). Orography drives the semistationary West Australian summer trough. Geophys. Res. Lett., 45, 9981–9986. https://doi.org/10.1029/2018GL079312.CrossRefGoogle Scholar
Piecuch, C. G. and Quinn, K. J. (2016). El Niño, La Niña, and the global sea level budget. Ocean Sci., 12(6), 1165–1177. https://doi.org/10.5194/os-12-1165-2016.CrossRefGoogle Scholar
Pierson, W. J., Jr. and Moskowitz, L. (1964). A proposed spectral form for fully developed wind seas based on the similarity theory of S. A. Kitaigorodskii. J. Geophys. Res., 69, 5181–5190.Google Scholar
Ponte, R. M. (2006). Low-frequency sea level variability and the inverted barometer effect. J. Atmos. Ocean. Technol., 23, 619–629.CrossRefGoogle Scholar
Prieto, M. R., et al. (2004). Early records of icebergs in the South Atlantic Ocean from Spanish documentary sources. Clim. Change, 66, 29–48.Google Scholar
Pugh, D. T. (2004). Changing Sea Levels. Effects of tides, weather and climate. Cambridge University Press, UK, pp265.Google Scholar
Pugh, D. T. and Woodworth, P. L. (2014). Sea-Level Science: Understanding Tides, Surges, Tsunamis and Mean Sea- Level Changes. Cambridge University Press, Cambridge. ISBN 9781107028197.CrossRefGoogle Scholar
Rahn, D. A. and Garreaud, R. D. (2014). A synoptic climatology of the near-surface wind along the west coast of South America. Int. J. Climatol., 34, 780–792.CrossRefGoogle Scholar
Ranjha, R., et al. (2013). Global distribution and seasonal variability of coastal low-level jets derived from ERA-Interim reanalysis. Tellus A: Dyn. Meteorol. Oceanogr., 65(1), 20412. https://doi.org/10.3402/tellusa.v65i0.20412.CrossRefGoogle Scholar
Renault, L., et al. (2009). Impact of atmospheric coastal jet off central Chile on sea surface temperature from satellite observations (2000–2007). J. Geophys. Res., 114, C08006. https://doi.org/10.1029/2008JC005083.Google Scholar
Ribal, A. and Young, I. R. (2019). 33 years of globally calibrated wave height and wind speed data based on altimeter observations. Sci. Data, 6, 77. https://doi.org/10.1038/s41597-019-0083-9.Google ScholarPubMed
Ribbe, J. and Brieva, D. (2016). A western boundary current eddy characterisation study. Estuarine. Coast. Shelf Sci., 183, 203–212.CrossRefGoogle Scholar
Ribbe, J., et al. (2018). Frontal eddies along a western boundary current. Cont. Shelf Res., 165, 51–59.CrossRefGoogle Scholar
Ridgway, K. R. (2007a). Long-term trend and decadal variability of the southward penetration of the East Australian Current. J. Geophys. Res., 34, L13613. https://doi.org/10.1029/2007GL030393.Google Scholar
Ridgway, K. R. (2007b). Seasonal circulation around Tasmania: An interface between eastern and western boundary dynamics. J. Geophys. Res., 112, C10016. https://doi.org/10.1029/2006JC003898.Google Scholar
Ridgway, K. R. and Godfrey, J. S. (1997). Seasonal cycle of the East Australian Current. J. Geophys. Res., 102(C10), 22921–22936.Google Scholar
Ridgway, K. and Dunn, J. (2003). Mesoscale structure of the mean East Australian Current system and its relationship with topography. Prog. Oceanogr., 56(2), 189–222.CrossRefGoogle Scholar
Ridgway, K. R. and Dunn, J. R. (2007). Observational evidence for a Southern Hemisphere oceanic supergyre. J. Geophys. Res., 34, L13612. https://doi.org/10.1029/2007GL030392.Google Scholar
Rignot, E., et al. (2019). Four decades of Antarctic Ice Sheet mass balance from 1979–2017. Proc. Natl. Acad. Sci. USA, 116(4), 1095–1103. https://doi.org/10.1073/pnas.1812883116.CrossRefGoogle ScholarPubMed
Robson, N. A., et al. (2017). Use of particle tracking to determine optimal release dates and locations for rehabilitated neonate sea turtles. Front. Mar. Sci., 4, 173. https://doi.org/10.3389/fmars.2017.00173.CrossRefGoogle Scholar
Roughan, M. and Middleton, J. H. (2002). A comparison of observed upwelling mechanisms off the east coast of Australia. Cont. Shelf Res., 22(17), 2551–2572. https://doi.org/10.1016/S0278-4343(02)00101-2.CrossRefGoogle Scholar
Rubin, M. J. (1982). James Cook’s scientific programme in the Southern Ocean, 1772–75. Polar Rec., 21(130), 33–49.CrossRefGoogle Scholar
Rueda, A., et al. (2018). Marine climate variability based on weather patterns for a complicated island setting: The New Zealand case. Int. J. Climatol., 39, 1777–1786. https://doi.org/10.1002/joc.5912.Google Scholar
Ruiz-Etcheverry, L. A. and Saraceno, M. (2020). Sea level trend and fronts in the South Atlantic Ocean. Geosci., 10, 218. https://doi.org/10.3390/geosciences10060218.CrossRefGoogle Scholar
Russell, H. C. (1895). Icebergs in the Southern Ocean. J. Proc. Roy. Soc. NSW, 29, www.biodiversitylibrary.org/item/131951.Google Scholar
Russell, H. C. (1897). Icebergs in the Southern Ocean, No. 2. J. Proc. Roy. Soc. NSW, 31, www.biodiversitylibrary.org/item/130955.Google Scholar
Saha, S., et al. (2010). The NCEP climate forecast system reanalysis. Bull. Am. Meteorol. Soc., 91(8), 1015–1057.CrossRefGoogle Scholar
Santamaria-Aguilar, S., et al. (2017). Long-term trends and variability of water levels and tides in Buenos Aires and Mar del Plata, Argentina. Front. Mar. Sci. 4, 380. https://doi.org/10.3389/fmars.2017.00380.Google Scholar
Schaeffer, A., et al. (2013). Cross-shelf dynamics in a western boundary current. Implications for upwelling. J. Phys. Oceanogr., 43, 1042–1059. https://doi.org/10.1175/JPO-D-12–0177.1.CrossRefGoogle Scholar
Schloesser, F., et al. (2019). Antarctic iceberg impacts on future Southern Hemisphere climate. Nat. Clim. Change, 9(9), 672–677. https://doi.org/10.1038/s41558-019-0546-1.CrossRefGoogle Scholar
Schumann, E. H. and Brink, K. H. (1990). Coastal-trapped waves off the coast of South Africa: Generation, propagation and current structures. J. Phys. Oceanogr., 20, 1206–1218. https://doi.org/10.1175/1520-0485(1990)020<1206:CTWOTC>2.0.CO;2.2.0.CO;2>CrossRefGoogle Scholar
Schwerdtfeger, W. (1975). The effect of the Antarctic Peninsula on the temperature regime of the Weddell Sea. Mon. Weather Rev., 103, 45–51.2.0.CO;2>CrossRefGoogle Scholar
Seidel, D. J., Fu, Q., Randel, W. J. and Reichler, T. J. (2008). Widening of the tropical belt in a changing climate. Nat. Geosci., 1(1), 21–24. https://doi.org/10.1038/ngeo.2007.38.CrossRefGoogle Scholar
Semedo, A., et al. (2011). A global view on the wind sea and swell climate and variability from ERA-40. J. Clim., 24(5), 1461–1479. https://doi.org/10.1175/2010JCLI3718.1.CrossRefGoogle Scholar
Sen Gupta, A. and England, M. H. (2006). Coupled ocean–atmosphere– ice response to variations in the Southern Annular Mode. J. Clim., 19, 4457–4486. https://doi.org/10.1175/JCLI3843.1.CrossRefGoogle Scholar
Short, A. D., (ed.), (1999). Beach and Shoreface Morphodynamics. John Wiley and Sons, Chichester, 379pp.Google Scholar
Smith, S. R., et al. (2019). Ship-based contributions to global ocean, weather, and climate observing systems. Front. Mar. Sci., 6, 434. https://doi.org/10.3389/fmars.2019.00434.CrossRefGoogle Scholar
Snodgrass, F. E., et al. (1966). Propagation of ocean swell across the Pacific. Phil. Trans. Roy. Soc. London, A, 259, 431–497.Google Scholar
Spence, P., et al. (2014). Rapid subsurface warming and circulation changes of Antarctic coastal waters by poleward shifting winds. Geophys Res Lett, 41, 4601–4610.CrossRefGoogle Scholar
Spence, P., et al. (2017). Localized rapid warming of West Antarctic subsurface waters by remote winds. Nat. Clim. Change, 7, 595–603. https://doi.org/10.1038/NCLIMATE3335.CrossRefGoogle Scholar
Stammer, D., et al. (2013). Causes for contemporary regional sea-level changes. Ann. Rev. Mar. Sci., 5, 21–46. https://doi.org/10.1146/annurev-marine-121211-172406.CrossRefGoogle ScholarPubMed
Steinberg, C. R. and Lawrey, E. (2018). Circulation and upwelling. In eAtlas.org.au, Australia’s tropical land and seas.Google Scholar
Stenchikov, G., et al. (2009). Volcanic signals in oceans. J. Geophys. Res., 114, D16104. https://doi.org/10.1029/2008JD011673.Google Scholar
Sterle, A. and Caires, S. (2005). Climatology, variability and extrema of ocean waves: The web-based KNMI/ERA-40 wave atlas. Int. J. Climatol., 25(7), 963–977. https://doi.org/10.1002/joc.1175.Google Scholar
Storlazzi, C. D. and Wingfield, D. K. (2005). Spatial and temporal variations in oceanographic and meteorologic forcing along the Central California Coast, 1980–2002. Scientific Investigations Report 2005–5085, U.S. Department of the Interior U.S. Geological Survey, 39pp.CrossRefGoogle Scholar
Sturges, W. and Hong, B. G. (2001). Decadal variability of sea level. In Douglas, B. C., et al. (eds.), Sea Level Rise: History and Consequences. Academic Press, San Diego, 165–180.Google Scholar
Sun, X., et al. (2019). Land–atmosphere–ocean interactions in the southeastern Atlantic: Interannual variability. Clim. Dyn., 52, 539–561. https://doi.org/10.1007/s00382-018-4155-x.CrossRefGoogle Scholar
Teleti, P. R., et al. (2019). A historical Southern Ocean climate dataset from whaling ships’ logbooks. Geosci. Data J., 6, 30–40. https://doi.org/10.1002/gdj3.65.CrossRefGoogle Scholar
Thompson, A. F., et al. (2018). The Antarctic Slope Current in a changing climate. Rev. Geophys., 56, 741–770. https://doi.org/10.1029/2018RG000624.CrossRefGoogle Scholar
Thompson, D. W. J. and Solomon, S. (2002). Interpretation of recent Southern Hemisphere climate change. Science, 296, 895–899.Google ScholarPubMed
Thompson, D. W. J., et al. (2011). Signatures of the Antarctic ozone hole in Southern Hemisphere surface climate change. Nat. Geosci., 4(11), 741–749. https://doi.org/10.1038/NGEO1296.CrossRefGoogle Scholar
Timmermann, A., et al. (2010). Wind effects on past and future regional sea level trends in the southern Indo-Pacific. J. Clim., 23, 4429–4437.CrossRefGoogle Scholar
Toggweiler, J. R. and Russell, J. (2008). Ocean circulation in a warming climate. Nature, 45. https://do10.1038/nature06590.Google Scholar
Towson, J. T. (1858). Icebergs in the Southern Ocean. Hist. Soc. Lancashire Cheshire Transact., 10, 239–254.Google Scholar
Trouet, V. and van Oldenborgh, G. J. (2013). KNMI Climate Explorer: A web-based research tool for high-resolution paleoclimatology. Tree-Ring Res., 69, 3–14.Google Scholar
Trouet, V., et al. (2016). Shipwreck rates reveal Caribbean tropical cyclone response to past radiative forcing. Proc. Natl. Acad. Sci. USA, 113(12), 3169–3174. https://doi.org/10.1073/pnas.1519566113.CrossRefGoogle ScholarPubMed
Veitch, J. A. and Penven, P. (2017). The role of the Agulhas in the Benguela Current system: A numerical modeling approach. J. Geophys. Res., Oceans, 122, 3375–3393. https://doi.org/10.1002/2016JC012247.CrossRefGoogle Scholar
Von Storch, H. and van Zwiers, F. W. (1999). Statistical Analysis in Climate Research. Cambridge University Press, Cambridge, 484pp.Google Scholar
Waelbroeck, C., et al. (2002). Sea-level and deep water temperature changes derived from benthic foraminifera isotopic records. Quat. Sci. Rev., 21, 295–305.CrossRefGoogle Scholar
Wang, J., et al. (2021). Reconciling global mean and regional sea level change in projections and observations. Nat. Commun., 12, 990. https://doi.org/10.1038/s41467-021-21265-6.Google ScholarPubMed
Weeks, S. J., et al. (2010). The Capricorn Eddy: A prominent driver of the ecology and future of the southern Great Barrier Reef. Coral Reefs, 29, 975–985. https://doi.org/10.1007/s00338-010-0644-z.CrossRefGoogle Scholar
Weisse, R. and von Storch, H. (2010). Marine Climate and Climate Change: Storms, Wind Waves and Storm Surges. Praxis Publishing Ltd, Chichester, 419pp.CrossRefGoogle Scholar
Wheeler, D. and García-Herrera, R. (2008). Ships’ logbooks in climatological research. Ann. New York Acad. Sci., 1146, 1–15.CrossRefGoogle ScholarPubMed
Wheeler, D., et al. (2009). Atmospheric circulation and storminess derived from Royal Navy logbooks: 1685 to 1750. Clim. Change, 101, 257–280. https://doi.org/10.1007/s10584-009-9732-x.Google Scholar
Wheeler, D., et al. (2010). Air circulation and storminess in the Atlantic-European region derived from logbooks: 1658 to 1750. Clim. Change, 101, 257–280.Google Scholar
WhitworthIII, T., et al. (1998). Water masses and mixing near the Antarctic Slope Front. Antarctic Research Series, 75, 1–27.Google Scholar
Widlansky, M. J., et al. (2015). Future extreme sea level seesaws in the tropical Pacific. Sci. Adv., 1(8), e1500560–e1500560.CrossRefGoogle ScholarPubMed
Wilkinson, C., et al. (2011). Recovery of logbooks and international marine data: The RECLAIM project. Int. J. Climatol., 31, 968–979. https://doi.org/10.1002/joc.2102.CrossRefGoogle Scholar
Williamson, F., et al. (2015). New directions in hydro-climatic histories: Observational data recovery, proxy records and the Atmospheric Circulation Reconstructions over the Earth (ACRE) initiative in Southeast Asia. Geoscience Letters, 2, 2. https://doi.org/10.1186/s40562-015-0018-z.CrossRefGoogle Scholar
Wolanski, E., et al. (1984). Island wakes in shallow coastal waters. J. Geophys. Res., 89, 10553–10569.Google Scholar
Woodruff, S. D. (2007). Archival of data other than in IMMT format: The International Maritime Meteorological Archive (IMMA) Format. In Second Session of the JCOMM Expert Team on Marine Climatology (ETMC), Geneva, Switzerland, 26–27 March 2007. JCOMM Meeting Report No. 50, 68–101. http://icoads.noaa.gov/e-doc/imma/.Google Scholar
Woodworth, P. L., et al. (eds.), (1992). Sea Level Changes: Determination and Effects. Vol. 69. American Geophysical Union, Washington, DC, pp. 1–196.CrossRefGoogle Scholar
Woodworth, P. L., et al. (2019). Forcing factors affecting sea-level changes at the coast. Surv. Geophys., 1–47. https://doi.org/10.1007/s10712-019-09531-1.CrossRefGoogle Scholar
Wu, L., et al. (2012). Enhanced warming over the global subtropical western boundary currents. Nat. Clim. Change, 2, 3, 161–166. https://doi.org/10.1038/NCLIMATE1353.CrossRefGoogle Scholar
Wu, Q., et al. (2017). Variability and change of sea level and its components in the Indo-Pacific region during the altimetry era. J. Geophys. Res., Oceans, 122, 1862–1881. https://doi.org/10.1002/2016JC012345.CrossRefGoogle Scholar
Wunsch, C. (1992). Decade to century changes in the ocean circulation. Oceanogr., 5(2), 99–106.CrossRefGoogle Scholar
Wyrtki, K. (1985). Sea level fluctuations in the Pacific during the 1982–83 El Niño. Geophys. Res. Lett., 12(3), 125–128.CrossRefGoogle Scholar
Yang, Y., et al. (2008). Observations of the trade wind wakes of Kauai and Oahu. Geophys. Res. Lett., 35, L04807. https://doi.org/10.1029/2007GL031742.CrossRefGoogle Scholar
Young, I. R. and Donelan, M. A. (2018). On the determination of global ocean wind and wave climate from satellite observations. Remote Sensing of Environ., 215, 228–241.CrossRefGoogle Scholar
Young, I. R., et al. (2020). The global wind resource observed by scatterometer. Remote Sens., 12, 2920. https://doi.org/10.3390/rs12182920.CrossRefGoogle Scholar
Zhang, X. and Church, J. A. (2012). Sea level trends, interannual and decadal variability in the Pacific Ocean. Geophys. Res. Lett., 39, L21701. https://doi.org/10.1029/2012GL053240.CrossRefGoogle Scholar
Ackerley, D., et al. (2011). Using synoptic type analysis to understand New Zealand climate during the mid-Holocene. Clim. Past., 7(4), 1189–1207.CrossRefGoogle Scholar
Andersen, H., et al. (2019). Synoptic-scale controls of fog and low clouds in the Namib Desert. Atmos. Chem. Phys. Discuss., 20(6), 3415–3438. https://doi.org/10.5194/acp-2019-924.Google Scholar
Arias, P. A., et al. (2021). Hydroclimate of the Andes Part II: Hydroclimate variability and sub- continental patterns. Front. Earth Sci., 8, 505467. https://doi.org/10.3389/feart.2020.505467.CrossRefGoogle Scholar
Bannister, D. and King, J. C., (2019). The characteristics and temporal variability of föhn winds at King Edward Point, South Georgia. Int. J. Climatol., 40, 2778–2794. https://doi.org/10.1002/joc.6366.Google Scholar
Bárdossy, A., et al. (2015). Circulation patterns identified by spatial rainfall and ocean wave fields in Southern Africa. Front. Environ. Sci., 3, 31. https://doi.org/10.3389/fenvs.2015.00031.CrossRefGoogle Scholar
Barimalala, R. F., et al. (2018). Madagascar influence on the South Indian Ocean Convergence Zone, the Mozambique Channel Trough and southern African rainfall. Geophys. Res. Lett., 45, 11380–11389. https://doi.org/10.1029/2018GL079964.CrossRefGoogle Scholar
Barriopedro, D., et al. (2014). Witnessing North Atlantic westerlies variability from ships’ logbooks (1685–2008). Clim. Dyn., 43, 939–955. https://doi.org/10.1007/s00382-013-1957-8.CrossRefGoogle Scholar
Barry, R. G. and Chorley, R. J. (2003). Atmosphere, Weather and Climate, 8th ed. Routledge, London, UK, 421pp.Google Scholar
Behera, S. K. and Yamagata, T. (2001). Subtropical SST dipole events in the southern Indian Ocean. Geophys. Res. Lett., 28(2), 327–330.CrossRefGoogle Scholar
Bettolli, L., et al. (2010). Synoptic weather types in the south of South America and their relationship to daily rainfall in the core crop-producing region in Argentina. Aust. Meteorol. Oceanograph. J., 60, 37–48.Google Scholar
Birnbaum, G., et al. (2006). Synoptic situations causing high precipitation rates on the Antarctic Plateau: Observations from Kohnen Station, Dronning Maud Land. Antarct. Sci., 18(2), 279–288.CrossRefGoogle Scholar
Blackmore, K. and Goodwin, I. D. (2009). Report No. 3. Climate change impacts for the Hunter, Lower North Coast and Central Coast Region of NSW. Newcastle Innovation Technical Report, 99pp. The University of Newcastle, Australia.Google Scholar
Blamey, R. and Reason, C. J. C. (2007). Relationships between Antarctic sea-ice and South African winter rainfall. Clim. Res., 33, 183–193.CrossRefGoogle Scholar
Blamey, R. C., et al. (2018). The influence of atmospheric rivers over the South Atlantic on winter rainfall in South Africa. J. Hydrometeorol., 19, 127–142. https://doi.org/10.1175/JHM-D-17-0111.1.CrossRefGoogle Scholar
Bonshoms, M., et al. (2019). Dry season circulation-type classification applied to precipitation and temperature in the Peruvian Andes. Int. J. Climatol., 40, 6473–6491. https://doi.org/10.1002/joc.6593.Google Scholar
Bozkurt, D., et al. (2018). Foehn event triggered by an atmospheric river underlies record-setting temperature along continental Antarctica. J. Geophys. Res. Atmos., 123, 3871–3892. https://doi.org/10.1002/2017JD027796.CrossRefGoogle Scholar
Brands, S., et al. (2023). A global climate model performance atlas for the Southern Hemisphere extratropics based on regional atmospheric circulation patterns. Geophys. Res. Lett., 50, e2023GL103531. https://doi.org/10.1029/2023GL103531.CrossRefGoogle Scholar
Burls, N. J. (2019). The Cape Town ‘Day Zero’ drought and Hadley cell expansion. npj Clim. Atmos., 2, 27. https://doi.org/10.1038/s41612-019-0084-6.Google Scholar
Cai, W., et al. (2020). Climate impacts of El Niño-Southern Oscillation on South America. Nat. Rev. Earth Environ., 1, 215–231. https://doi.org/10.1038/s43017-020-0040-3.CrossRefGoogle Scholar
Camus, P., et al. (2014). A weather-type statistical downscaling framework for ocean wave climate, J. Geophys. Res. Oceans, 119, https://doi.org/10.1002/2014JC010141.CrossRefGoogle Scholar
Carpenedo, C. B., et al. (2022). Sea ice in the Weddell Sea and its relationship with the South Atlantic Subtropical Highand precipitation in South America. An. Acad. Bras. Cienc., 94(Suppl. 4): e20211623. https://doi.org/10.1590/0001-3765202220211623.CrossRefGoogle Scholar
Catto, J. L., et al. (2012). Relating global precipitation to atmospheric fronts. Geophys. Res. Lett., 39, L10805. https://doi.org/10.1029/2012GL051736.CrossRefGoogle Scholar
Charles, S. P., et al. (2015). Hydroclimate of the Pilbara: Past, present and future. A report to the Government of Western Australia and industry partners from the CSIRO Pilbara Water Resource Assessment. CSIRO Land and Water, Australia.Google Scholar
Chen, H. W., et al. (2013). A robust mode of climate variability in the Arctic: The Barents Oscillation. Geophys. Res. Lett., 40, 2856–2861. https://doi.org/10.1002/grl.50551.CrossRefGoogle Scholar
Coelho, C. A. S., et al. (2016). Precipitation diagnostics of an exceptionally dry event in São Paulo, Brazil. Theor. Appl. Climatol., 125(3–4), 769–784. https://doi.org/10.1007/s00704-015-1540-9.CrossRefGoogle Scholar
Coggins, J. H. J., et al. (2014). Synoptic climatology of the Ross Ice Shelf and Ross Sea region of Antarctica: k-means clustering and validation, Int. J. Climatol., 34(7), 2330–2348. https://doi.org/10.1002/joc.3842.CrossRefGoogle Scholar
Coggins, J. H. J., et al. (2015). An assessment of the ocean wave climate of New Zealand as represented in Kidson’s synoptic types. Int. J. Climatol., 2496, 2481–2496. https://doi.org/10.1002/joc.4507.Google Scholar
Coggins, J. H. J. and McDonald, A. J. (2015). The influence of the Amundsen Sea Low on the winds in the Ross Sea and surroundings: Insights from a synoptic climatology. J. Geophys. Res. Atmos., 120, 2167–2189, https://doi.org/10.1002/2014JD022830.CrossRefGoogle Scholar
Cohen, L., et al. (2013). Synoptic weather types for the Ross Sea region, Antarctica. J. Clim., 26(2), 636–649. https://doi.org/10.1175/JCLI-D-11-00690.1.CrossRefGoogle Scholar
Compo, G. P., et al. (2011). The twentieth century reanalysis project. Q. J. R. Meteorol. Soc., 137, 1–28. https://doi.org/10.1002/qj.776.CrossRefGoogle Scholar
Cortesi, N., et al. (2021). Yearly evolution of Euro-Atlantic weather regimes and of their sub-seasonal predictability. Clim. Dyn., 56(11–12), 3933–3964. https://doi.org/10.1007/s00382-021-05679-y.Google Scholar
Cretat, J., et al. (2019). The Angola Low: Relationship with southern African rainfall and ENSO. Clim. Dyn., 52, 1783–1803. https://doi.org/10.1007/s00382-018-4222-3.CrossRefGoogle Scholar
da Anunciação, Y. M. T., et al. (2014). Observed summer weather regimes and associated extreme precipitation over Distrito Federal, west-central Brazil. Environ. Earth Sci., 72, 4835–4848. https://doi.org/10.1007/s12665-014-3607-9.Google Scholar
De Kock, W. M., et al. (2022). Large-scale mechanisms linked to anomalously wet summers over the southwestern Cape, South Africa. Clim. Dyn., 59, 3503–3517. https://doi.org/10.1007/s00382-022-06280-7.CrossRefGoogle Scholar
De Luca, P., et al. (2019). Past and projected weather pattern persistence with associated multi-hazards in the British Isles. Atmos., 10(10), 577. https://doi.org/10.3390/atmos10100577.Google Scholar
Delaygue, G., et al. (2019). Reconstruction of Lamb weather type series back to the eighteenth century. Clim. Dyn., 52, 6131–6148. https://doi.org/10.1007/s00382-018-4506-7.CrossRefGoogle Scholar
Desbiolles, F., et al. (2020). Role of ocean mesoscale structures in shaping the Angola-Low pressure system and the southern Africa rainfall. Clim. Dyn., 54, 3685–3704. https://doi.org/10.1007/s00382-020-05199-1.CrossRefGoogle Scholar
Diab, R. D., et al. (1991). Distribution of rainfall by synoptic type over Natal, South Africa. Int. J. Climatol., 11, 877–888.CrossRefGoogle Scholar
Dittmann, A., et al. (2016). Precipitation regime and stable isotopes at Dome Fuji, East Antarctica. Atmos. Chem. Phys., 16, 6883–6900. https://doi.org/10.5194/acp-16-6883-2016.CrossRefGoogle Scholar
Domensino, B. (2010). Using synoptic climate pattern typing to resolve snow accumulation and glaciochemical variability at Mill Island, East Antarctica, Department of Environment and Geography, Macquarie University, Honours thesis, 2010.Google Scholar
Dutheil, C., et al. (2021). Fine-scale rainfall over New Caledonia under climate change. Clim. Dyn., 56, 87–108. https://doi.org/10.1007/s00382-020-05467-0.CrossRefGoogle Scholar
El Kadi, A. K. A. and Smithson, P. A. (1992). Atmospheric classifications and synoptic climatology. Prog. Physic. Geogr., 16(4), 432–455.Google Scholar
Emanuelsson, B. D., et al. (2018). The role of Amundsen-Bellingshausen Sea anticyclonic circulation in forcing marine air mass intrusions into West Antarctica. Clim. Dyn., 51, 3579–3596. https://doi.org/10.1007/s00382-018-4097-3.CrossRefGoogle Scholar
Engelbrecht, C. J., et al. (2015). A synoptic decomposition of rainfall over the Cape south coast of South Africa. Clim Dyn., 44, 2589–2607. https://doi.org/10.1007/s00382-014-2230-5.CrossRefGoogle Scholar
Engelbrecht, C. J. and Landman, W. A. (2016). Interannual variability of seasonal rainfall over the Cape south coast of South Africa and synoptic type association. Clim. Dyn., 47, 295–313. https://doi.org/10.1007/s00382-015-2836-2.CrossRefGoogle Scholar
Espinoza, J. C., et al. (2020). Hydroclimate of the Andes Part I: Main climatic features. Front. Earth Sci., 8, 64. https://doi.org/10.3389/feart.2020.00064.CrossRefGoogle Scholar
Fauchereau, N., et al. (2003). Sea-surface temperature co-variability in the Southern Atlantic and Indian oceans and its connections with the atmospheric circulation in the Southern Hemisphere. Int. J. Climatol., 23, 663–677. https://doi.org/10.1002/joc.905.CrossRefGoogle Scholar
Fauchereau, N., et al. (2016). Extratropical impacts of the Madden–Julian oscillation over New Zealand from a weather regime perspective. J. Clim., 29(6), 2161–2175.CrossRefGoogle Scholar
Favre, A., et al. (2013). Cut-off lows in the South Africa region and their contribution to precipitation. Clim. Dyn., 41, 2331–2351. https://doi.org/10.1007/s00382-012-1579-6.CrossRefGoogle Scholar
Feistel, R., et al. (2003). Climatic changes in the subtropical Southeast Atlantic: The St. Helena Island Climate Index (1893–1999). Prog. Oceanogr., 59, 321–337.CrossRefGoogle Scholar
Fernández-Granja, J. A., et al. (2023). Exploring the limits of the Jenkinson–Collison weather types classification scheme: A global assessment based on various reanalyses. Clim. Dyn. https://doi.org/10.1007/s00382-022-06658-7.CrossRefGoogle Scholar
Flores-Aqueveque, V., et al. (2020). South Pacific Subtropical High from the late Holocene to the end of the 21st century: Insights from climate proxies and general circulation models. Clim. Past, 16, 79–99. https://doi.org/10.5194/cp-16-79-2020.CrossRefGoogle Scholar
Fogt, R. L., et al. (2012). The characteristic variability and connection to the underlying synoptic activity of the Amundsen-Bellingshausen Seas Low. J. Geophys. Res., 117, D07111. https://doi.org/10.1029/2011JD017337.Google Scholar
Fonseca, R. and Martin-Torres, J. M. (2019). High-resolution dynamical downscaling of re-analysis data over the Kerguelen Islands using the WRF model. Theor. Appl. Climatol., 135, 1259–1277. https://doi.org/10.1007/s00704-018-2438-0.CrossRefGoogle Scholar
Francis, D., et al. (2020). On the crucial role of atmospheric rivers in the two major Weddell Polynya events in 1973 and 2017 in Antarctica. Sci. Adv., 6, eabc2695. https://doi.org/10.1126/sciadv.abc2695.CrossRefGoogle ScholarPubMed
Garreaud, R. D. (2009). The Andes climate and weather. Adv. Geosci., 22, 3–11. https://doi.org/10.5194/adgeo-22-3-2009.CrossRefGoogle Scholar
Garreaud, R. D. (2018). A plausible atmospheric trigger for the 2017 coastal El Niño. Int. J. Climatol. 38 (Suppl.1): 1296–1302.CrossRefGoogle Scholar
Garreaud, R. D., et al. (2009). Present-day South American climate. Palaeogeogr., Palaeoclimatol., Palaeoecol., 281, 180–195.CrossRefGoogle Scholar
Gatebe, C. K., et al. (1999). A seasonal air transport climatology for Kenya. J. Geophys. Res., 104(D12), 14237–14244.Google Scholar
Geirinhas, J. L., et al. (2017). Climatic and synoptic characterization of heat waves in Brazil. Int. J. Climatol., 38(4), 1760–1776. https://doi.org/10.1002/joc.5294.Google Scholar
Gibson, P. B., et al. (2017). On the use of self-organizing maps for studying climate extremes. J. Geophys. Res. Atmos., 122, 3891–3903. https://doi.org/10.1002/2016JD026256.CrossRefGoogle Scholar
Gillett, Z. E., et al. (2023). Linking ENSO to synoptic weather systems in eastern Australia. Geophys. Res. Lett., 50, e2023GL104814. https://doi.org/10.1029/2023GL104814.CrossRefGoogle Scholar
Gilliard, J. M. and Keim, B. D. (2018). Position of the South Atlantic anticyclone and its impact on surface conditions across Brazil. J. Appl. Meteorol. Climatol., 57, 535–553. https://doi.org/10.1175/JAMC-D-17-0178.1.Google Scholar
Gong, D. and Wang, S. (1999). The definition of the Antarctic oscillation index. Geophys. Res. Lett., 26(4), 459–462.CrossRefGoogle Scholar
Gonzalez, S., et al. (2018). Atmospheric patterns over the Antarctic Peninsula. J. Clim., 31, 3597–3608. https://doi.org/10.1175/JCLI-D-17-0598.1.CrossRefGoogle Scholar
Goodwin, I. D. (2005). A mid-shelf wave direction climatology for south-eastern Australia, and its relationship to the El Nino – Southern Oscillation, since 1877 AD. Int. J. Climatol., 25, 1715–1729.Google Scholar
Goodwin, I. D., et al. (2003). Snow accumulation variability in Wilkes Land, East Antarctica, and the relationship to atmospheric ridging in the 130°–170°E region since 1930. J. Geophys. Res., 108(D21), 4673. https://doi.org/10.1029/2002JD002995.Google Scholar
Goodwin, I. D., et al. (2004). Mid latitude winter climate variability in the South Indian and southwest Pacific regions since 1300 AD. Clim. Dyn., 22, 783–794.CrossRefGoogle Scholar
Goodwin, I. D., et al. (2016). Tropical and extratropical-origin storm wave types and their influence on the East Australian longshore sand transport system under a changing climate. J. Geophys. Res. Oceans, 121, 4833–4853.CrossRefGoogle Scholar
Gorodetskaya, I. V., et al. (2014). The role of atmospheric rivers in anomalous snow accumulation in East Antarctic. Geophys. Res. Lett., 41, 6199–6206. https://doi.org/10.1002/2014GL060881.CrossRefGoogle Scholar
Grams, C. M., et al. (2017). Balancing Europe’s wind-power output through spatial deployment informed by weather regimes. Nature Clim. Change, 7, 557–562. https://doi.org/10.1038/NCLIMATE3338.CrossRefGoogle ScholarPubMed
Hagen, E. and Agenbag, J. J. (2018). Namibian rainfall and the 1933/34 Benguela Niño. Metorologische Zeitschrift, 27(2), 125–134. https://doi.org/10.1127/metz/2018/0745.Google Scholar
Hanna, E., et al. (2016). Greenland Blocking Index 1851–2015: A regional climate change signal. Int. J. Climatol., 36, 4847–4861.CrossRefGoogle Scholar
Harrison, M. S. J. (1984). A generalized classification of South African rain-bearing synoptic systems. J. Climatol., 4, 547–560. https://doi.org/10.1002/joc.3370040510.CrossRefGoogle Scholar
Hart, N. C. G., et al. (2013). Cloud bands over southern Africa: Seasonality, contribution to rainfall variability and modulation by the MJO. Clim. Dyn., 42, 1199–1212. https://doi.org/10.1007/s00382-012-1589-4.Google Scholar
Hastenrath, S. (1984). The Glaciers of Equatorial East Africa. Reidel, Dordrecht, Boston, Lancaster, p. 353.CrossRefGoogle Scholar
Hertig, E. and Jacobeit, J. (2014). Variability of weather regimes in the North Atlantic-European area: Past and future. Atmos. Sci. Lett., 15, 314–320. https://doi.org/10.1002/asl2.505.CrossRefGoogle Scholar
Hess, P. and Brezowsky, H. (1969). Katalogder Grosswetterlagen Europas, Bd. 15.2. Neubearbeitete und Ergänzte Aufl. Berichte des Deutschen Wetterdienstes 113. Deutscher Wetterdienst: Offenbach am Main, Germany.Google Scholar
Hewitson, B. C. and Crane, R. G. (2002). Self-organizing maps: Applications to synoptic climatology. Clim. Res., 22, 13–26.CrossRefGoogle Scholar
Hildebrandsson, H. H. (1897). Quelque recherches sure les centres d’action de l’atmosphère. K. Sven. Vetenskaps akad. Handl. 29, 1–33.Google Scholar
Hirasawa, N., et al. (2013). The role of synoptic-scale features and advection in prolonged warming and generation of different forms of precipitation at Dome Fuji station, Antarctica, following a prominent blocking event. J. Geophys. Res. Atmos., 118(13), 6916–6928. https://doi.org/10.1002/jgrd.50532.CrossRefGoogle Scholar
Hope, P. K., (2006). Projected future changes in synoptic systems influencing southwest Western Australia. Clim. Dyn., 26, 765–780. https://doi.org/10.1007/s00382-006-0116-x.Google Scholar
Hope, P. K., et al. (2006). Shifts in the synoptic systems influencing southwest Western Australia. Clim. Dyn., 26, 751–764.Google Scholar
Hosking, J. S., et al. (2013). The influence of the Amundsen–Bellingshausen Seas Low on the climate of West Antarctica and its representation in coupled climate model simulations. J. Clim., 26, 6633–6648.CrossRefGoogle Scholar
Howard, E. and Washington, R. (2018). Characterizing the synoptic expression of the Angola Low. J. Clim., 31, 7147–7165. https://doi.org/10.1175/JCLI-D-18-0017.1.CrossRefGoogle Scholar
Ibebuchi, C. C. (2021). Can synoptic patterns influence the track and formation of tropical cyclones in the Mozambique Channel? AIMS Geosci., 8(1), 33–51.Google Scholar
Ibebuchi, C. C. and Lee, C. C. (2024). Circulation pattern controls of summer temperature anomalies in southern Africa. Adv. Atmos. Sci., 41(2), 341−354. https://doi.org/10.1007/s00376-023-2392-3.CrossRefGoogle Scholar
Illig, S. and Bachelery, M.-L. (2019). Propagation of subseasonal equatorially-forced coastal trapped waves down to the Benguela Upwelling system. Sci. Rep., 9, 5306. https://doi.org/10.1038/s41598-019-41847-1.CrossRefGoogle Scholar
Indian Ocean Climate Initiative (IOCI). (2012). Western Australia’s Weather and Climate: A Synthesis of Indian Ocean Climate Initiative Stage 3 Research. CSIRO and BoM, Australia.Google Scholar
Jenkinson, A. F. and Collison, F. P. (1977). An Initial Climatology of Gales over the North Sea, Synoptic Climatology Branch Memorandum No. 62. Meteorological Office, Bracknell.Google Scholar
Jiang, N. (2011). A new objective procedure for classifying New Zealand synoptic weather types during 1958–2008. Int. J. Climatol., 31, 863–879.CrossRefGoogle Scholar
Jiang, N., et al. (2012a). On two different objective procedures for classifying synoptic weather types over east Australia. Int. J. Climatol., 32, 1475–1494.CrossRefGoogle Scholar
Jiang, N., et al. (2012b). Influence of large-scale climate modes on daily synoptic weather types over New Zealand. Int. J. Climatol., 33(2), 499–519.Google Scholar
Jones, P. D., et al. (1993). A comparison of Lamb circulation types with an objective classification scheme. Int. J. Climatol., 13, 655–663.CrossRefGoogle Scholar
Jones, P. D., et al. (2013). Lamb weather types derived from reanalysis products. Int. J. Climatol., 33, 1129–1139. https://doi.org/10.1002/joc.3498.CrossRefGoogle Scholar
Jones, P. D., et al. (2014). The development of Lamb weather types: From subjective analysis of weather charts to objective approaches using reanalyses. Weather, 69(5), 128–132.CrossRefGoogle Scholar
Jones, P. D., et al. (2015). Long-term trends in gale days and storminess for the Falkland Islands. Int. J. Climatol., 36, 1413–1427 (2016). https://doi.org/10.1002/joc.4434.Google Scholar
Jury, M. R., et al. (1990). Pulsing of the Benguela upwelling region: Large-scale atmospheric controls, S. Afr. J. Mar. Sci., 9(1), 27–41. https://doi.org/10.2989/025776190784378826.CrossRefGoogle Scholar
Kageyama, M., et al. (1999). Weather regimes in past climate atmospheric general circulation model simulations. Clim. Dyn., 15, 773–793.CrossRefGoogle Scholar
Kataoka, T., et al. (2014). On the Ningaloo Niño/Niña. Clim. Dyn., 43(5–6), 1463–1482. https://doi.org/10.1007/s00382-013-1961-z.CrossRefGoogle Scholar
Kataoka, T., et al. (2018). Can Ningaloo Niño/Niña develop without El Niño–Southern Oscillation? Geophys. Res. Lett., 45, 7040–7048. https://doi.org/10.1029/2018GL078188.CrossRefGoogle Scholar
Kidson, J. W. (2000). An analysis of New Zealand synoptic types and their use in defining weather regimes. Int. J. Climatol., 20, 299–316.3.0.CO;2-B>CrossRefGoogle Scholar
Lamb, H. H. (1950). Types and spells of weather around the year in the British Isles. Q. J. R. Meteorol. Soc., 76, 393–438.CrossRefGoogle Scholar
Lamb, H. H. (1972). British Isles Weather Types and a Register of Daily Sequence of Circulation Patterns, 1861–1971. Geophysical Memoir 116. HMSO, London, 85pp.Google Scholar
Lee, J.-W., et al. (2019). Weather noise leading to El Niño diversity in an ocean general circulation model. Clim. Dyn., 52, 7235–7247. https://doi.org/10.1007/s00382-016-3438-3.CrossRefGoogle Scholar
Lefèvre, J., et al. (2010). Weather regimes and orographic circulation around New Caledonia. Mar. Pollut. Bull., 61(7), 413–431.CrossRefGoogle ScholarPubMed
Lennard, C. and Hegerl, G. (2015). Relating changes in synoptic circulation to the surface rainfall response using self-organizing maps. Clim. Dyn., 44, 861–879. https://doi.org/10.1007/s00382-014-2169-6.CrossRefGoogle Scholar
Levick, R. B. M. (1949). Fifty years of British weather. Weather, 4, 206–211.CrossRefGoogle Scholar
Levick, R. B. M. (1950). Fifty years of British weather. Weather, 5, 245–247.CrossRefGoogle Scholar
Liu, W., et al. (2020). Influence of Indian Ocean SST regionality on the East African short rains. Clim. Dyn., 54, 4991–5011. https://doi.org/10.1007/s00382-020-05265-8.CrossRefGoogle Scholar
Loikith, P. C., et al. (2019). A climatology of daily synoptic circulation patterns and associated surface meteorology over southern South America. Clim. Dyn., 53, 4019–4035. https://doi.org/10.1007/s00382-019-04768-3.CrossRefGoogle Scholar
Lorrey, A., et al. (2007). Regional climate regime classification as a qualitative tool for interpreting multi-proxy palaeoclimate data spatial patterns: A New Zealand case study. Palaeogeogr. Palaeoclimatol. Palaeoecol., 253(3), 407–433.CrossRefGoogle Scholar
Lorrey, A., et al. (2008). Speleothem stable isotope records interpreted within a multi-proxy framework and implications for New Zealand palaeoclimate reconstruction. Quat. Int., 187, 52–75.CrossRefGoogle Scholar
Lorrey, A. M. and Fauchereau, N. C. (2017). Southwest Pacific atmospheric weather regimes: Linkages to ENSO and extratropical teleconnections. Int. J. Climatol., 38(4), 1893–1909.Google Scholar
Lübbecke, J. F., et al. (2014). Variability in the South Atlantic anticyclone and the Atlantic Niño Mode. J. Clim., 27, 8135–8150. https://doi.org/10.1175/JCLI-D-14-00202.1.CrossRefGoogle Scholar
Lübbecke, J. F. and McPhaden, M. J. (2017). Symmetry of the Atlantic Niño mode. Geophys. Res. Lett., 44, 965–973, https://doi.org/10.1002/2016GL071829.CrossRefGoogle Scholar
Lübbecke, J. F., et al. (2018). Equatorial Atlantic variability – Modes, mechanisms, and global teleconnections. WIREs Clim. Change, 9, e527. https://doi.org/10.1002/wcc.527.CrossRefGoogle Scholar
Manatsa, D., et al. (2016). Linking major shifts in East Africa ‘short rains’ to the Southern Annular Mode. Int. J. Climatol., 36, 1590–1599. https://doi.org/10.1002/joc.4443.CrossRefGoogle Scholar
Marchant, R., et al. (2006). The Indian Ocean dipole – The unsung driver of climatic variability in East Africa. Afr. J. Ecol., 45, 4–16.Google Scholar
Markle, B. R., et al. (2012). Synoptic variability in the Ross Sea region, Antarctica, as seen from back-trajectory modeling and ice core analysis. J. Geophys. Res., 117, D02113. https://doi.org/10.1029/2011JD016437.Google Scholar
Massom, R. A., et al. (2004). Precipitation over the interior East Antarctic Ice Sheet related to midlatitude blocking-high activity. J. Clim., 17, 1914–1928.2.0.CO;2>CrossRefGoogle Scholar
Mawren, D., and Reason, C. J. C. (2017). Variability of upper-ocean characteristics and tropical cyclones in the South West Indian Ocean. J. Geophys. Res. Oceans, 122. https://doi.org/10.1002/2016JC012028.CrossRefGoogle Scholar
Michelangeli, P.-A., et al. (1995). Weather regimes: Recurrence and quasi stationarity. J. Atmos. Sci., 52(8), 1237–1256.2.0.CO;2>CrossRefGoogle Scholar
Morioka, Y., et al. (2013). How is the Indian Ocean Subtropical Dipole excited? Clim. Dyn., 41, 1955–1968. https://doi.org/10.1007/s00382-012-1584-9.CrossRefGoogle Scholar
Morioka, Y., et al. (2014). Role of tropical SST variability on the formation of subtropical dipoles. J. Clim., 27, 4486–4507. https://doi.org/10.1175/JCLI-D-13-00506.1.Google Scholar
Moron, V., et al. (2015). Weather types across the Maritime Continent: From the diurnal cycle of interannual variations. Front. Environ. Sci., 2, 65. https://doi.org/10.3389/fenvs.2014.00065.CrossRefGoogle Scholar
Moron, V., et al. (2019). Weather types and hourly to multiday rainfall characteristics in Tropical Australia. J. Clim., 32, 3983–4011. https://doi.org/10.1175/JCLI-D-18-0384.1.CrossRefGoogle Scholar
Mortlock, T. R. and Goodwin, I. D. (2015). Directional wave climate and power variability in the Tasman Sea. Cont. Shelf Res., 98, 36–53.CrossRefGoogle Scholar
Mossman, R. C. (1913). Southern Hemisphere seasonal correlations. Symon’s Meteorol. Mag., 48, 1–34.Google Scholar
Mulenga, H. M., et al. (2003). Dry summers over northeastern South Africa and associated circulation anomalies. Clim. Res., 25, 29–41. https://doi.org/10.3354/cr025029.CrossRefGoogle Scholar
Muller, A., et al. (2008). Extreme rainfall in the Namib desert during late summer 2006 and influences of regional ocean variability. Int. J. Climatol., 28(8), 1061–1070. https://doi.org/10.1002/joc.1603.CrossRefGoogle Scholar
Munday, C. and Washington, R. (2017). Circulation controls on southern African precipitation in coupled models: The role of the Angola Low. J. Geophys. Res. Atmos., 122, 861–877. https://doi.org/10.1002/2016JD025736.CrossRefGoogle Scholar
Nicholson, S. E. (2010). A low-level jet along the Benguela coast, an integral part of the Benguela current ecosystem. Clim. Change., 99(3–4), 613–624.CrossRefGoogle Scholar
Nnamchi, H. C., et al. (2021). Diabatic heating governs the seasonality of the Atlantic Niño. Nature Comm., 12(1), 376. https://doi.org/10.1038/s41467-020-20452-1.CrossRefGoogle ScholarPubMed
Noone, D., et al. (1999). Atmospheric signals and characteristics of accumulation in Dronning Maud Land, Antarctica. J. Geophys. Res., 104(D16), 19191–19211.Google Scholar
O Kane, T. J., et al. (2017). A multiscale reexamination of the Pacific–South American pattern. Mon. Wea. Rev., 145, 379–402. https://doi.org/10.1175/MWR-D-16-0291.1.Google Scholar
O’ Hare, G. and Sweeney, J. (1993). Lamb’s circulation types and British Weather: An evaluation. Geography, January 1993, 78(1), 43–60.Google Scholar
Odoulami, R. C., et al. (2021). A SOM-based analysis of the drivers of the 2015–2017 Western Cape drought in South Africa. Int. J. Climatol., 41(S1), E1518–1530. https://doi.org/10.1002/joc.6785.CrossRefGoogle Scholar
Ohishi, S., et al. (2015). Zonal movement of the Mascarene High in austral summer. Clim. Dyn., 45, 1739–1745. https://doi.org/10.1007/s00382-014-2427-7.CrossRefGoogle Scholar
Pascale, S., et al. (2019). On the Angola Low interannual variability and its role in modulating ENSO effects in Southern Africa. J. Clim., 32, 4783–4803. https://doi.org/10.1175/JCLI-D-18-0745.1.CrossRefGoogle Scholar
Pepler, A., et al. (2019). A global climatology of surface anticyclones, their variability, associated drivers and long- term trends. Clim. Dyn., 52(9–10), 5397–5412. https://doi.org/10.1007/s00382-018-4451-5.CrossRefGoogle Scholar
Pepler, A. S., et al. (2020). The contributions of fronts, lows and thunderstorms to southern Australian rainfall. Clim. Dyn., 55, 1489–1505. https://doi.org/10.1007/s00382-020-05338-8.CrossRefGoogle Scholar
Petit, J. R., et al. (1991). Deuterium excess in recent Antarctic snow. J. Geophys. Res., 96, 5113–5122. https://doi.org/10.1029/90JD02232.Google Scholar
Pohl, B. and Fauchereau, N. (2012). The Southern Annular Mode seen through weather regimes. J. Clim., 25, 3336–3354.CrossRefGoogle Scholar
Pohl, B., et al. (2018). From synoptic to interdecadal variability in Southern African rainfall: Toward a unified view across time scales. J. Clim., 31, 5845–5872. https://doi.org/10.1175/JCLI-D-17-0405.1.CrossRefGoogle Scholar
Pohl, B., et al. (2021). Relationship between weather regimes and atmospheric rivers in East Antarctica. J. Geophys. Res.: Atmos, 126, e2021JD035294. https://doi.org/10.1029/2021JD035294.CrossRefGoogle Scholar
Pope, M., et al. (2009). Regimes of the North Australian wet season. J. Clim., 22, 6699–6715. https://doi.org/10.1175/2009JCLI3057.1.CrossRefGoogle Scholar
Poveda, G., et al. (2020). High impact weather events in the Andes. Front. Earth Sci. 8, 162. https://doi.org/10.3389/feart.2020.00162.CrossRefGoogle Scholar
Prabhat, , et al. (2021). ClimateNet: An expert-labeled open dataset and Deep Learning architecture for enabling high-precision analyses of extreme weather. Geosci. Model Dev., 14(1), 107–124. https://doi.org/10.5194/gmd-2020-72.CrossRefGoogle Scholar
Puard, Y., et al. (2017). Climate covariability between South America and Southern Africa at interannual, intraseasonal and synoptic scales. Clim. Dyn., 48, 4029–4050. https://doi.org/10.1007/s00382-016-3318-x.Google Scholar
Ramos, A. M., et al. (eds.) (2015). Circulation weather types as a tool in atmospheric, climate and environmental research. Front. Environ. Sci., 3, 44. https://doi.org/10.3389/fenvs.2015.00044.CrossRefGoogle Scholar
Reason, C. J. C. (2001). Subtropical Indian Ocean SST dipole events and southern African rainfall. Geophys. Res. Lett., 28, 2225–2227. https://doi.org/10.1029/2000GL012735.CrossRefGoogle Scholar
Reason, C. J. C. (2017). Climate of Southern Africa. Oxford University Press, Oxford. https://doi.org/10.1093/acrefore/9780190228620.013.513.Google Scholar
Reboita, M. S., et al. (2019). The South Atlantic Subtropical anticyclone: Present and future climate. Front. Earth Sci., 7, 8. https://doi.org/10.3389/feart.2019.00008.CrossRefGoogle Scholar
Rodríguez-Morata, C., et al. (2018). Linking atmospheric circulation patterns with hydro-geomorphic disasters in Peru. Int. J. Climatol., 38, 3388–3404.CrossRefGoogle Scholar
Rouault, M., et al. (2005). Climate variability at Marion Island, Southern Ocean, since 1960, J. Geophys. Res., 110, C05007.Google Scholar
Rueda, A., et al. (2018). Marine climate variability based on weather patterns for a complicated island setting: The New Zealand case. Int. J. Climatol., 39, 1777–1786. https://doi.org/10.1002/joc.5912.Google Scholar
Russell, A., et al. (2004). An examination of the precipitation delivery mechanisms for Dolleman Island, eastern Antarctic Peninsula. Tellus, 56A, 501–513.Google Scholar
Russell, A., et al. (2008). Eastern Antarctic Peninsula precipitation delivery mechanisms: Process studies and back trajectory evaluation. Atmos. Sci. Lett., 9(4), 214–221. https://doi.org/10.1002/asl.190.CrossRefGoogle Scholar
Santos, E. B., et al. (2017). Synoptic patterns of atmospheric circulation associated with intense precipitation events over the Brazilian Amazon. Theor. Appl. Climatol., 128, 343–358. https://doi.org/10.1007/s00704-015-1712-7.CrossRefGoogle Scholar
Scarchilli, C., et al. (2011). Snow precipitation at four ice core sites in East Antarctica: Provenance, seasonality and blocking factors. Clim. Dyn., 37, 2107–2125. https://doi.org/10.1007/s00382-010-0946-4.CrossRefGoogle Scholar
Schlosser, E., et al. (2011). Interaction between Antarctic sea ice and synoptic activity in the circumpolar trough: Implications for ice-core interpretation. Ann. Glaciol., 52(57), 9–17.CrossRefGoogle Scholar
Seefeldt, M. W. and Cassano, J. J. (2012). A description of the Ross Ice Shelf air stream (RAS) through the use of self-organizing maps (SOMs), J. Geophys. Res., 117, D09112. https://doi.org/10.1029/2011JD016857.Google Scholar
Shannon, L. V., et al. (1986). On the existence of an El Niño-type phenomenon in the Benguela system. J. Mar. Res., 44(3), 495–520.CrossRefGoogle Scholar
Sierra, J. P., et al. (2021). The Choco low-level jet: Past, present and future. Clim. Dyn., 56(7–8), 2667–2692. https://doi.org/10.1007/s00382-020-05611-w.CrossRefGoogle Scholar
Sinclair, M. R. (1981). Record-high temperatures in the Antarctic – A synoptic case study. Mon. Weather Rev., 109, 2234–2242. https://doi.org/10.1175/1520-0493(1981)109<2234:rhtita>2.0.co;2.2.0.CO;2>CrossRefGoogle Scholar
Sinclair, K. E., et al. (2013). Seasonality of airmass pathways to coastal Antarctica: Ramifications for interpreting high-resolution ice core records. J. Clim., 26, 2065–2076. https://doi.org/10.1175/JCLI-D-12-00167.1.CrossRefGoogle Scholar
Singleton, A. T. and Reason, C. J. C. (2007). Variability in the characteristics of cut-off low pressure systems over subtropical Southern Africa. Int. J. Climatol., 27, 295–310. https://doi.org/10.1002/joc.1399.CrossRefGoogle Scholar
Sodemann, H. and Stohl, A. (2009). Asymmetries in the moisture origin of Antarctic precipitation. Geophys. Res. Lett., 36, L22803. https://doi.org/10.1029/2009GL040242.CrossRefGoogle Scholar
Solman, S. A. and Menendez, C. G. (2003). Weather regimes in the South American sector and neighbouring oceans during winter. Clim. Dyn., 21, 91–104. https://doi.org/10.1007/s00382-003-0320-x.CrossRefGoogle Scholar
Sousa, P. M., et al. (2018). The ‘Day Zero’ Cape Town drought and the poleward migration of moisture corridors. Environ. Res. Lett., 13(12), 124025.CrossRefGoogle Scholar
Souverijns, N., et al. (2016). Drivers of future changes in East African precipitation, Environ. Res. Lett., 11, 114011. https://doi.org/10.1088/1748-9326/11/11/114011.CrossRefGoogle Scholar
Sulca, J., et al. (2018). Impacts of different ENSO flavors and tropical Pacific convection variability (ITCZ, SPCZ) on austral summer rainfall in South America, with a focus on Peru. Int. J. Climatol., 38, 420–435.CrossRefGoogle Scholar
Sun, X., et al. (2017). The South Atlantic subtropical high: Climatology and interannual variability. J. Clim., 30, 3279–3296. https://doi.org/10.1175/JCLI-D-16-0705.1.CrossRefGoogle Scholar
Taljaard, J. J. (1995). Atmospheric circulation systems, synoptic climatology and weather phenomena of South Africa. Part 2: Atmospheric circulation systems in the South African region. South African Weather Bureau, Technical paper 28.Google Scholar
Todd, M. and Washington, R. (1999). Circulation anomalies associated with tropical-temperate troughs in Southern Africa and the south west Indian Ocean. Clim. Dyn., 15, 937–951. https://doi.org/10.1007/s003820050323.CrossRefGoogle Scholar
Tozer, C. R., et al. (2018). The relationship between wave trains in the Southern Hemisphere storm track and rainfall extremes over Tasmania. Mon. Weather Rev., 146, 4201–4230. https://doi.org/10.1175/MWR-D-18-0135.1.CrossRefGoogle Scholar
Tozuka, T., et al. (2008). Tropical Indian Ocean variability revealed by self-organizing maps. Clim. Dyn., 31, 333–343.CrossRefGoogle Scholar
Turner, J., et al. (2013). The Amundsen Sea low. Int. J. Climatol., 33, 1818–1829. https://doi.org/10.1002/joc.3558.CrossRefGoogle Scholar
Turner, J., et al. (2019). The dominant role of extreme precipitation events in Antarctic snowfall variability. Geophys. Res. Lett., 46, 3502–3511.CrossRefGoogle Scholar
Turner, J., et al. (2021). Extreme temperatures in the Antarctic. J. Clim., 34, 2653–2668. https://doi.org/10.1175/jcli-d-20-0538.1.Google Scholar
Turner, J., et al. (2022). An extreme high temperature event in coastal East Antarctica associated with an atmospheric river and record summer downslope winds. Geophys. Res. Lett., 49, e2021GL097108. https://doi.org/10.1029/2021GL097108.CrossRefGoogle Scholar
Tyson, P. D. and Preston-Whyte, R. A. (2000). The Weather and Climate of Southern Africa. Oxford University Press, South Africa.Google Scholar
Udy, D. G., et al. (2021). Links between large-scale modes of climate variability and synoptic weather patterns in the southern Indian Ocean. J. Clim., 34, 883–899.CrossRefGoogle Scholar
Ummenhofer, C. C., et al. (2009). What causes south- east Australia’s worst droughts? Geophys. Res. Lett., 36, L04706.CrossRefGoogle Scholar
Vallès-Casanova, I., et al. (2020). On the spatiotemporal diversity of Atlantic Niño and associated rainfall variability over West Africa and South America. Geophys. Res. Lett., 47, e2020GL087108. https://doi.org/10.1029/2020GL087108.CrossRefGoogle Scholar
Verdon-Kidd, D. C. and Kiem, A. S. (2009). On the relationship between large-scale climate modes and regional synoptic patterns that drive Victorian rainfall. Hydrol. Earth Syst. Sci., 13, 467–479. https://doi.org/10.5194/hess-13-467-2009.CrossRefGoogle Scholar
Weidemann, S. S., et al. (2018). A 17-year record of meteorological observations across the Gran Campo Nevado Ice Cap in Southern Patagonia, Chile, pelated to synoptic weather types and climate modes. Front. Earth Sci., 6, 53. https://doi.org/10.3389/feart.2018.00053.CrossRefGoogle Scholar
Wheeler, M. C. and Hendon, H. H. (2004). An all-season real- time multivariate MJO index: Development of an index for monitoring and prediction. Mon. Weather. Rev., 132, 1917–1932. https://doi.org/10.1175/1520-0493(2004)132,1917:AARMMI.2.0.CO;2.2.0.CO;2>CrossRefGoogle Scholar
Widlansky, M. J., et al. (2011). On the location and orientation of the South Pacific convergence zone. Clim. Dyn., 36, 561–578. https://doi.org/10.1007/s00382-010-0871-6.CrossRefGoogle Scholar
Willie, J. D., et al. (2019). West Antarctic surface melt triggered by atmospheric rivers. Nature Geo., 12, 911–916. https://doi.org/10.1038/s41561-019-0460-1.Google Scholar
Wilson, L., et al. (2013). Relationship between rainfall and weather regimes in southeastern Queensland, Australia. Int. J. Climatol., 33, 979–991. https://doi.org/10.1002/joc.3484.CrossRefGoogle Scholar
Xulu, N., et al. (2020). Climatology of the Mascarene High and its influence on weather and climate over Southern Africa. Climate, 8, 86. https://doi.org/10.3390/cli8070086.CrossRefGoogle Scholar
Yarnal, B. (1993). Synoptic Climatology in Environmental Analysis: A Primer. Belhaven Press, London, 195.Google Scholar
Yarnal, B., et al. (2001). Review: Developments and prospects in synoptic climatology. Int. J. Climatol., 21, 1923–1950.CrossRefGoogle Scholar
Zhang, L. and Han, W. (2021). Indian Ocean Dipole leads to Atlantic Niño. Nature Comm., 12, 5952. https://doi.org/10.1038/s41467-021-26223-w.Google ScholarPubMed
Acworth, C. and Lawson, S. (2012). The Tweed Rive Entrance Sand Bypassing Project, ten years of managing operations in a highly variable coastal system. Proc. 20th NSW Coastal Conference, Tweed Heads, pp. 1–23.Google Scholar
Ainsworth, R. B., et al. (2011). Dynamic spatial and temporal prediction of changes in depositional processes on clastic shorelines: Toward improved subsurface uncertainty reduction and management. Am. Assoc. Pet. Geol. Bull., 95(2), 267–297.Google Scholar
Allard, J., et al. (2008). Sand spit rhythmic development: A potential record of wave climate variations? Arçay spit, western coast of France. Mar. Geol., 253, 107–131.CrossRefGoogle Scholar
Almar, R., et al. (2015). Response of the Bight of Benin (Gulf of Guinea, West Africa) coastline to anthropogenic and natural forcing, Part1: Wave climate variability and impacts on the longshore sediment transport. Cont. Shelf Res., 110, 48–59.CrossRefGoogle Scholar
Alomar, M., et al. (2014). Wave growth and forecasting in variable, semi enclosed domains. Cont. Shelf Res., 87(15), 28–40.CrossRefGoogle Scholar
Anthony, E. J. (1995). Beach-ridge progradation in response to sediment supply: Examples from West Africa. Mar. Geol., 129, 175–186.CrossRefGoogle Scholar
Anthony, E. J. (2009). Shore Processes and Their Palaeoenvironmental Applications. Dev. Mar. Geol., 4. Elsevier Science, Amsterdam, 519pp.Google Scholar
Anthony, E. J. (2015). Wave influence in the construction, shaping and destruction of river deltas: A review. Mar. Geol., 361, 53–78.CrossRefGoogle Scholar
Anthony, E. J., et al. (2019). Response of the Bight of Benin (Gulf of Guinea, West Africa) coastline to anthropogenic and natural forcing, Part 2: Sources and patterns of sediment supply, sediment cells, and recent shoreline change. Cont. Shelf Res., 173, 93–103.CrossRefGoogle Scholar
Ashton, A. D., et al. (2001). Formation of coastline features by large-scale instabilities induced by high-angle waves. Nature, 414, 296–300. https://doi.org/10.1038/35104541.CrossRefGoogle ScholarPubMed
Ashton, A. D. and Murray, A. B. (2006a). High-angle wave instability and emergent shoreline shapes: 1. Modeling of sand waves, flying spits, and capes. J. Geophys. Res., 111, F04011. https://doi.org/10.1029/2005JF000422.Google Scholar
Ashton, A. D. and Murray, A. B. (2006b). High-angle wave instability and emergent shoreline shapes: 2. Wave climate analysis and comparisons to nature. J. Geophys. Res., 111, F04012. https://doi.org/10.1029/2005JF000423.Google Scholar
Ashton, A. D. and Giosan, L. (2011). Wave-angle control of delta evolution. Geophys. Res. Lett., 38, L13405. https://doi.org/10.1029/2011GL047630.CrossRefGoogle Scholar
Ashton, A. D., et al. (2013). Progress in coupling models of coastline and fluvial dynamics. Comp. Geosci., 53, 21–29.CrossRefGoogle Scholar
Ashton, A. D., et al. (2016). On a neck, on a spit: Controls on the shape of free spits. Earth Surf. Dynam., 4, 193–210. https://doi.org/10.5194/esurf-4-193-2016.CrossRefGoogle Scholar
Bagnold, R. A. (1941). The Physics of Blown Sand and Desert Dunes. Methuen, London, 265pp.Google Scholar
Bárdossy, A., et al. (2015). Circulation patterns identified by spatial rainfall and ocean wave fields in Southern Africa. Front. Environ. Sci., 3, 31. https://doi.org/10.3389/fenvs.2015.00031.CrossRefGoogle Scholar
Bauer, E. (2001). Interannual changes of the ocean wave variability in the North Atlantic and in the North Sea. Clim. Res., 18(1–2), pp. 63–69. Available at: www.jstor.org/stable/24861559.CrossRefGoogle Scholar
Bhattacharya, J. P. and Giosan, L. (2003). Wave-influenced deltas: Geomorphological implications for facies reconstruction. Sedimentol., 50, 187–210.CrossRefGoogle Scholar
Bird, E. C. F. (2000). Coastal Geomorphology: An Introduction. Wiley, Chichester, 322pp.Google Scholar
Boukhanovsky, A. V., et al. (2007). Spectral wave climates of the North Sea. Appl. Ocean Res., 29(3), 146–154.CrossRefGoogle Scholar
Boyd, R., et al. (1992). Classification of clastic coastal depositional environments. Sed. Geol., 80, 139–150.CrossRefGoogle Scholar
Bristow, C. S., et al. (2000). The structure and development of foredunes on a locally prograding coast: Insights from ground-penetrating radar surveys, Norfolk, UK. Sedimentol., 47, 923–944.CrossRefGoogle Scholar
Bromirski, P. D., et al. (2013). Wave power variability and trends across the North Pacific. J Geophys, Res. Oceans., 118(12), 6329–6348. https://doi.org/10.1002/2013JC009189.CrossRefGoogle Scholar
Bromirski, P. D. and Cayan, D. R. (2015). Wave power variability and trends across the North Atlantic influenced by decadal climate patterns. J Geophys, Res. Oceans., 120. https://doi.org/10.1002/2014JC010440.CrossRefGoogle Scholar
Brooke, B., et al. (2008). Influence of climate fluctuations and changes in catchment land use on Late Holocene and modern beach-ridge sedimentation on a tropical macrotidal coast: Keppell Bay, Queensland, Australia. Mar. Geol., 251, 195–208.CrossRefGoogle Scholar
Bryant, E. A. (2001). Tsunami – The Underrated Hazard. Cambridge University Press, 320pp.Google Scholar
Bujalesky, G. G., et al. (2014). Holocene coastal environments and processes in subantarctic/temperate cold Tierra del Fuego, Argentina-Chile. In Martini, I. P. and Wanless, H. R. (eds.), Sedimentary Coastal Zones from High to Low Latitudes: Similarities and Differences. Geological Society, London, Special Publications, 388. http://dx.doi.org/10.1144/SP388.10.Google Scholar
Bullard, J. (1997). A note on the use of the ‘Fryberger Method’ for evaluating potential sand transport by wind. J. Sed. Res., 67, 499–501.CrossRefGoogle Scholar
Buynevich, I. V., et al. (2007). A 1500 yr record of North Atlantic storm activity based on optically dated relict beach scarps. Geol., 35(6), 543–546.CrossRefGoogle Scholar
Caldwell, R. L., et al. (2019). A global delta dataset and the environmental variables that predict delta formation on marine coastlines. Earth Surf. Dynam., 7, 773–787. https://doi.org/10.5194/esurf-7-773-2019.CrossRefGoogle Scholar
Callaghan, J. and Helman, P. (2008). Severe storms on the east coast of Australia 1770–2008. Griffith Centre for Coastal Management Report, Griffith University, 1–252.Google Scholar
Camus, P., et al. (2011a). Multivariate wave climate using self-organizing maps. J. Atmos. Ocean. Technol., 28, 1554–1658.CrossRefGoogle Scholar
Camus, P., et al. (2011b). Analysis of clustering and selection algorithms for the study of multivariate wave climate. Coast. Eng., 58, 453–462.CrossRefGoogle Scholar
Camus, P., et al. (2013). High resolution downscaled ocean waves (DOW) reanalysis in coastal areas. Coast. Eng., 72, 56–68.CrossRefGoogle Scholar
Camus, P., et al. (2014). A weather-type statistical downscaling framework for ocean wave climate. J Geophys, Res. Oceans., 119, 7389–7405. https://doi.org/10.1002/2014JC010141.CrossRefGoogle Scholar
Camus, P., et al. (2016). An atmospheric-to-marine synoptic classification for statistical downscaling marine climate. Ocean Dyn., 66, 1589–1601. https://doi.org/10.1007/s10236-016-1004-5.CrossRefGoogle Scholar
Carter, R. W. G. and Woodroffe, C. D., (eds.), (1994). Coastal Evolution: Late Quaternary Shoreline Morphodynamics. Cambridge University Press, Cambridge, 517pp.Google Scholar
Castelle, B., et al. (2017). Foredune morphological changes and beach recovery from the extreme 2013/2014 winter at a high-energy sandy coast. Mar. Geol., 385, 41–55.CrossRefGoogle Scholar
Chapman, D. M. (1978). Zetaform or logarithmic spiral beach. Australian Geographer, 14, 44–45.CrossRefGoogle Scholar
Claudino-Sales, V., et al. (2018). Interactions between various headlands, beaches, and dunes along the Coast of Ceara State, Northeast Brazil. J. Coast. Res., 34(2), 413–428.Google Scholar
Coco, G. and Murray, A. B. (2007). Patterns in the sand: From forcing templates to self organization. Geomorphol., 91, 271–290.CrossRefGoogle Scholar
Corbella, S., et al. (2015). Assimilation of ocean wave spectra and atmospheric circulation patterns to improve wave modeling. Coast. Eng., 100, 1–10.CrossRefGoogle Scholar
Costa, M. B., et al. (2017). Planimetric and volumetric changes of reef islands in response to wave conditions. Earth Surf. Process. Landforms, 42, 2663–2678. https://doi.org/10.1002/esp.4215.CrossRefGoogle Scholar
Cowell, P. J., et al. (1995). Simulation of large-scale coastal change using a morphological behavior model. Mar. Geol., 126, 45–61.CrossRefGoogle Scholar
Cowell, P. J., et al. (2003a). The coastal-tract (part 1): A conceptual approach to aggregated modeling of low-order coastal change. J. Coast. Res., 19, 812–827.Google Scholar
Cowell, P. J., et al. (2003b). The coastal-tract (part 2): Applications of aggregated modeling of lower-order coastal change. J. Coast. Res., 19, 828–848.Google Scholar
Cuttler, M. V. W. (2020). Interannual response of reef islands to climate-driven variations in water level and wave climate. Remote Sens., 12, 4089. https://doi.org/10.3390/rs12244089.CrossRefGoogle Scholar
Cuttler, M. V. W., et al. (2019). Source and supply of sediment to a shoreline salient in a fringing reef environment. Earth Surf. Process. Landforms, 44, 552–564. https://doi.org/10.1002/esp.4516.CrossRefGoogle Scholar
Daly, C. J., et al. (2014). Wave energy distribution and morphological development in and around the shadow zone of an embayed beach. Coast. Eng., 93, 40–54.CrossRefGoogle Scholar
Davies, J. L. (1958). Wave refraction and the evolution of shoreline curves. Aust. Geogr. Stud., 5, 1–14.Google Scholar
Davies, J. L. (1974). The coastal sediment compartment. Aust. Geogr. Stud., 12, 139–151.CrossRefGoogle Scholar
Davies, J. L. (1980). Geographical Variation in Coastal Development, 2nd ed. Longman, London, 212pp.Google Scholar
Davis, R. A., (ed). (1985). Coastal Sedimentary Environments, 2nd ed. Springer-Verlag, New York, 716pp.CrossRefGoogle Scholar
Deltares. (2024). User manual SWAN Cycle III version 41.45AB. www.swan.tudelft.nl.Google Scholar
Denniston, R. F., et al. (2015). Extreme rainfall activity in the Australian tropics reflects changes in the El Niño/Southern Oscillation over the last two millennia. Proc. Natl. Acad. Sci. USA, 112, 4576–4581.CrossRefGoogle ScholarPubMed
DHI. (2018). MIKE 21 Flow Model FM, Hydrodynamic Module, User Guide. DHI Water & Environment, Denmark.Google Scholar
Dillenburg, S. and Hesp, P. (2009). Geology and geomorphology of Holocene coastal barriers of Brazil. Lecture Notes in Earth Sciences 107, Springer-Verlag, Berlin, Heidelberg, 380pp.CrossRefGoogle Scholar
Dissanayake, P., et al. (2015). Effects of storm clustering on beach/dune evolution. Mar. Geol., 370, 63–75, https://doi.org/10.1016/jmargeo.2015.10.010.CrossRefGoogle Scholar
Dodet, G., et al. (2010). Wave climate variability in the North-East Atlantic Ocean over the last six decades. Ocean Model., 31(3–4), 120–131. https://doi.org/10.1016/j.ocemod.2009.10.010.CrossRefGoogle Scholar
Dominey-Howes, D. (2007). Geological and historical records of tsunami in Australia. Mar. Geol., 239, 99–123.CrossRefGoogle Scholar
Donnelly, J. P. and Woodruff, J. D. (2007). Intense hurricane activity over the past 5,000 years controlled by El Niño and the West African monsoon. Nature, 447, 465–468.CrossRefGoogle ScholarPubMed
Donnelly, J. P. and Giosan, L. (2008). Tempestuous highs and lows in the Gulf of Mexico. Geol., 9, 751–752. https://doi.org/10.1130/focus092008.1.Google Scholar
Donnelly, J. P., et al. (2015). Climate forcing of unprecedented intense-hurricane activity in the last 2000 years. Earth’s Future, 3, 49–65. https://doi.org/10.1002/2014EF000274.CrossRefGoogle Scholar
Dougherty, A. J. (2014). Extracting a record of Holocene storm erosion and deposition preserved in the morphostratigraphy of a prograded coastal barrier. Cont. Shelf Res., 86, 116–131.CrossRefGoogle Scholar
Elshinnawy, A. I., et al. (2018). On the influence of wave directional spreading on the equilibrium planform of embayed beaches. Coast. Eng., 133, 59–75.CrossRefGoogle Scholar
Elsner, J. B. and Jagger, T. H., (eds.). (2009). Hurricanes and Climate Change. Springer, New York, 419pp.CrossRefGoogle Scholar
Fairbridge, R. W. (2004). Classification of coasts. J. Coast. Res., 20(1), 155–165. https://doi.org/10.2112/1551-5036(2004)20[155:COC]2.0.CO;2.Google Scholar
Fitzgerald, D., et al. (2012). Morphodynamics and facies architecture of tidal inlets and tidal deltas. Chapter 12. In Davis, R. A. Jr. and Dalrymple, R. W. (eds.), Principles of Tidal Sedimentology. Springer, Dordrecht, pp. 301–333. https://doi.org/10.1007/978-94-007-0123-6_12.Google Scholar
Flood, P. G. (1986). Sensitivity of coral cays to climatic variations, southern Great Barrier Reel Australia. Coral Reefs, 5, 13–18.CrossRefGoogle Scholar
Forsyth, A., et al. (2010). Beach ridge plain evidence of a variable late-Holocene tropical cyclone climate, North Queensland, Australia. Palaeogeogr. Palaeoclimatol. Palaeoecol., 297, 707–716. https://doi.org/10.1016/j.palaeo.2010.09.024.CrossRefGoogle Scholar
Frappier, A., et al. (2007). Perspective: Coordinating paleoclimate research on tropical cyclones with hurricane–climate theory and modeling. Tellus A: Dyn. Meteorol. Oceanogr., 59A(4), 529–537. http://dx.doi.org/10.1111/j.1600-0870.2007.00250.x.Google Scholar
Fruergaard, M., et al. (2013). Major coastal impact induced by a 1000-year storm event. Sci. Rep., 3(1), 1051. https://doi.org/10.1038/srep01051.CrossRefGoogle Scholar
Fryberger, S. G. and Dean, G. (1979). Dune forms and wind regime. In McKee, E. D. (ed), A Study of Global Sand Seas. Geological Survey Professional Paper 1052. US Government Printing Office, Washington, pp. 137–170.Google Scholar
George, D. A., et al. (2015). Classification of rocky headlands in California with relevance to littoral cell boundary delineation. Mar. Geol., 369, 137–152. https://doi.org/10.1016/j.margeo.2015.08.010.CrossRefGoogle Scholar
Giosan, L. (2007). Morphodynamic feedbacks on deltaic coasts: Lessons from the wave-dominated Danube delta. Coastal Sediments 2007, 828–841. https://doi.org/10.1061/40926(239)63.CrossRefGoogle Scholar
Godoi, V. A., et al. (2018). Storm wave clustering around New Zealand and its connection to climatic patterns. Int. J. Climatol., 38, S1, 401–417. https://doi.org/10.1002/joc.5380.CrossRefGoogle Scholar
Goff, J., et al. (2011). Palaeotsunamis in the Pacific Islands. Earth-Sci. Rev., 107, 141–146.CrossRefGoogle Scholar
Goff, J., et al. (2014). What is a mega-tsunami? Mar. Geol., 358(2014), 12–17.CrossRefGoogle Scholar
Goodwin, I. D. (2005). A mid-shelf wave direction climatology for south-eastern Australia, and its relationship to the El Nino – Southern Oscillation, since 1877 AD. Int. J. Climatol., 25, 1715–1729.Google Scholar
Goodwin, I. D., et al. (2006). Wave climate, sediment budget and shoreline alignment evolution of the Iluka-Woody Bay sand barrier, northern NSW, Australia, since 3000 yr BP. Mar. Geol., 226, 127–144.CrossRefGoogle Scholar
Goodwin, I. D., et al. (2013). An insight into headland sand bypassing and wave climate variability from shoreface bathymetric change at Byron Bay, New South Wales, Australia. Mar. Geol., 341, 29–45.CrossRefGoogle Scholar
Goodwin, I. D., et al. (2016). Tropical-Extratropical origin storm wave types and their influence on the East Australian Longshore Sand Transport System under a changing climate. J. Geophys. Res. Oceans, 121, 4833–4853, https://doi.org/10.1002/2016JC011769.CrossRefGoogle Scholar
Goodwin, I. D., et al. (2020). Coastal sediment compartments, wave climate and centennial-scale sediment budget: The south-eastern Australian example. Chapter 25. In Jackson, D., Short, A. (eds.), Sandy Beach Morphodynamics. Elsevier, Amsterdam, pp. 615–640.Google Scholar
Goodwin, I. D., et al. (2023). Robbins Island: The index site for regional Last Interglacial sea level, wave climate and the subtropical ridge around Bass Strait, Australia. Quat. Sc. Rev., 305, 107996. https://doi.org/10.1016/j.quascirev.2023.107996.CrossRefGoogle Scholar
Goodwin, I. D., et al. (in prep 2025). Tasman Sea wave climate, coastal impacts and extreme storm activity in the context of large-scale atmospheric circulation, since 1500 CE.Google Scholar
Goslin, J. and Clemmensen, L. B. (2017). Proxy records of Holocene storm events in coastal barrier systems: Storm-wave induced markers. Quat. Sci. Rev., 174, 80–119.CrossRefGoogle Scholar
Gourlay, M. R. (1983). Accretion and erosion of coral cays and some implications for the management of Marine Parks. In Baker, J. T., et al. (eds.), Proceedings of the Inaugural Great Barrier Reef Conference. JCU Press, Townsville, pp. 475–482.Google Scholar
Guanche, Y., et al. (2013). Climate-based Monte Carlo simulation of trivariate sea states. Coast. Eng., 80, 107–121.CrossRefGoogle Scholar
Guisado-Pintado, E. and Jackson, D. W. T. (2019). Coastal impact from high-energy events and the importance of concurrent forcing parameters: The cases of Storm Ophelia (2017) and Storm Hector (2018) in NW Ireland. Front. Earth Sci., 7, 190. https://doi.org/10.3389/feart.2019.00190.CrossRefGoogle Scholar
Haig, J., et al. (2014). Australian tropical cyclone activity lower than at any time over the past 550–1,500 years. Nature, 505, 667–671.CrossRefGoogle Scholar
Hallermeier, R. J. (1981). A profile zonation for seasonal sand beaches from wave climate. Coast. Eng., 4, 253–277.Google Scholar
Halligan, G. H. (1906). Sand movement on the New South Wales coast. Proc. Linnean Soc. New South Wales, 31, 619–640.Google Scholar
Hamilton, L. J. (2010). Characterising spectral sea wave conditions with statistical clustering of actual spectra. Appl. Ocean Res., 32, 332–342.CrossRefGoogle Scholar
Hamon-Kerivel, K., et al. (2020). Shoreface mesoscale morphodynamics: A review. Earth-Sci. Rev., 209, 103330.CrossRefGoogle Scholar
Hamylton, S. M. and Puotinen, M. (2015). A meta-analysis of reef island response to environmental change on the Great Barrier Reef. Earth Surf. Process. Landforms, 40, 1006–1016. https://doi.org/10.1002/esp.3694.CrossRefGoogle Scholar
Hanson, H. (1989). Genesis: A generalised shoreline change numerical model. J. Coast. Res., 5(1), 1–27.Google Scholar
Harley, M. D., et al. (2015). New insights into embayed beach rotation: The importance of wave exposure and cross- shore processes. J. Geophys. Res. F Earth Surf., 120, 1470–1484.Google Scholar
Hayes, M. O. (1979). Barrier island morphology as a function of tidal and wave regime. In Leatherman, S. P. (ed), Barrier Islands: From the Gulf of St. Lawrence to the Gulf of Mexico. Academic, New York. 1–27.Google Scholar
Hayne, M. and Chappell, J. (2001). Cyclone frequency during the last 5000 years at Curacoa Island, north Queensland, Australia. Palaeogeogr. Palaeoclimatol. Palaeoecol., 168, 207–219.CrossRefGoogle Scholar
Hearty, P. J. and Tormey, B. R. (2017). Sea-level change and super- storms; geologic evidence from the last interglacial (MIS 5e) in the Bahamas and Bermuda offers ominous prospects for a warming Earth. Mar. Geol., 390, 347–365. http://dx.doi.org/10.1016/j.margeo.2017.05.009.CrossRefGoogle Scholar
Hein, C. and Ashton, A. (2020). Long-term shoreline morphodynamics: Processes and preservation of environmental signals. In Jackson, D. W. T. and Short, A. D. (eds.), Sandy Beach Morphodynamics. https://doi.org/10.1016/B978-0-08-102927-5.00021-7.Google Scholar
Hemer, M. A., et al. (2008). A classification of wave generation characteristics during large wave events on the Southern Australian margin. Cont. Shelf Res., 28, 634–652.CrossRefGoogle Scholar
Hemer, M. A., et al. (2009). Variability and trends in the directional wave climate of the Southern Hemisphere. Int. J. Climatol., 30, 475–491.Google Scholar
Hemer, M. A., et al. (2013). Projected changes in wave climate from a multi-model ensemble. Nat. Clim. Change, 3, 471–476. https://doi.org/10.1038/NCLIMATE1791.CrossRefGoogle Scholar
Hesp, P. A. and Thom, B. G. (1990). Geomorphology and evolution of active transgressive dunefields. In Nordstrom, K. F., et al. (eds.), Coastal Dunes: Form and Process. John Wiley and Sons, Chichester, pp. 253–288.Google Scholar
Hesp, P. A. (2002). Foredunes and blowouts: Initiation, geomorphology and dynamics. Geomorphol., 48, 245–268.CrossRefGoogle Scholar
Hesp, P. A. (2013). Conceptual models of the evolution of transgressive dune field systems. Geomorphol., 199, 138–149.CrossRefGoogle Scholar
Hesp, P. A., et al. (2007). Regional wind fields and dunefield migration, southern Brazil. Earth Surf. Process. Landf., 32, 561–573. https://doi.org/10.1002/esp.1406.CrossRefGoogle Scholar
Hesp, P. A., et al. (2015). Flow deflection over a foredune. Geomorphol., 230, 64–74.CrossRefGoogle Scholar
Hesp, P. A., et al. (2022). Review and direct evidence of transgressive Aeolian sand sheet and dunefield initiation. Earth Surf. Process. Landf., 47, 2660–2675.CrossRefGoogle Scholar
Hesse, P. P., et al. (2017). Complexity confers stability: Climate variability, vegetation response and sand transport on longitudinal sand dunes in Australia’s deserts. Aeolian Res., 25, 45–61.CrossRefGoogle Scholar
Holthuijsen, L. H. (2007). Waves in Oceanic and Coastal Waters. Cambridge University Press, Cambridge, 379pp.CrossRefGoogle Scholar
Horikawa, K. (1988). Nearshore Dynamics and Coastal Processes: Theory, Measurement and Predictive Models. University of Tokyo Press, Tokyo, 522pp.Google Scholar
Hsu, J. R. C. and Evans, C. (1989). Parabolic bay shapes and applications. Proc., Inst. Civ. Eng., 87(2), 557–570.Google Scholar
Hsu, J. R. C., et al. (2008). Appreciation of static bay beach concept for coastal management and protection. J. Coast. Res., 24(1), 198–215.Google Scholar
Hsu, J. R. C., et al. (2010). Static bay beach concept for scientists and engineers: A review. Coast. Eng., 57, 76–91.CrossRefGoogle Scholar
Hubbert, G. D. and Mclnnes, K. L. (1999). A storm surge inundation model for coastal planning and impact studies. J. Coast. Res., 15, 168–185.Google Scholar
Hutton, E. W. H. and Syvitski, J. P. M. (2008). Sedflux 2.0: An advanced process-response model that generates three-dimensional stratigraphy. Comput. Geosci., 34, 1319–1337.CrossRefGoogle Scholar
Hyder, Consulting Pty Ltd, et al. (1997). Tweed River Entrance Sand Bypassing Project permanent sand bypassing system environmental impact statement/impact assessment study (June 1997).Google Scholar
Iglesias, G., et al. (2010). Artificial intelligence and headland bay beaches. Coast. Eng., 57, 176–183.CrossRefGoogle Scholar
Jackson, P. S. and Hunt, J. C. R. (1975). Turbulent wind flow over a low hill. Q. J. Roy. Meteorol. Soc., 101, 929–955.CrossRefGoogle Scholar
Jackson, D. and Short, A. (eds.). (2020). Sandy Beach Morphodynamics. Elsevier, 793pp. https://doi.org/10.1016/B978-0-08-102927-5.00021-7.Google Scholar
Jahnert, R., et al. (2012). Evolution of a coquina barrier in Shark Bay, Australia by GPR imaging: Architecture of a Holocene reservoir analog. Sed. Geol., 281, 59–74.CrossRefGoogle Scholar
Jelgersma, S., et al. (1995). Holocene storm surge signatures in the coastal dunes of the western Netherlands. Mar. Geol., 125, 95–110.CrossRefGoogle Scholar
Jol, H. M., (ed). (2009). Ground Penetrating Radar: Theory and Applications. Elsevier, Amsterdam, The Netherlands, 524pp.Google Scholar
Jol, H. M. and Bristow, C. S. (2003). GPR in sediments: Advice on data collection, basic processing and interpretation, a good practice guide. In Bristow, C. S. and Jol, H. M. (eds.), Ground Penetrating Radar in Sediments. Geological Society of London Special Publication 211, Geological Society of London, London, pp. 9–27.Google Scholar
Kamphuis, J. W., et al. (1986). Calculation of littoral sand transport rate. Coast. Eng., 10, 1–21.CrossRefGoogle Scholar
Karunarathna, H., et al. (2014). The effects of storm clustering on beach profile variability. Mar. Geol., 348, 103–112.CrossRefGoogle Scholar
Kench, P. S. and Brander, R. W. (2006). Response of reef island shorelines to seasonal climate oscillations: South Maalhosmadulu atoll, Maldives. J. Geophys. Res., 111, F01001, https://doi.org/10.1029/2005JF000323.Google Scholar
Kench, P., et al. (2009). Coral reefs. In Slaymaker, O., et al. (eds.), Geomorphology and Global Environmental Change. Cambridge University Press, Cambridge, pp. 180–213.Google Scholar
Klein, A. H. F., et al. (2003). Visual assessment of bayed beach stability using computer software. Comput. Geosci., 29, 1249–1257.Google Scholar
Klein, A. H. F., et al. (2020). Headland bypassing and overpassing: Form, processes and applications. Chapter 23. In Jackson, D. and Short, A. (eds.), Sandy Beach Morphodynamics. Elsevier, Amsterdam, pp. 557–591. 793pp. https://doi.org/10.1016/B978-0-08-102927-5.00021-7.Google Scholar
Komar, P. D. (1998). Beach Processes and Sedimentation, 2nd ed. Prentice Hall, New Jersey, 544pp.Google Scholar
Lancaster, N. (1988). Development of linear dunes in the southwestern Kalahari, southern Africa. J. Arid Environ., 14, 23–244.CrossRefGoogle Scholar
Lane, C. S., et al. (2017). Verification of tropical cyclone deposits with oxygen isotope analyses of coeval ostracod valves. J. Paleolimnol., 57, 245–255. https://doi.org/10.1007/s10933-017-9943-5.Google Scholar
Lausman, R., et al. (2010a). Uncertainty in the application of parabolic bay shape equation: Part 1. Coast. Eng., 57, 132–141.Google Scholar
Lausman, R., et al. (2010b). A uncertainty in the application of parabolic bay shape equation: Part 2. Coast. Eng., 57, 142–151.Google Scholar
Lazarus, E. D., et al. (2019). Environmental signal shredding on sandy coastlines. Earth Surf. Dynam., 7, 77–86. https://doi.org/10.5194/esurf-7-77-2019.CrossRefGoogle Scholar
Lebrec, U., et al. (2022). Morphology and distribution of submerged palaeoshorelines: Insights from the North West Shelf of Australia. Earth Sci. Rev. 224, 103864. https://doi.org/10.1016/j.earscirev.2021.103864.CrossRefGoogle Scholar
Lettau, H., et al. (2022). Experimental and micro-meteorological field studies of dune migration. IES Report 101. In Lettau, H. H. and Lettau, K. (eds.), Exploring the World’s Driest Climate. University of Wisconsin, Madison, pp. 110–147.Google Scholar
Lim, C., et al. (2022). MeePaSoL: A MATLAB-based GUI software tool for shoreline management. Comput. Geosci., 16, 105059. https://doi.org/10.1016/j.cageo.2022.105059.Google Scholar
Liu, K.-B. (2004). Paleotempestology: Principles, methods, and examples from Gulf coast lake-sediments. In Murnane, R. and Liu, K.-B. (eds.), Hurricanes and Typhoons: Past, Present, and Future. Columbia University Press, New York, pp. 13–57.Google Scholar
Liu, K. and Fearn, M. (2000). Reconstruction of prehistoric landfall frequencies of catastrophic hurricanes in northwestern Florida from lake sediment records. Quat. Res., 54, 238–245. https://doi.org/10.1006/qres.2000.2166.CrossRefGoogle Scholar
Liu, E., et al. (2016). U-Th age distribution of coral fragments from multiple rubble ridges within the Frankland Islands, Great Barrier Reef: Implications for past storminess history. Quat. Sci. Rev., 143, 51–68. http://dx.doi.org/10.1016/j.quascirev.2016.05.006.CrossRefGoogle Scholar
Loope, D. B., et al. (2004). Tropical westerlies over Pangean sand seas. Sedimentol., 51, 315–322. https://doi.org/10.1046/j.1365-3091.2003.00623.x.CrossRefGoogle Scholar
Loureiro, C. and Ferreira, O. (2020). Mechanisms and timescales of beach rotation. Chapter 24. In Jackson, D. and Short, A. (eds.), Sandy Beach Morphodynamics. Elsevier, pp. 593–614. 793pp. https://doi.org/10.1016/B978-0-08-102927-5.00021-7.Google Scholar
Masselink, G. and Pattiaratchi, C. B. (2001). Characteristics of the sea-breeze system in Perth, Western Australia, and its effect on the nearshore wave climate. J. Coast. Res., 17(1), 173–187.Google Scholar
Masselink, G. and van Heteren, S. (2014). Response of wave-dominated and mixed-energy barriers to storms. Mar. Geol., 352, 321–347. https://doi.org/10.1016/j.margeo.2013.11.004.CrossRefGoogle Scholar
Masselink, G., et al. (2014). Role of wave forcing, storms and NAO in outer bar dynamics on a high-energy, macro-tidal beach. Geomorphol., 226, 76–93.CrossRefGoogle Scholar
Masselink, G., et al. (2016). Extreme wave activity during 2013/2014 winter and morphological impacts along the Atlantic coast of Europe. Geophys. Res. Lett., 43, https://doi.org/10.1002/2015GL067492.CrossRefGoogle Scholar
Menéndez, M., et al. (2008). Variability of extreme wave heights in the northeast Pacific Ocean based on buoy measurements. Geophys. Res. Lett., 35(22), 1–6. https://doi.org/10.1029/2008GL035394.CrossRefGoogle Scholar
Miguel, L. L. A. J. and Castro, J. W. A., et al. (2018). Aeolian dynamics of transgressive dunefields on the southern Mozambique coast, Africa. Earth Surf. Process. Landf., 43, 2533–2546.CrossRefGoogle Scholar
Mortlock, T. R. and Goodwin, I. D. (2015). Directional wave climate and power variability in the Tasman Sea. Cont. Shelf Res., 98, 36–53.Google Scholar
Mortlock, T. R. and Goodwin, I. D. (2016). Impacts of enhanced central Pacific ENSO on wave climate and headland-bay beach morphology. Cont. Shelf Res., 120, 14–25.CrossRefGoogle Scholar
Mortlock, T. R., et al. (2017). Open Beaches Project 1A – Quantification of Regional Rates of Sand Supply to the NSW Coast: Numerical Modelling Report. Department of Environmental Sciences and Risk Frontiers, Macquarie University, Sydney, pp. 155.Google Scholar
Mortlock, T. R., et al. (2020). Influence of the subtropical ridge on directional wave power in the Southeast Indian Ocean. Int. J. Climatol., 40(12), 5352–5367. https://doi.org/10.1002/joc.6522.CrossRefGoogle Scholar
Mortlock, T. R., et al. (2023). A long-term view of tropical cyclone risk in Australia. Nat. Hazards., 118(1), 571–588. https://doi.org/10.1007/s11069-023-06019-5.CrossRefGoogle Scholar
Morton, R. A. and Sallenger, A. H., Jr. (2003). Morphological impacts of extreme storms on sandy beaches and barriers. J. Coast. Res., 19(3), 560–573.Google Scholar
Munk, W. H. and Traylor, M. A. (1947). Refraction of ocean waves: A process linking under- water topography to beach erosion. J. Geol., 55(1), 1–26.CrossRefGoogle Scholar
Murray, A. S. and Wintle, A. G. (2000). Luminescence dating of quartz using an improved single-aliquot regenerative-dose protocol. Radiat. Meas., 32, 57–73.Google Scholar
Murray, A. B. and Ashton, A. D. (2013). Instability and finite- amplitude self-organization of large-scale coastline shapes. Philos. Trans. R. Soc. A Math. Physical Eng. Sci., 371, 20120363, https://doi.org/10.1098/rsta.2012.0363.Google ScholarPubMed
Murray, A. B., et al. (2020). From cusps to capes: Self-organised shoreline shapes. Chapter 12. In Jackson, D. and Short, A. (eds.), Sandy Beach Morphodynamics. Elsevier, pp. 277–295, 793pp. https://doi.org/10.1016/B978-0-08-102927-5.00021-7.Google Scholar
Nahon, A., et al. (2019). Imprints of wave climate and mean sea level variations in the dynamics of a coastal spit over the last 250 years: Cap Ferret, SW France. Earth Surf. Process. Landf., 44, 2112–2125. https://doi.org/10.1002/esp.4634.CrossRefGoogle Scholar
Neal, A. (2004). Ground-penetrating radar and its use in sedimentology: Principles, problems, and progress. Earth-Sci. Rev., 66, 261–330. https://doi.org/10.1016/j.earscirev.2004.01.004.CrossRefGoogle Scholar
Nienhuis, J. H., et al. (2013). Wave reworking of abandoned deltas. Geophys. Res. Lett., 40, 5899–5903. https://doi.org/10.1002/2013GL058231.CrossRefGoogle Scholar
Nott, J. (2004). The tsunami hypothesis – Comparison of field evidence against the effects, on the Western Australian coast, of some of the most powerful storms on Earth. Mar. Geol., 208, 1–12.CrossRefGoogle Scholar
Nott, J. (2006). Extreme Events: A Physical Reconstruction and Risk Assessment. Cambridge University Press, Cambridge, 297pp.CrossRefGoogle Scholar
Nott, J. (2011). A 6000 year tropical cyclone record from Western Australia. Quat. Sci. Rev., 30, 713–722.CrossRefGoogle Scholar
Nott, J. (2012). Storm tide recurrence intervals – A statistical approach using beach ridge plains in Northern Australia. Geograph. Res., 50(4), 368–376.CrossRefGoogle Scholar
Nott, J. and Hayne, M. (2001). High frequency of ‘super- cyclones’ along the Great Barrier Reef over the past 5,000 years. Nature, 413, 508–512.CrossRefGoogle ScholarPubMed
Nott, J., et al. (2009). Sand beach ridges record 6000 year history of extreme tropical cyclone activity in northeastern Australia. Quat. Sci. Rev., 28(15–16), 1511–1520.CrossRefGoogle Scholar
Nott, J. and Forsyth, A. (2012). Punctuated global tropical cyclone activity over the past 5,000 years. Geophys. Res. Lett., 39, L14703. https://doi.org/10.1029/2012GL052236.CrossRefGoogle Scholar
Nott, J. F. and Jagger, T. H. (2013). Deriving robust return periods for tropical cyclone inundations from sediments. Geophys. Res. Lett., 40, 370–373. https://doi.org/10.1029/2012GL054455.CrossRefGoogle Scholar
Nott, J., et al. (2013). Anatomy of sand beach ridges: Evidence from severe Tropical Cyclone Yasi and its predecessors, northeast Queensland, Australia. J. Geophys. Res. Earth Surf., 118, 1710–1719.Google Scholar
Nott, J. A., et al. (2015). The origin of centennial- to millennial-scale chronological gaps in storm emplaced beach ridge plains. Mar. Geol., 367, 83–93. http://dx.doi.org/10.1016/j.margeo.2015.05.011.CrossRefGoogle Scholar
Nutz, A. et al. (2016). Wind-driven waterbodies: a new category of lake within an alternative sedimentologically-based lake classification. J. Paleolimnol., 59, 189–199.CrossRefGoogle Scholar
Oliver, T. S. N., et al. (2015). Towards more robust chronologies of coastal progradation: Optically stimulated luminescence ages for the coastal plain at Moruya, south-eastern Australia. The Holocene, 25(3), 536–546. https://doi.org/10.1177/0959683614561886.CrossRefGoogle Scholar
Otvos, E. G. (2000). Beach ridges and definitions and significance. Geomorphol., 32, 83–108. http://dx.doi.org/10.1016/S0169-555X(99)00075-6.Google Scholar
Pringle, J. J., et al. (2014). Automated classification of the atmospheric circulation patterns that drive regional wave climates. Nat. Hazards Earth Syst. Sci., 14, 2145–2155.CrossRefGoogle Scholar
Pringle, J. J., et al. (2015). Atmospheric circulation patterns that drive extreme wave events: A new framework for coastal vulnerability assessment. Nat. Hazards, 79, 45–59. https://doi.org/10.1007/s11069-015-1825-4.CrossRefGoogle Scholar
Pringle, J. and Stretch, D. D. (2019). A new approach for the stochastic simulation of regional wave climates conditioned on synoptic scale meteorology. J. Coast. Res., 35(6), 1331–1342.CrossRefGoogle Scholar
Pye, K. and Lancaster, N. (1993). Late quaternary development of coastal parabolic megadune complexes in Northeastern Australia. In Pye, K. (ed), Aeolian Sediments. Blackwell Publishing Ltd, Oxford, pp. 23–44. https://doi.org/10.1002/9781444303971.ch3.CrossRefGoogle Scholar
Pye, K. and Tsoar, H. (2009). Aeolian Sand and Sand Dunes. Springer, Berlin, Heidelberg.CrossRefGoogle Scholar
Raabe, A. L. A., et al. (2010). MEPBAY and SMC: Software tools to support different operational levels of headland bay beaches in coastal engineering projects. Coast. Eng., 57, 213–226.CrossRefGoogle Scholar
Rhodes, E. G. (1982). Depositional model for a chenier plain, Gulf of Carpentaria, Australia. Sedimentol., 29, 201–221.Google Scholar
Ribo, M., et al. (2020). Shelf sand supply determined by glacial-age sea-level modes, submerged coastlines and wave climate. Sci. Rep., 10, 462. https://doi.org/10.1038/s41598-019-57049-8.CrossRefGoogle ScholarPubMed
Roelvink, D. and Reniers, A. (2012). A Guide to Modeling Coastal Morphology. Advances in Coastal and Ocean Engineering, Volume 12, World Scientific Publishing, Singapore, 274pp.Google Scholar
Rogers, S. S., et al. (2004). Coastal change and beach ridges along the northwest coast of Peru: Image and GIS analysis of the Chira, Piura, and Colán beach ridge plains. J. Coast. Res., 20(4), 1102–1125.Google Scholar
Rosati, J. D., et al. (2002). Longshore sediment transport. In Vincent, L. and Demirbilek, Z. (eds.), Coastal Engineering Manual, Part III, Engineer Manual 5421110-2-1100, chap. III-2, U.S. Army Corps of Eng., Washington, DC.Google Scholar
Rouillard, A., et al. (2015). Evidence for extreme floods in arid subtropical northwest Australia during the Little Ice Age chronozone (CE 1400–1850). Quat. Sci. Rev., 144, 107–122.Google Scholar
Rovere, A., et al. (2017). Giant boulders and Last Interglacial storm intensity in the North Atlantic. Proc. Natl. Acad. Sci. USA, 114(46), 12144–12149. https://doi.org/10.1073/pnas.1712433114.CrossRefGoogle ScholarPubMed
Roy, P. S. and Thom, B. G. (1981). Late Quaternary marine deposition in New South Wales and southern Queensland, an evolutionary model. J. Geol. Soc. Aust., 28, 471–489.CrossRefGoogle Scholar
Roy, P. S., et al. (1994). Wave dominated coasts. Chapter 4. In Carter, R. W. G. and Woodroffe, C. D. (eds.), Coastal Evolution: Late Quaternary Shoreline Morphodynamics. Cambridge University Press, Cambridge, pp. 121–186.Google Scholar
Roy, P. S., (1999). Heavy mineral beach sand placers in Southeastern Australia their nature and genesis. Econ. Geol., 94, 567–588.CrossRefGoogle Scholar
Sallenger, A. H. (2000). Storm impact scale for barrier islands. J. Coast Res., 16, 890–895.Google Scholar
Sanderson, P. G. and Eliot, I. (1996). Shoreline salients, cuspate forelands and tombolos on the Coast of Western Australia. J. Coast Res., 12(3), 761–773.Google Scholar
Sandweiss, D. H. (1986). The beach ridges at Santa, Peru: El Niño, uplift and prehistory. Geoarchaeology, 1, 17–28.CrossRefGoogle Scholar
Sandweiss, D. H., et al. (2007). Mid-Holocene climate and cultural change in coastal Peru. In Anderson, D. G., et al., (eds.), Climatic Change and Cultural Dynamics: A Global Perspective on Mid-Holocene Transitions. Academic, San Diego, pp. 25–50.Google Scholar
Scheffers, A. and Kelletat, D. (2003). Sedimentologic and geomorphologic tsunami imprints worldwide – A review. Earth Sci. Rev., 63, 83–92.CrossRefGoogle Scholar
Scoffin, T. P. (1993). The geological effects of hurricanes on coral reefs and the interpretation of storm deposits. Coral Reefs, 12, 203–221.CrossRefGoogle Scholar
Short, A. D., (ed), (1999). Beach and Shoreface Morphodynamics. John Wiley and Sons, Chichester, 379pp.Google Scholar
Short, A. D. and Masselink, G. (1999). Embayed and structurally controlled beaches. In Short, A. D. (ed), Handbook of Beach and Shoreface Morphodynamics. Wiley, New York, pp. 230–249.Google Scholar
Silva, A. P., et al. (2020). Climate-induced variability in South Atlantic wave direction over the past three millennia. Sci. Rep., 10, 18553. https://doi.org/10.1038/s41598-020-75265-5.CrossRefGoogle ScholarPubMed
Silveira, L. F., et al. (2010). Headland-bay beach planform stability of Santa Catarina State and of the northern coast of Sao Paulo State. Brazilian J. Oceanogr., 58(2), 101–122.CrossRefGoogle Scholar
Silvester, R. (1960). Stabilization of sedimentary coastlines. Nature, 188, 467–469.CrossRefGoogle Scholar
Silvester, R. and Hsu, J. R. C. (1997). Coastal Stabilization. World Scientific, Singapore, 578p.CrossRefGoogle Scholar
Smith, A. W. (2001). Headland bypassing. Coasts & Ports 2001: Proceedings of the 15th Australasian Coastal and Ocean Engineering Conference, the 8th Australasian Port and Harbour Conference. Institution of Engineers, Australia, Barton, A.C.T., pp. 214–216.Google Scholar
Splinter, K. D., et al. (2014). A relationship to describe the cumulative impact of storm clusters on beach erosion. Coast. Eng., 83, 49–55.CrossRefGoogle Scholar
Stolper, D., et al. (2005). Simulating the evolution of coastal morphology and stratigraphy with a new morphological-behaviour model (GEOMBEST). Mar. Geol., 218, 17–36.CrossRefGoogle Scholar
Storlazzi, C. D. and Griggs, G. B. (2000). Influence of El Niño-Southern Oscillation (ENSO) events on the evolution of central California’s shoreline. Geol. Soc. Amer. Bull., 112(2), 236–249.2.0.CO;2>CrossRefGoogle Scholar
Storlazzi, C. D. and Wingfield, D. K. (2005). The spatial and temporal variability in oceanographic and meteorologic forcing along Central California – 1980–2002. U.S. Geological Survey Scientific Investigations Report 2005–5085, 39p.Google Scholar
Storlazzi, C. D., et al. (2018). Most atolls will be uninhabitable by the mid-21st century because of sea-level rise exacerbating wave-driven flooding. Sci. Adv., 4(4), eaap9741.CrossRefGoogle Scholar
Storms, J. E. A., et al. (2002). Process-response modelling of wave-dominated coastal systems: Simulating evolution and stratigraphy on geologic timescales. J. Sed. Res., 72, 226–239.CrossRefGoogle Scholar
Sunamura, T. and Misuzo, O. (1987). A study on depositional shoreline forms behind an island. Ann. Rep. Inst. Geosci. Univ. Tsukuba, 13, 71–73.Google Scholar
Suter, J. R. (1994). Deltaic coasts. In Carter, R. W. G. and Woodroffe, C. D. (eds.), Coastal Evolution: Late Quaternary Shoreline Morphodynamics. Cambridge: Cambridge University Press, Tulsa, pp. 87–114.Google Scholar
Tamura, T. (2012). Beach ridges and prograded beach deposits as palaeoenvironment records. Earth Sci. Rev., 114, 279–297.CrossRefGoogle Scholar
Tamura, T., et al. (2018). Coarse-sand beach ridges at Cowley Beach, north-eastern Australia: Their formative processes and potential as records of tropical cyclone history. Sedimentology, 65(3), 721–744.CrossRefGoogle Scholar
Tamura, T., et al. (2019). Recurrence of extreme coastal erosion in SE Australia beyond historical timescales inferred from beach ridge morphostratigraphy. Geophys. Res. Lett., 46, 4705–4714. https://doi.org/10.1029/2019GL083061.CrossRefGoogle Scholar
Thom, B. G. (1978). Coastal sand deposition in southeast Australia during the Holocene. Chapter 9. In Davies, J. J. and Williams, M. A. J. (eds.), Landform Evolution in Australasia. ANU Press, Canberra, Australia, pp. 197–214, 376pp.Google Scholar
Thomas, C. W., et al. (2016). Complex coastlines responding to climate change: Do shoreline shapes reflect present forcing or ‘remember’ the distant past? Earth Surf. Dynam., 4, 871–884. https://doi.org/10.5194/esurf-4-871-2016.CrossRefGoogle Scholar
Trouet, V., et al. (2016). Shipwreck rates reveal Caribbean tropical cyclone response to past radiative forcing. Proc. Natl. Acad. Sci. USA, 113(12), 3169–3174. https://doi.org/10.1073/pnas.1519566113.CrossRefGoogle ScholarPubMed
Vakarelov, B. K. and Ainsworth, R. B. (2013). A hierarchical approach to architectural classification in marginal-marine systems: Bridging the gap between sedimentology and sequence stratigraphy. AAPG Bulletin, 97(7), 1121–1161.CrossRefGoogle Scholar
Valiente, N. G., et al. (2019). Role of waves and tides on depth of closure and potential for headland bypassing. Mar. Geol., 407, 60–75.CrossRefGoogle Scholar
Vimpere, L., et al. (2019). Chevrons: Origin and relevance for the reconstruction of past wind regimes. Earth-Sci Rev., 193, 317–332.CrossRefGoogle Scholar
Walker, I. J. and Hesp, P. A. (2013). Fundamentals of aeolian sediment transport: Airflow over dunes. Ch. 11.7. In Lancaster, N., et al. (eds.), Vol. 11: Aeolian Geomorphology. In Shroder, J. F. (ed.), Treatise on Geomorphology. Elsevier, Oxford, pp. 109–133.Google Scholar
Walker, I. J., et al. (2017). Scale-dependent perspectives on the geomorphology and evolution of beach- dune systems. Earth-Sci Rev., 171, 220–253.CrossRefGoogle Scholar
Wandres, M., et al. (2017). The effect of the Leeuwin current on offshore surface gravity waves in southwest western Australia. J. Geophys. Res. Oceans, 122, 9047–9067. https://doi.org/10.1002/2017JC013006.CrossRefGoogle Scholar
Wang, D. W. and Hwang, P. A. (2001). An operational method for separating wind sea and swell from ocean wave spectra. J. Atmos. Oceanic. Technol., 18(12), 2052–2063.2.0.CO;2>CrossRefGoogle Scholar
Williams, H. F. L. (2013). 600-year sedimentary archive of hurricane strikes in a pro- grading beach ridge plain, southwestern Louisiana. Mar. Geol., 336, 170–183.Google Scholar
Woodroffe, C. D. (2003). Coasts: Form, Process, Evolution. Cambridge: University Press. 623p.Google Scholar
Wright, L. D. (1977). Sediment transport and deposition at river mouths: A synthesis. Geol. Soc. Am. Bull., 88, 857–868.2.0.CO;2>CrossRefGoogle Scholar
Wright, L. D. (1995). Morphodynamics of Inner Continental Shelves. CRC Press, Boca Raton, Florida, 241pp.Google Scholar
Wright, L. D. and Coleman, J. M. (1973). Variations in morphology of major river deltas as functions of ocean wave and river discharge regimes. Am. Assoc. Pet. Geol. Bull., 57, 370–398.Google Scholar
Wright, L. D. and Short, A. D. (1984). Morphodynamic variability of surf zones and beaches: A synthesis. Mar. Geol., 56, 93–118.Google Scholar
Yan, N. and Baas, A. C. W. (2015). Parabolic dunes and their transformations under environmental and climatic changes: Towards a conceptual framework for understanding and prediction. Glob. Planet. Change., 124, 123–148.CrossRefGoogle Scholar
Yizhaq, H., et al. (2020). The effect of wind speed averaging time on the calculation of sand drift potential: New scaling laws. Earth Planet. Sci. Lett., 544, 116373.CrossRefGoogle Scholar
Zӑinescu, F., et al. (2024). Wave-influenced delta morphodynamics, long-term sediment bypass and trapping controlled by relative magnitudes of riverine and wave-driven sediment transport. Geophys. Res. Lett., 51, e2024GL111069. https://doi.org/10.1029/2024GL111069.CrossRefGoogle Scholar
Zenkovich, V. P. (Ed. Steers, J. A.) (1967). Processes of Coastal Development. Oliver and Boyd, London, 738pp.Google Scholar
Abram, N. J., et al. (2008). Recent intensification of tropical climate variability in the Indian Ocean. Nat. Geosci., 1, 849–853.CrossRefGoogle Scholar
Abram, N. J., et al. (2016). Early onset of industrial-era warming across the oceans and continents. Nature, 536, 411–418, https://doi.org/10.1038/nature19082.CrossRefGoogle ScholarPubMed
Aharon, P. (1991). Recorders of reef environment histories: Stable isotopes in corals, giant clams, and calcareous algae. Coral Reefs, 10, 71–90.CrossRefGoogle Scholar
Alpert, A. E., et al. (2016). Comparison of equatorial Pacific sea surface temperature variability and trends with Sr/Ca records from multiple corals. Paleoceanogr., 31(2), 252–265.CrossRefGoogle Scholar
Andrus, C. F. T. (2011). Shell midden sclerochronology. Quat. Sci. Rev., 30, 2892–2905.Google Scholar
Atsawawaranunt, K., et al. (2018). The SISAL database: A global resource to document oxygen and carbon isotope records from speleothems. Earth Syst. Sci. Data, 10, 1687–1713. https://doi.org/10.5194/essd-10-1687-2018.CrossRefGoogle Scholar
Aubert, A., et al. (2009). The tropical giant clam Hippopus hippopus shell, a new archive of environmental conditions as revealed by sclerochronological and δ18O profiles. Coral Reefs, 28, 989–998. https://doi.org/10.1007/s00338-009-0538-0.CrossRefGoogle Scholar
Azzoug, M., et al. (2012). Reconstructing the duration of the West African Monsoon season from growth patterns and isotopic signals of shells of Anadara senilis (Saloum Delta, Senegal). Palaeogeogr. Palaeoclimatol. Palaeoecol., 346–347, 145–152.Google Scholar
Baker, P. A. and Fritz, S. C. (2016). Nature and causes of quaternary climate variation of tropical South America. Quat. Sci. Rev., 124, 31–47.Google Scholar
Baker, J. C. A., et al. (2016). What drives interannual variation in tree ring oxygen isotopes in the Amazon?, Geophys. Res. Lett., 43, 11831–11840. https://doi.org/10.1002/2016GL071507.CrossRefGoogle Scholar
Baldini, L. M., et al. (2016). Persistent northward North Atlantic tropical cyclone track migration over the past five centuries. Sci. Rep., 6, 375522. https://doi.org/10.1038/srep37522.CrossRefGoogle ScholarPubMed
Bard, E., et al. (1996a). Deglacial sea-level record from Tahiti corals and the timing of global meltwater discharge. Nature, 382, 241–244. https://doi.org/10.1038/382241a0.CrossRefGoogle Scholar
Barnes, D. J. and Taylor, R. B. (2001). On the nature and causes of luminescent lines and bands in coral skeletons. Coral Reefs, 19, 221–230.CrossRefGoogle Scholar
Barnes, D. J., et al. (2003). Measurement of luminescence in coral skeletons. J. Exp. Mar. Biol. Ecol., 295, 91–106. https://doi.org/10.1016/S0022-0981(03)00274-0.CrossRefGoogle Scholar
Barnes, D. J. and Taylor, R. B. (2005). On the nature and causes of luminescent lines and bands in coral skeleton: II. Contribution of skeletal crystals. J. Exp. Mar. Biol. Ecol., 322, 135–142.CrossRefGoogle Scholar
Beck, J. W., et al. (1992). Sea-surface temperature from coral skeletal Strontium/Calcium ratios. Science, 257, 644–647.CrossRefGoogle ScholarPubMed
Berkman, P. A., et al. (2004). Polar emergence and the influence of increased sea-ice extent on the Cenozoic biogeography of pectinid molluscs in Antarctic coastal areas. Deep Sea Res., Part II, Topical Studies in Oceanography, 5(14), 1839–1855.Google Scholar
Berlage, H. P. (1931). Over het verband tusschen de dikte der jaarringen van djatiboomen (Tectona grandis L. f.) en den regenval op Java. Tectona, 24, 939–953.Google Scholar
Bird, M. I., et al. (2020). Stable isotope proxy records in tropical terrestrial environments. Palaeogeogr. Palaeoclimatol. Palaeoecol., 538, 109445.CrossRefGoogle Scholar
Blanchon, P. and Shaw, J. (1995). Reef drowning during the last deglaciation: Evidence for catastrophic sea- level rise and ice -sheet collapse. Geol., 23(1), 4–8.2.3.CO;2>CrossRefGoogle Scholar
Boninsegna, J. A., et al. (2009). Dendroclimatological reconstructions in South America: A review. Palaeogeogr. Palaeoclimatol. Palaeoecol., 281, 210–228.CrossRefGoogle Scholar
Bourman, R. P., et al. (2016). Last interglacial (MIS 5e) sea-level determined from a tectonically stable, far-field location, Eyre Peninsula, southern Australia. Aust. J. Earth Sci., 63, 611–630.Google Scholar
Bradley, R. S. (1999). Paleoclimatology: Reconstructing Climates of the Quaternary. 2nd ed. International Geophysics Press, Volume 64, 610pp, Academic Press, San Diego.Google Scholar
Bradley, R. S. (2011). Dendrochronology: Progress and prospects, developments in Paleoenvironmental research. In Bradley, R. S. (ed), High Resolution Paleoclimatology, vol. 11. Springer, New York, pp. 3–16.CrossRefGoogle Scholar
Brienen, R. J. W., et al. (2012). Oxygen isotopes in tree rings are a good proxy for Amazon precipitation and El Niño-Southern Oscillation variability. Proc. Natl. Acad. Sci. U.S.A., 109, 16957–16962. https://doi.org/10.1073/pnas.1205977109.CrossRefGoogle ScholarPubMed
Brienen, R. J. W., et al. (2016). Tree rings in the tropics: Insights into the ecology and climate sensitivity of tropical trees. In Goldstein, G. and Santiago, L. S. (eds.), Tropical Tree Physiology 6, 439–461. https://doi.org/10.1007/978-3-319-27422-5_20.Google Scholar
Buckley, B. M., et al. (1995). Dendrochronological investigations in Thailand. IAWA Journal, 16(4), 393–409.CrossRefGoogle Scholar
Buckley, B. M., et al. (2007). Decadal scale droughts over northwestern Thailand over the past 448 years: Links to the tropical Pacific and Indian Ocean sectors. Clim Dyn., 29, 63–71.CrossRefGoogle Scholar
Buckley, B., et al. (2018). Blue intensity from a tropical conifer’s annual rings for climate reconstruction: An ecophysiological perspective. Dendrochronologia, 50, 10–22.CrossRefGoogle Scholar
Carré, M., et al. (2005a). Stable isotopes and sclerochronology of the bivalve Mesodesma donacium: Potential application to Peruvian paleoceanographic reconstructions. Palaeogeogr. Palaeoclimatol. Palaeoecol., 228, 4e25.CrossRefGoogle Scholar
Carré, M., et al. (2005b). Strong El Niño events during the early Holocene: Stable isotope evidence from Peruvian sea-shells. Holocene, 15, 42–47.CrossRefGoogle Scholar
Carré, M., et al. (2012). Exploring errors in paleoclimate proxy reconstructions using Monte Carlo simulations: Paleotemperature from mollusk and coral geochemistry. Clim. Past., 8, 433–450.CrossRefGoogle Scholar
Carre, M., et al. (2014). Holocene history of ENSO variance and asymmetry in the eastern tropical Pacific. Science, 345(6200), 1045–1048.CrossRefGoogle ScholarPubMed
Carton, J. A., and Giese, B. S. (2008). A reanalysis of ocean climate using simple ocean data assimilation (SODA). Mon. Weather Rev., 2999–3017.CrossRefGoogle Scholar
Chadwick, M., et al. (2020). Analysing the timing of peak warming and minimum winter sea-ice extent in the Southern Ocean during MIS 5e. Quat. Sci. Rev., 229, 106134.CrossRefGoogle Scholar
Chappell, J., (1983). Evidence for smoothly falling sea-level relative to North Queensland, Australia, during the past 6,000 yr. Nature, 302, 406–408.CrossRefGoogle Scholar
Chappell, J., et al. (1996). Reconciliation of late Quaternary sea levels derived from coral terraces at Huon Peninsula with deep sea oxygen isotope records. Earth Planet. Sci. Lett., 141(1–4), 227–236. https://doi.org/10.1016/0012-821X(96)00062.CrossRefGoogle Scholar
Chen, T., et al. (2018). Coral-derived western Pacific tropical sea surface temperatures during the last millennium. Geophys. Res. Lett., 45(8), 3542–3549. https://doi.org/10.1002/2018GL077619.CrossRefGoogle Scholar
Chen, T., et al. (2019). Tropical sand cays as natural paleocyclone archives. Geophys. Res. Lett., 46, 9796–9803. https://doi.org/10.1029/2019GL084274.CrossRefGoogle Scholar
Cheng, L., et al. (2020). Improved estimates of changes in upper ocean salinity and the hydrological cycle. J. Clim., 33(23), 10357–10381. https://doi.org/10.1175/JCLI-D-20-0366.1.CrossRefGoogle Scholar
Clark, J. A., et al. (1978). Global changes in postglacial sea level: A numerical calculation. Quat. Res., 9, 265–278.CrossRefGoogle Scholar
Clark, P. U., et al. (2002). Sea-level fingerprinting as a direct test for the source of global meltwater pulse 1A. Science, 295(5564), 2438–2441.CrossRefGoogle Scholar
Cobb, K. M., et al. (2003). El Niño/Southern Oscillation and tropical Pacific climate during the last millennium. Nature, 424, 271–276.CrossRefGoogle ScholarPubMed
Cobb, K. M., et al. (2013). Highly variable El Niño- Southern Oscillation throughout the Holocene. Science, 339, 67–70.CrossRefGoogle ScholarPubMed
Cohen, A. L., et al. (1992). A Holocene sea surface temperature record in mollusk shells from the southwest African coast. Quat. Res., 38, 379–385.CrossRefGoogle Scholar
Cohen, A. L. and Tyson, P. D. (1995). Sea-surface temperature fluctuations during the Holocene off the south coast of Africa: Implications for terrestrial climate and rainfall. Holocene, 5(3), 304–312.CrossRefGoogle Scholar
Cole, J. E. and Fairbanks, R. G. (1990). The Southern Oscillation in the δ18O of corals from Tarawa Atoll. Paleoceanogr., 5(5), 669–683.CrossRefGoogle Scholar
Cole, J. E., et al. (2000). Tropical Pacific forcing of decadal SST variability in the Western Indian Ocean over the past two centuries. Science, 287(5453), 617–619.CrossRefGoogle ScholarPubMed
Cole, J. E. (2003). Holocene coral records: Windows on tropical climate variability. In Mackay, A., Battarbee, R., Birks, J. and Oldfield, F. (eds.), Global Change in the Holocene. Edward Arnold, London, pp. 168–184.Google Scholar
Compo, G. P., et al. (2011). The twentieth century reanalysis project. Q. J. Roy. Meteorol. Soc., 137(654), 1–28.CrossRefGoogle Scholar
Conroy, J. L., et al. (2017). Spatiotemporal variability in the δ18O-salinity relationship of seawater across the tropical Pacific Ocean. Paleoceanogr., 32(5), 484–497.CrossRefGoogle Scholar
Cook, E. R. and Kairiukstis, L. A. (ed). (1990). Methods of Dendrochronology – Applications in the Environmental Sciences. The Netherlands, Kluwer.CrossRefGoogle Scholar
Cook, E. R. and Peters, K. (1997). Calculating unbiased tree-ring indices for the study of climatic and environmental change. Holocene, 7(3), 361–370.CrossRefGoogle Scholar
Cook, E. R., et al. (2006). Millennia-long tree-ring records from Tasmania and New Zealand: A basis for modelling climate variability and forcing, past, present and future. J. Quat. Sci., 21, 689–699.CrossRefGoogle Scholar
Cook, E. R., et al. (2010). Asian monsoon failure and megadrought during the last millennium Science, 328, 486–489.CrossRefGoogle ScholarPubMed
Corrège, T. (2006). Sea surface temperature and salinity reconstruction from coral geochemical tracers. Palaeogeogr. Palaeoclimatol. Palaeoecol., 232, 408–428.CrossRefGoogle Scholar
Corrège, T., et al., (2001). Little Ice Age sea surface temperature variability in the southwest tropical Pacific. Geophys. Res. Lett., 28, 3477–3480.CrossRefGoogle Scholar
Cortese, G., et al. (2013). Southwest Pacific Ocean response to a warmer world: Insights from Marine Isotope Stage 5e. Paleoceanogr., 28(3), 585–598.CrossRefGoogle Scholar
D’Arrigo, R., et al. (2006). Monsoon drought over Java, Indonesia, during the past two centuries. Geophys. Res. Lett., 33, 4, L04709. https://doi.org/10.1029/2005GL025465.Google Scholar
D’Olivo, J. and McCulloch, M. (2017). Response of coral calcification and calcifying fluid composition to thermally induced bleaching stress. Scientific Rep., 7(1), 1–15.Google ScholarPubMed
D’Olivo, J. P., et al. (2018). A universal multi-trace element calibration for reconstructing sea surface temperatures from long-lived Porites corals: Removing ‘vital-effects’. Geochim. Cosmochim. Acta, 239, 109–135. https://doi.org/10.1016/j.gca.2018.07.035.Google Scholar
Dassié, E. P., et al. (2014). A Fiji multi-coral δ18O composite approach to obtaining a more accurate reconstruction of the last two-centuries of the ocean-climate variability in the South Pacific Convergence Zone region: Fiji coral network and SPCZ variability. Paleoceanogr., 29(12), 1196–1213.CrossRefGoogle Scholar
De Deckker, P. D. (2020). Airborne dust traffic from Australia in modern and Late Quaternary times. Glob. Planet. Change, 184, 103056. https://doi.org/10.1016/j.gloplacha.2019.103056.CrossRefGoogle Scholar
De Ridder, F., et al. (2004). Decoding nonlinear growth rates in biogenic environmental archives. Geochem., Geophys., Geosyst., 5, 12.CrossRefGoogle Scholar
de Villiers, S., et al. (1995). Biological controls on coral Sr/Ca and δ18O reconstructions of sea surface temperatures. Science, 269, 1247–1249.CrossRefGoogle Scholar
DeCarlo, T. M., et al. (2016). Coral Sr-U thermometry. Paleoceanogr., 31, 626–638, https://doi.org/10.1002/2015PA002908.CrossRefGoogle Scholar
Dee, S., et al. (2015). PRYSM: An open-source framework for PRoxY System Modeling, with applications to oxygen-isotope systems. J. Adv. Mod. Earth Syst., 7, 1220–1247.Google Scholar
Dee, S. K., et al. (2020). No consistent ENSO response to volcanic forcing over the last millennium. Science, 367, 6485. https://doi.org/10.1126/science.aax2000.CrossRefGoogle ScholarPubMed
DeFlorio, M., et al. (2015). Interannual modulation of subtropical Atlantic boreal summer dust variability by ENSO, Clim. Dyn., 1–15, https://doi.org/10.1007/s00382-015-2600-7.Google Scholar
Delcroix, T. and McPhaden, M. (2002). Interannual sea surface salinity and temperature changes in the western Pacific warm pool during 1992–2000. J. Geophys. Res., 107(C12), 8002. https://doi.org/10.1029/2001JC000862.Google Scholar
Delcroix, T., et al. (2007). Decadal variations and trends in tropical Pacific sea surface salinity since 1970. J. Geophys. Res., 112, C03012. https://doi.org/10.1029/2006JC003801.Google Scholar
Delcroix, T., et al. (2011). A gridded sea surface salinity data set for the tropical Pacific with sample applications (1950–2008). Deep-Sea Res., I, 58, 38–48.Google Scholar
DeLong, K. L., et al. (2013). Improving coral-base paleoclimate reconstructions by replicating 350 years of coral Sr/Ca variations. Palaeogeogr. Palaeoclimatol. Palaeoecol., 373, 6–24.CrossRefGoogle Scholar
Druffel, E. R. M. and Griffin, S. (1993). Large variations of surface ocean radiocarbon: Evidence of circulation changes in the southwestern Pacific. J. Geophys. Res., 98(C11), 20249–20259.Google Scholar
Druffel, E. M. and Griffin, S. (1999). Variability of surface ocean radiocarbon and stable isotopes in the southwestern Pacific. J. Geophys. Res., 104, 23607–23613.Google Scholar
Druffel, E. R. M., et al. (2007). Oceanic climate and circulation changes during the past four centuries from radiocarbon in corals. Geophys. Res. Lett., 34, L09601.CrossRefGoogle Scholar
Dunbar, R. B. and Cole, J. E. (1999). Annual records of tropical systems (ARTS). PAGES Workshop Report, Series 99–1.Google Scholar
Dutton, A. (2015). Uranium-thorium dating. In Shennan, I., et al., (eds.), Handbook of Sea-Level Research. John Wiley & Sons Ltd. 386–403.Google Scholar
Dutton, A., et al. (2017). Data reporting standards for publication of U-series data for geochronology and timescale assessment in the earth sciences. Quat. Geochronol., 39, 142–149. https://doi.org/10.1016/j.quageo.2017.03.001, 2017.CrossRefGoogle Scholar
Edwards, R. L., et al. (1987). 238U–234U–230Th–232Th systematics and the precise measurement of time over the past 500,000 years. Earth Planet. Sci. Lett., 81(2–3), 175–192.Google Scholar
Emile-Geay, J., et al. (2016). Links between tropical Pacific seasonal, interannual and orbital variability during the Holocene. Nature Geosci., 9, 168–173, https://doi.org/10.1038/NGEO2608.CrossRefGoogle Scholar
Esat, T. M., et al. (1999). Rapid fluctuations in sea level recorded during the penultimate deglaciation. Science, 283, 197–201.CrossRefGoogle ScholarPubMed
Etayo-Cadavid, M. F., et al. (2013). Marine radiocarbon reservoir age variation in Donax obesulus shells from northern Peru: Late Holocene evidence for extended El Niño. Geol., 41(5), 599–602.CrossRefGoogle Scholar
Evans, M. N. and Schrag, D. P. (2004). A stable isotope-based approach to tropical dendroclimatology. Geochim. Cosmochim. Acta, 68, 3295–3305.CrossRefGoogle Scholar
Evans, M. N., et al. (2013). Applications of proxy system modeling in high resolution paleoclimatology. Quat. Sci. Rev., 76, 16–28.CrossRefGoogle Scholar
Fairbanks, R. G. (1989). A 17,000-year glacio-eustatic sea level record: Influence of glacial melting rates on the Younger Dryas event and deep-ocean circulation. Nature, 342, 637–642.CrossRefGoogle Scholar
Fairbanks, R. G., et al. (1997). Evaluating climate indices and their geochemical proxies measured in corals. Coral Reefs, 16(5), S93–S100.CrossRefGoogle Scholar
Fairchild, I. J. and Baker, A. (2012). Speleothem Science: From Process to Past Environments. Wiley.CrossRefGoogle Scholar
Felis, T. (2020). Extending the instrumental record of ocean-atmosphere variability into the last inter- glacial using tropical corals. Oceanography 33, 2, https://doi.org/10.5670/oceanog.2020.209.CrossRefGoogle Scholar
Felis, T. and Patzold, J. (2004). Climate reconstructions from annually banded corals. In Shiyomi, M., et al. (eds.), Global Environmental Change in the Ocean and on Land, pp. 205–227. TERRAPUB, Tokyo.Google Scholar
Felis, T., et al. (2009). Subtropical coral reveals abrupt early-twentieth-century freshening in the western North Pacific Ocean. Geology, 37, 527–530.CrossRefGoogle Scholar
Felis, T., et al. (2012). Pronounced interannual variability in tropical South Pacific temperatures during Heinrich Stadial 1. Nature Commun., 3, 965. https://doi.org/10.1038/ncomms1973.CrossRefGoogle ScholarPubMed
Feng, M., et al. (2013). La Niña forces unprecedented Leeuwin Current warming in 2011. Scientific Rep., 2, 1277. https://doi.org/10.1038/srep01277.Google Scholar
Flood, P. G. (1986). Sensitivity of coral cays to climatic variations, southern Great Barrier Reel Australia. Coral Reefs, 5, 13–18.CrossRefGoogle Scholar
Flora, C. J., et al. (2009). Microatoll edge to ENSO annulus growth suggests sea level change. Atoll Res. Bull., 571, 10pp.CrossRefGoogle Scholar
Fowler, A. M., et al. (2012) Multi-centennial ENSO insights from New Zealand forest giants. Nat. Clim. Chang., 2(3), 172–176.Google Scholar
Freund, M. B., et al. (2019). Higher frequency of Central Pacific El Niño events in recent decades relative to past centuries. Nat. Geosci., https://doi.org/10.1038/s41561-019-0353-3.CrossRefGoogle Scholar
Gagan, M. K., et al. (2000). New views of tropical paleoclimates from corals. Quat. Sci. Rev., 19, 45–64.CrossRefGoogle Scholar
Garreaud, R. D. (2018). Short Communication. A plausible atmospheric trigger for the 2017 coastal El Niño. Int. J. Climatol. 38(Suppl.1), 1296–1302. https://doi.org/10.1002/joc.5426.CrossRefGoogle Scholar
Garrison, V. H., et al. (2003). African and Asian dust: From desert soils to coral reefs. Biosci., 53(5), 469–480.CrossRefGoogle Scholar
Gelhar, L. W. and Wilson, J. L. (1974). Ground-water quality modeling, Groundwater, 12(6), 399–408.CrossRefGoogle Scholar
Genty, D., et al. (1998). Calculation of past dead carbon proportion and variability by the comparison of AMS 14C and TIMS U/Th ages on two Holocene stalagmites. Radiocarbon, 41, 251–270.Google Scholar
Genty, D. and Moreno, A. (2021). Air-ground interface: Reconstruction of paleoclimates using speleothem. In Ramstein, G. et al. (eds.), Paleoclimatology, Frontiers in Earth Sciences, Springer Nature, Switzerland, pp. 169–177. https://doi.org/10.1007/978-3-030-24982-3_14.Google Scholar
Gill, E. D. and Amin, B. S. (1975). Interpretation of 7.5 and 4 metre Last Interglacial shore platforms in southeast Australia. Search, 6, 394–396.Google Scholar
Giry, C., et al. (2010). Geochemistry and skeletal structure of Diploria strigosa, implications for coral-based climate reconstruction. Palaeogeogr. Palaeoclimatol. Palaeoecol., 298, 378–387. https://doi.org/10.1016/j.palaeo.2010.10.022.CrossRefGoogle Scholar
Gläser, G., et al. (2015). The transatlantic dust transport from North Africato the Americas – Its characteristics and source regions. J. Geophys. Res. Atmos., 120, 11231–11252. https://doi.org/10.1002/2015JD023792.CrossRefGoogle Scholar
Good, S. A., et al. (2013). EN4: Quality controlled ocean temperature and salinity profiles and monthly objective analyses with uncertainty estimates. J. Geophys. Res. Oceans, 118, 6704–6716. https://doi.org/10.1002/2013JC009067.CrossRefGoogle Scholar
Goodkin, N. F., et al. (2005). Record of Little Ice Age sea surface temperatures at Bermuda using a growth-dependent calibration of coral Sr/Ca. Paleoceanogr., 20(4), PA4016.CrossRefGoogle Scholar
Goodkin, N. F., et al. (2019). East Asian Monsoon variability since the sixteenth century. Geophys. Res. Lett., 46, 4790–4798. https://doi.org/10.1029/2019GL081939.CrossRefGoogle Scholar
Goodwin, I. D. (2003). Unraveling climate influences on late Holocene sea-level and coastal evolution. In Mackay, A., et al. (eds.), Global Change in the Holocene. Edward Arnold, London, pp. 406–421.Google Scholar
Goodwin, I. D. and Harvey, N. (2008). Sub-tropical sea- level history from coral microatolls in the Southern Cook Islands, since 300 AD. Mar. Geol., 253(1–2), 14–25.CrossRefGoogle Scholar
Goodwin, I. D. and Howard, W. R. (2012). Evidence of environmental change from the marine realm. Chapter 9. In Matthews, J. A. (ed), The Handbook of Environmental Change. Sage Publications, London, pp. 181–210.Google Scholar
Goodwin, I. D., et al. (2023). Robbins Island: The index site for regional Last Interglacial sea level, wave climate and the subtropical ridge around Bass Strait, Australia. Quat. Sc. Rev., 305, 107996. https://doi.org/10.1016/j.quascirev.2023.107996.CrossRefGoogle Scholar
Goodwin, I. D. and Mitrovica, J. X. M. (in prep 2025). Fingerprinting Last Interglacial sea-level rise from Australia reveals minimal East Antarctic ice mass loss.Google Scholar
Gordillo, S., et al. (2014). Mollusk shells as bio-geo-archives: Evaluating environmental changes during the Quaternary. Springer Briefs in Earth System Sciences South America and the Southern Hemisphere, 80pp. https://doi.org/10.1007/978-3-319-03476-8_9.CrossRefGoogle Scholar
Gourlay, M. R. (1983) Accretion and erosion of coral cays and some implications for the management of Marine Parks. In Baker, J. T., et al. (eds.), Proceedings of the Inaugural Great Barrier Reef Conference. JCU Press, Townsville, pp. 475–482.Google Scholar
Griffiths, M. L., et al. (2010). Evidence for Holocene changes in Australian–Indonesian monsoon rainfall from stalagmite trace element and stable isotope ratios. Earth. Planet. Sci. Lett., 292, 27–38.CrossRefGoogle Scholar
Grossman, E. E., et al. (1998). The Holocene sea-level highstand in the equatorial Pacific: Analysis of the insular paleosea-level database. Coral Reefs, 17, 309–327.CrossRefGoogle Scholar
Grottoli, A. G. (2000). Stable carbon isotopes (δ13C) in coral skeletons. Oceanogr., 13, 93–97.CrossRefGoogle Scholar
Grottoli, A. G. and Eakin, M. (2007). A review of coral δ18O and Δ14C proxy records. Earth Sci. Rev., 81, 67–91.CrossRefGoogle Scholar
Grottoli, A. G., et al. (2013). High resolution coral Cd measurements using LA-ICP-MS and ID-ICP-MS: Calibration and interpretation. Chem. Geol., 356, 151–159.CrossRefGoogle Scholar
Grove, C. A., et al. (2010). River runoff reconstructions from novel spectral luminescence scanning of massive coral skeletons. Coral Reefs, 29, 579–591.CrossRefGoogle Scholar
Grove, C. A., et al. (2012).Spatial linkages between coral proxies of terrestrial runoff across a large embayment in Madagascar. Biogeosci., 9, 3063–3081. https://doi.org/10.5194/bg-9-3063-2012.CrossRefGoogle Scholar
Grove, C. A., et al. (2013). Madagascar corals reveal a multidecadal signature of rainfall and river runoff since 1708. Clim. Past, 9, 641–656, https://doi.org/10.5194/cp-9-641-2013.CrossRefGoogle Scholar
Hallmann, N., et al. (2018). Ice volume and climate changes from a 6000 year sea-level record in French Polynesia. Nature Commun., 9, 285. https://doi.org/10.1038/s41467-017-02695-7.CrossRefGoogle ScholarPubMed
Hamylton, S. and Puotinen, M. (2015). A meta-analysis of reef island response to environmental change on the Great Barrier Reef. Earth Surf. Process. Landforms, 40, 1006–1016.CrossRefGoogle Scholar
Hathorne, E. C., et al. (2011). Laser ablation ICP-MS screening of corals for diagenetically affected areas applied to Tahiti corals from the last deglaciation. Geochim. Cosmochim. Acta, 75(6), 1490–1506.CrossRefGoogle Scholar
Hathorne, E. C., et al. (2013a). Lithium in the aragonite skeletons of massive Porites corals: A new tool to reconstruct tropical sea surface temperatures. Paleoceanogr., 28, 143–152.CrossRefGoogle Scholar
Hathorne, E. C., et al. (2013b). Interlaboratory study for coral Sr/Ca and other element/Ca ratio measurements. Geochem. Geophys, Geosyst., 14, 3730–3750.CrossRefGoogle Scholar
Hay, C., et al. (2014). The sea-level fingerprints of ice-sheet collapse during interglacial periods. Quat. Sci. Rev., 87, 60–69. https://doi.org/10.1016/j.quascirev.2013.12.022,2014.CrossRefGoogle Scholar
Hearty, P. J. and Neumann, A. C. (2001). Rapid sea level and climate change at the close of the Last Interglaciation (MIS5e): Evidence from the Bahama Islands. Quat. Sci. Rev., 20, 1881–1895.CrossRefGoogle Scholar
Hellstrom, J. (2006). U-Th dating of speleothems with high initial Th-230 using stratigraphical constraint. Quat. Geochron., 1, 289–295.CrossRefGoogle Scholar
Hellstrom, J. and Pickering, R. (2015). Recent advances and future prospects of the U–Th and U–Pb chronometers applicable to archaeology. J. Archaeol. Sci., 56, 32–40.CrossRefGoogle Scholar
Hendy, E. J., et al. (2002). Abrupt decrease in tropical Pacific sea surface salinity at end of Little Ice Age. Science, 295, 1511–1514.CrossRefGoogle ScholarPubMed
Hendy, E. J., et al. (2003). Chronological control of coral records using luminescent lines and evidence for non-stationary ENSO teleconnections in northeast Australia. Holocene, 13, 187–199, https://doi.org/10.1191/0959683603hl606rp.CrossRefGoogle Scholar
Hendy, E. J., et al. (2007). Impact of skeletal dissolution and secondary aragonite on trace element and isotopic climate proxies in Porites corals. Paleoceanogr., 22(4), PA4101.CrossRefGoogle Scholar
Hereid, K. A., et al. (2013). Assessing spatial variability in El Niño–Southern Oscillation event detection skill using coral geochemistry. Paleoceanogr., 28(1), 14–23.CrossRefGoogle Scholar
Hersbach, H., et al. (2020). The ERA5 global reanalysis. Quart. J. Roy. Meteorol. Soc., 146, 1999–2049.CrossRefGoogle Scholar
Hetzinger, S., et al. (2008). Caribbean coral tracks Atlantic Multidecadal Oscillation and past hurricane activity. Geology, 36(1), 11–14.CrossRefGoogle Scholar
Hetzinger, S., et al. (2016). A change in coral extension rates and stable isotopes after El Niño- induced coral bleaching and regional stress events. Sci. Rep., 6, 32879. https://doi.org/10.1038/srep32879.CrossRefGoogle ScholarPubMed
Hibbert, F. D., et al. (2016). Coral indicators of past sea-level change: A global repository of U-series dated benchmarks. Quat. Sci. Rev., 145, 1–56. https://doi.org/10.1016/j.quascirev.2016.04.019.CrossRefGoogle Scholar
Higgins, P. A., et al. (2020). One thousand three hundred years of variability in the position of the South Pacific Convergence Zone. Geophys. Res. Lett., 47, e2020GL088238. https://doi.org/10.1029/2020GL088238.CrossRefGoogle Scholar
Hopley, D. (1982). Geomorphology of the Great Barrier Reef: Quaternary Development of Coral Reefs. John Wiley-Interscience, New York.Google Scholar
Hua, Q., et al. (2005). Radiocarbon in corals from the Cocos (Keeling) Islands and implications for Indian Ocean circulation. Geophys. Res. Lett., 32(21), l21602.CrossRefGoogle Scholar
Hughes, M. K. (2011). Dendroclimatology in high-resolution paleoclimatology. In Hughes, M. K., et al. (eds.), Dendrochronology: Progress and Prospects, Developments in Paleoenvironmental Research, vol. 11. Springer, New York, pp. 17–34, 365pp.CrossRefGoogle Scholar
Humanes-Fuente, V., et al. (2020). Two centuries of hydroclimatic variability reconstructed from tree-ring records over the Amazonian Andes of Peru. J. Geophys. Res: Atmospheres, 125(18), e2020JD032565. https://doi.org/10.1029/2020JD032565.CrossRefGoogle Scholar
Huyghe, D., et al. (2022). Clumped isotopes in modern marine bivalves. Geochim. Cosmochim. Acta, 316, 41–58.CrossRefGoogle Scholar
Isdale, P. (1984). Fluorescent bands in massive corals record centuries of coastal rainfall. Nature, 310, 578–579.CrossRefGoogle Scholar
Isdale, P. J., et al. (1998). Palaeohydrological variation in a tropical river catchment: A reconstruction using fluorescent bands in corals of the Great Barrier Reef, Australia. Holocene, 8, 1–8.CrossRefGoogle Scholar
Jones, D. S. (1980). Annual cycle of shell growth increment formation in two continental shelf bivalves and its paleoecologic significance. Paleobiol., 6, 331–340.CrossRefGoogle Scholar
Jones, D. S. (1983). Sclerochronology: Reading the record of the molluscan shell. Am. Sci., 71, 384–391.Google Scholar
Jones, D. S., et al. (1989). Sclerochronological records of temperature and growth from shells of Mercenaria mercenaria from Narragansett Bay, Rhode Island. Mar. Biol., 102, 225–234.CrossRefGoogle Scholar
Jones, K. B., et al. (2009). Upwelling signals in radiocarbon from early 20th-century Peruvian bay scallop (Argopecten purpuratus) shells. Quat. Res., 72, 452–456.CrossRefGoogle Scholar
Kataoka, T., et al. (2014). On the Ningaloo Niño/Niña. Clim. Dyn., 43(5–6), 1463–1482.CrossRefGoogle Scholar
Kataoka, T., et al. (2018). Can Ningaloo Niño/Niña develop without El Niño–Southern Oscillation? Geophys. Res. Lett., 45, 7040–7048.CrossRefGoogle Scholar
Kayanne, H., et al. (2006). Indian Ocean Dipole index recorded in Kenyan coral annual density bands. Geophys. Res. Lett., 33(19), L19709. https://doi.org/10.1029/2006GL027168.CrossRefGoogle Scholar
Kench, P. S., et al. (2009). Holocene reef growth in the Maldives: Evidence of a mid-Holocene sea level highstand in the central Indian Ocean. Geology, 37, 455–458.CrossRefGoogle Scholar
Kench, P. S., et al. (2018). Storm-deposited coral blocks: A mechanism of island genesis, Tutaga island, Funafuti atoll, Tuvalu. Geology, 46(10), 915–918.CrossRefGoogle Scholar
Kench, P. S., et al. (2020). Climate-forced sea-level lowstands in the Indian Ocean during the last two millennia. Nature Geosci., 13, 61–64, https://doi.org/10.1038/s41561-019-0503-7.CrossRefGoogle Scholar
King, T. M., et al. (2018). Large-scale intrusion of Circumpolar Deep Water on Antarctic margin recorded by stylasterid corals. Paleoceanogr. Paleoclimatol., 33, 1306–1321. https://doi.org/10.1029/2018PA003439.CrossRefGoogle Scholar
Kuhnert, H., et al. (1999). A 200-year coral stable oxygen isotope record from a high- latitude reef off Western Australia. Coral Reefs, 18, 1–12.CrossRefGoogle Scholar
Kuhnert, H., et al. (2000). Monitoring climate variability over the past 116 years in coral oxygen isotopes from Ningaloo Reef, Western Australia. Int. J. Earth Sci., 88, 725–732.CrossRefGoogle Scholar
Ladd, S. N., et al. (2021). Leaf wax hydrogen isotope s as a hydroclimate proxy in the tropical Pacific. J. Geophys. Res. Biogeosciences, 126(3), e2020JG005891. https://doi.org/10.1029/2020jg005891.CrossRefGoogle Scholar
Lambeck, K. (2002). Sea-level change from mid Holocene to recent time: An Australian example with global implications. In Mitrovica, J. X. and Vermeersen, B. L. A. (eds.), Ice Sheets, Sea Level and the Dynamic Earth. Geodynamics Series 29, American Geophysical Union, Washington, DC, pp. 33–50.Google Scholar
Lambeck, K. and Chappell, J. (2001). Sea level change through the last glacial cycle. Science, 292(5517), 679–686.CrossRefGoogle ScholarPubMed
Lambeck, K., et al. (2002). Links between climate and sea levels for the past three million years. Nature, 419, 199–205.CrossRefGoogle ScholarPubMed
LaVigne, M., et al. (2008). Skeletal P/Ca tracks upwelling in Gulf of Panama coral: Evidence for a new seawater phosphate proxy. Geophys. Res. Lett., 35(5), L05604.CrossRefGoogle Scholar
LaVigne, M., et al. (2010). Coral skeleton P/Ca proxy for seawater phosphate: Multi-colony calibration with a contemporaneous seawater phosphate record. Geochim. Cosmochim. Acta, 74(4), 1282–1293. https://doi.org/10.1016/j.gca.2009.11.002.CrossRefGoogle Scholar
LaVigne, M., et al. (2016). Multi-colony calibrations of coral Ba/Ca with a contemporaneous in situ seawater barium record. Geochim. Cosmochim. Acta, 179, 203–216. https://doi.org/10.1016/j.gca.2015.12.038.CrossRefGoogle Scholar
Lawman, A. E., et al. (2020). Developing a coral proxy system model to compare coral and climate model estimates of changes in paleo-ENSO variability. Paleoceanogr. Paleoclimatol., 35(7), e2019PA003836. https://doi.org/10.1029/2019PA003836.CrossRefGoogle Scholar
Lécuyer, C., et al. (2004). Stable isotope fractionation between mollusc shells and marine waters from Martinique Island. Chem. Geol., 213, 293–305.CrossRefGoogle Scholar
Leinen, M. and Sarnthein, M., (eds.). (1987). Paleoclimatology and Paleometeorology: Modern and Past Patterns of Global Atmospheric Transport. NATO ASI Series C. Mathematical and Physical Sciences, 282, Kluwer Academic Publishers, Dordrecht, The Netherlands, 909pp.Google Scholar
Leonard, N. D., et al. (2016). Holocene sea level instability in the southern Great Barrier Reef, Australia: High-precision U–Th dating of fossil microatolls. Coral Reefs, 35, 625–639.CrossRefGoogle Scholar
Leonard, N. D., et al. (2019). New evidence for ‘far-field’ Holocene sea level oscillations and links to global climate records. Earth Planet. Sci. Lett., 487, 67–73.Google Scholar
Linse, K., et al. (2006). Biodiversity and biogeography of Antarctic and sub-Antarctic Mollusca. Deep-Sea Res. II, 53, 985–1008.Google Scholar
Linsley, B. K., et al. (2000a). Decadal sea surface temperature variability in the subtropical Pacific from 1726 to 1997 A. D. Science, 290, 1145–1148.CrossRefGoogle Scholar
Linsley, B. K., et al. (2000b). ENSO and decadal-scale climate variability at 10°N in the Eastern Pacific from 1893 to 1994: A coral- based reconstruction from Clipperton Atoll. Paleoceanogr., 15, 322–335.CrossRefGoogle Scholar
Linsley, B. K., et al. (2004). Geochemical evidence from corals for changes in the amplitude and spatial pattern of South Pacific interdecadal climate variability over the last 300 years. Clim. Dyn., 22, 1, 1–11.Google Scholar
Linsley, B. K., et al. (2006). Tracking the extent of the South Pacific Convergence Zone since the early 1960s. Geochem. Geophys. Geosyst., 7(4), Q05003.CrossRefGoogle Scholar
Linsley, B. K., et al. (2008). Interdecadal-decadal climate variability from multi-coral oxygen isotope records in the South Pacific Convergence Zone region since 1650 A.D. Paleoceanogr., 23(2), PA2219.CrossRefGoogle Scholar
Linsley, B. K., et al. (2015). Decadal changes in South Pacific sea surface temperatures and the relationship to the Pacific decadal oscillation and upper ocean heat content. Geophys. Res. Lett., 42(7), 2358–2366.CrossRefGoogle Scholar
Linsley, B. K., et al. (2017). Abrupt Northward Shift of SPCZ position in the late-1920s indicates coordinated Atlantic and Pacific ITCZ change. Past Global Changes Magazine, 25, https://doi.org/10.22498/pages.25.1.52(2017).Google Scholar
Linsley, B. K., et al. (2019). Coral carbon isotope sensitivity to growth rate and water depth with paleo-sea level implications. Nature Commun., 10, 2056. https://doi.org/10.1038/s41467-019-10054-x.Google ScholarPubMed
Loftus, E., et al. (2017). A late quaternary record of seasonal sea surface temperatures off southern Africa. Quat. Sci. Rev., 171(2017), 73–84.CrossRefGoogle Scholar
Lohmann, G. and Schöne, B. R. (2013). Climate signatures on decadal to interdecadal time scales as obtained from mollusk shells (Arctica islandica) from Iceland. Palaeogeogr. Palaeoclimatol. Palaeoecol., 373, 152–162.CrossRefGoogle Scholar
Lough, J. M. (2007). Tropical river flow and rainfall reconstructions from coral luminescence: Great Barrier Reef, Australia. Paleoceanogr., 22, PA2218.CrossRefGoogle Scholar
Lough, J. M. (2010). Climate records from corals: Climate records from corals. WIREs Clim. Change, 1(3), 318–331. https://doi.org/10.1002/wcc.39.CrossRefGoogle Scholar
Lough, J. M. (2011a). Great Barrier Reef coral luminescence reveals rainfall variability over northeastern Australia since the 17th century. Paleoceanogr., 26, PA2201. https://doi.org/10.1029/2010PA002050.CrossRefGoogle Scholar
Lough, J. M. (2011b). Measured coral luminescence as a freshwater proxy: Comparison with visual indices and a potential age artefact, Coral Reefs, 30, 169–182, https://doi.org/10.1007/s00338-010-0688-0.CrossRefGoogle Scholar
Lough, J. M. and Barnes, D. J. (1997). Several centuries of variation in skeletal extension, density and calcification in massive Porites colonies from the Great Barrier Reef: A proxy for seawater temperature and a background of variability against which to identify unnatural change. J. Exp. Mar. Biol. Ecol., 211, 29–67.CrossRefGoogle Scholar
Lough, J. M. and Barnes, D. J. (2000). Environmental controls on growth of the massive coral Porites. J. Exp. Mar. Biol. Ecol., 245, 225–243.CrossRefGoogle ScholarPubMed
Lough, J. M., & Cooper, T. F. (2011). New insights from coral growth band studies in an era of rapid environmental change. Earth-Sci. Rev., 108(3–4), 170–184.CrossRefGoogle Scholar
Lough, J. M., et al. (2014). Evidence for suppressed mid-Holocene northeastern Australian monsoon variability from coral luminescence. Paleoceanogr., 29, 581–594. https://doi.org/10.1002/2014PA002630.CrossRefGoogle Scholar
Luetscher, M., et al. (2015). North Atlantic storm track changes during the Last Glacial Maximum recorded by Alpine speleothems. Nature Commun., 6, 7344. https://doi.org/10.1038/ncomms7344.CrossRefGoogle ScholarPubMed
Majewski, J. M., et al. (2022). Extending instrumental sea-level records using coral microatolls, an example from Southeast Asia. Geophys. Res. Lett., 49, e2021GL095710. https://doi.org/10.1029/2021GL095710.CrossRefGoogle Scholar
Malakoff, D. (2003). Cool corals become hot topic. Science, 299, 195.CrossRefGoogle ScholarPubMed
Maloney, A. E., et al. (2022). Contrasting Common Era climate and hydrology sensitivities from paired lake sediment dinosterol hydrogen isotope records in the South Pacific Convergence Zone. Quat. Sci. Rev., 281, 107421.CrossRefGoogle Scholar
Mann, M. E., et al. (1998). Global-scale temperature patterns and climate forcing over the past six centuries, Nature, 392, 779–787.CrossRefGoogle Scholar
Marshall, J. F. and Thom, B. G. (1976). The sea level in the last interglacial. Nature, 263, 120–121.CrossRefGoogle Scholar
Marshall, J. F. and McCulloch, M. T. (2002). An assessment of the Sr/Ca ratio in shallow water hermatypic corals as a proxy for sea surface temperature. Geochim. Cosmochim. Acta, 66,(18), 3263–3280.CrossRefGoogle Scholar
Marshall, A. G., et al. (2015). Initiation and amplification of the Ningaloo Niño. Clim. Dyn., 45, 2367–2385.CrossRefGoogle Scholar
Marx, S. K., et al. (2009). Long-range dust transport from eastern Australia: A proxy for Holocene aridity and ENSO-type climate variability. Earth Planet. Sci. Lett., 282, 167–177.CrossRefGoogle Scholar
Massel, S. R. (1996). Ocean Surface Waves: Their Physics and Prediction. Advanced Series on Ocean Engineering, Volume 11, World Scientific Publishing, Singapore, 491pp.Google Scholar
Maupin, C. R., et al. (2014). Persistent decadal-scale rainfall variability in the tropical South Pacific Convergence Zone through the past six centuries. Climate of the Past, 10(4), pp. 1319.CrossRefGoogle Scholar
McCulloch, M. T., et al. (1999). Coral record of equatorial sea-surface temperatures during the penultimate deglaciation at Huon Peninsula. Science, 283, 202–204.CrossRefGoogle ScholarPubMed
McCulloch, M. T. and Esat, T. (2000). The coral record of Last Interglacial sea levels and sea surface temperatures. Chem. Geol., 169, 107–129.CrossRefGoogle Scholar
McCulloch, M. T., et al. (2003). Coral record of river runoff and human impacts on the Inner Great Barrier Reef of Australia. Nature, 427, 727–730.Google Scholar
McCulloch, M. T. and Mortimer, G. E. (2008). Applications of the 238U-230Th decay series to dating of fossil and modern corals using MC-ICPMS. Aust. J. Earth Sci., 55(6), 955–965.CrossRefGoogle Scholar
McCulloch, M. T., et al. (2024). 300 years of sclerosponge thermometry shows global warming has exceeded 1.5°C. Nature. Clim. Change, 14, 171–177, https://doi.org/10.1038/s41558-023-01919-7.CrossRefGoogle Scholar
McDermott, F. (2004). Paleo-climate reconstruction from stable isotope variations in speleothems: A review. Quat. Sci. Rev., 23, 901–918.CrossRefGoogle Scholar
McGregor, H. V. and Gagan, M. K. (2003). Diagenesis and geochemistry of Porites corals from Papua New Guinea. Geochim. Cosmochim. Acta, 67(12), 2147–2156. https://doi.org/10.1016/S0016-7037(02)01050-5.CrossRefGoogle Scholar
McGregor, H. V., et al. (2011). Environmental control of the oxygen isotope composition of Porites coral microatolls. Geochim. Cosmochim. Acta, 75(14), 3930–3944. https://doi.org/10.1016/j.gca.2011.04.017.CrossRefGoogle Scholar
Meltzner, A. J. and Woodroffe, C. D. (2015). Chapter 8: Coral microatolls. In Shennan, I., et al. (eds.), Handbook of Sea Level Research. John Wiley and Sons, Chichester, pp. 125–145.Google Scholar
Mitrovica, J. X. and Milne, G. A. (2002). On the origin of late Holocene sea-level highstands within equatorial ocean basins. Quat. Sci. Rev., 21, 2179–2190.CrossRefGoogle Scholar
Mitsuguchi, T., et al. (1996). Mg/Ca thermometry in coral skeletons. Science, 274, 961–963.CrossRefGoogle ScholarPubMed
Mitsuguchi, T., et al. (2003). Mg/Ca and Sr/Ca ratios of Porites coral skeleton: Evaluation of the effect of skeletal growth rate. Coral Reefs, 22, 381–388.CrossRefGoogle Scholar
Montalbetti, E., et al. (2021) Manganese benefits heat-stressed corals at the cellular level. Front. Mar. Sci., 8, 681119. https://doi.org/10.3389/fmars.2021.681119.CrossRefGoogle Scholar
Moseley, M. E., et al. (1992). Space shuttle imagery of recent catastrophic change along the arid Andean coast. In Lewis Johnson, L. and Stright, M. (eds.), Paleoshorelines and Prehistory: An Investigation of Method. Boca Raton, CRC Press, pp. 215–235.Google Scholar
Murray Roberts, J., et al. (2022). Cold-Water Corals: The Biology and Geology of Deep-Sea Coral Habitats. Cambridge University Press, Cambridge, 334pp.Google Scholar
Murray-Wallace, C. V., et al. (2000). Palaeoclimatic implications of the occurrence of the arcoid bivalve Anadara trapezia (Deshayes) in the Quaternary of Australasia. Quat. Sci. Rev., 19, 559–590.CrossRefGoogle Scholar
Murray-Wallace, C. V., et al. (2016). Last interglacial (MIS 5e) sea-level determined from a tectonically stable, far-field location, Eyre Peninsula, southern Australia. Aust. J. Earth Sci., 63, 611–630.CrossRefGoogle Scholar
Neumann, A. C. and McIntyre, I. G. (1985). Reef response to sea-level – Catch up, keep up, or give up. Proc. 5th Int. Coral Reef Congress, Morea, French Polynesia, 105–110.Google Scholar
Noe, S. U., et al. (2008). Varying growth rates in bamboo corals: Sclerochronology and radiocarbon dating of a mid-Holocene deep-water gorgonian skeleton (Keratoisis sp.: Octocorallia) from Chatham Rise (New Zealand). Facies, 54, 151–166.CrossRefGoogle Scholar
Novello, V. F., et al. (2018). Two millennia of South Atlantic Convergence Zone variability reconstructed from isotopic proxies. Geophys. Res. Lett., 45, 5045–5051. https://doi.org/10.1029/2017GL076838.CrossRefGoogle Scholar
Nunn, P. (2000). Significance of emerged Holocene corals around Ovalau and Moturiki islands, Fiji, southwest Pacific. Mar. Geol., 163, 345–351.CrossRefGoogle Scholar
Nurhati, I. S., et al. (2011). Correction to ‘Late 20th century warming and freshening in the central tropical Pacific’. Geophys. Res. Lett., 38(24), L24707.CrossRefGoogle Scholar
Nyberg, Y. (2002). Luminescence intensity in coral skeletons from Mona Island in the Caribbean Sea and its link to precipitation and wind speed. Phil. Trans. R. Soc. Lond. A, 360, 749–766, https://doi.org/10.1098/rsta.2001.0963.CrossRefGoogle ScholarPubMed
Obert, J. C., et al. (2016). 230Th/U dating of Last Interglacial brain corals from Bonaire (southern Caribbean) using bulk and theca wall material. Geochim. Cosmochim. Acta, 178, 20–40. https://doi.org/10.1016/j.gca.2016.01.011.CrossRefGoogle Scholar
Ortlieb, L. and Machare, J. (1993). Former El Niño events: Records from western South-America. Glob. Planet. Change, 7(1), 181–202.CrossRefGoogle Scholar
Ortlieb, L., et al., (1995). Beach ridges and major Late Holocene El Niño events in Northern Peru. J. Coast. Res., S.I. No. 17. Holocene cycles: Climate, sea levels, and sedimentation, 109–117.Google Scholar
PAGES2k Consortium. (2017). A global multiproxy database for temperature reconstructions of the Common Era. Sci. Data, 4(1), 170088. https://doi.org/10.1038/sdata.2017.88.Google Scholar
PAGES 2k Consortium. (2019). Consistent multidecadal variability in global temperature reconstructions and simulations over the Common Era. Nature Geosci., 12(8), 643–649. https://doi.org/10.1038/s41561-019-0400-0.Google Scholar
PAGES Hydro2k Consortium. (2017). Comparing proxy and model estimates of hydroclimate variability and change over the Common Era. Climate of the Past, 13, 1851–1900. https://doi.org/10.5194/cp-13-1851-2017.Google Scholar
Palmer, J. G. and Xiong, L. M. (2004). New Zealand climate over the last 500 years reconstructed from Libocedrus bidwillii Hook. f. tree-ring chronologies. Holocene, 14, 282–289.CrossRefGoogle Scholar
Pan, T. Y., et al. (2020). The last interglacial (MIS 5e) sea level highstand from a tectonically stable far-field setting, Yorke Peninsula, southern Australia. Mar. Geol., 398, 126–136. https://doi.org/10.1016/j.margeo.2018.01.012.Google Scholar
Partin, J. W., et al. (2013). Multidecadal rainfall variability in South Pacific Convergence Zone as revealed by stalagmite geochemistry. Geol., 41(11), 1143–1146.CrossRefGoogle Scholar
Paterne, M., et al. (2019). Variability of marine 14C reservoir ages in the Southern Ocean highlighting circulation changes between 1910 and 1950. Earth Planet. Sci. Lett., 511, 99–104.CrossRefGoogle Scholar
Pearce, A. F. and Feng, M. (2013). The rise and fall of the ‘marine heat wave’ off Western Australia during the summer of 2010/2011. J. Mar. Sys., 111–112, 139–156.Google Scholar
Peltier, W. R. (2001). Global isostatic adjustment and modern instrumental records of relative sea level history. In Douglas, B. C., et al. (eds.), Sea-level Rise History and Consequences. Academic Press, San Diego. pp. 65–95.Google Scholar
Peltier, W. R. (2004). Global glacial isostasy and the surface of the ice-age Earth: The ICE–5G (VM2) model and GRACE. Ann. Rev. Earth Sci., 32, 111–149.CrossRefGoogle Scholar
Peltier, W. R. and Fairbanks, R. G. (2006). Global glacial ice volume and Last Glacial Maximum duration from an extended Barbados sea level record. Quat. Sci. Rev., 25, 3322–3337.CrossRefGoogle Scholar
Pickett, J. W. (1981). A late Pleistocene coral fauna from Evans Head. N.S.W. Alcheringa, 5, 71–83.CrossRefGoogle Scholar
Pickett, J. W., et al. (1989). Review of age determinations on Pleistocene corals in eastern Australia. Quat. Res., 31, 392–395.CrossRefGoogle Scholar
Pirazzoli, P. A. (1991). World Atlas of Holocene Sea- Level Changes. Elsevier Oceanography Series 58, Elsevier, Amsterdam, 300pp.Google Scholar
Poitevin, P., et al. (2018). Ligament, hinge, and shell cross-sections of the Atlantic surfclam (Spisula solidissima): Promising marine environmental archives in NE North America. PLoS ONE, 13(6), e0199212. https://doi.org/10.1371/journal.pone.0199212.CrossRefGoogle ScholarPubMed
Poussart, P. F., et al. (2004). Resolving seasonality in tropical trees: Multi-decade, high-resolution oxygen and carbon isotope records from Indonesia and Thailand. Earth Planet. Sci. Lett., 218(3–4), 301–316.CrossRefGoogle Scholar
Poussart, P. M., et al. (2006). Tropical dendrochemistry: A novel approach to estimate age and growth from ringless trees. Geophys. Res. Lett., 33, L17711. https://doi.org/10.1029/2006GL026929.CrossRefGoogle Scholar
Pretet, C., et al. (2013). Constraining calcium isotope fractionation (δ44/40Ca) in modern and fossil scleractinian coral skeleton. Chem. Geol., 340, 49–58.CrossRefGoogle Scholar
Pucha-Cofrep, D., et al. (2015). Wet season precipitation during the past century reconstructed from tree-rings of a tropical dry forest in Southern Ecuador. Glob. Planet. Change, 133, 65–78.CrossRefGoogle Scholar
Rein, B., et al. (2005). El Niño variability off Peru during the last 20,000 years. Paleoceanogr., 20(4), PA4003. https://doi.org/10.1029/2004PA001099.CrossRefGoogle Scholar
Ren, L., et al. (2003). Deconvolving the δ18O seawater component from sub-seasonal coral δ18O and Sr/Ca at Rarotonga in the southwestern subtropical Pacific for the period 1726 to 1997. Geochim. Cosmochim. Acta, 67(9), 1609–1621.CrossRefGoogle Scholar
Rhoads, D. C. and Pannella, G. (1970). The use of molluscan shell growth patterns in ecology and paleoecology. Lethaia, 3, 143–161.CrossRefGoogle Scholar
Rhoads, D. C. and Lutz, R. A. (eds.). (1980) Skeletal Growth of Aquatic Organisms: Biological Records of Environmental Change. Plenum Press, New-York and London, 750pp.CrossRefGoogle Scholar
Richardson, J. B. (1983). The Chira beach ridges, sea level change, and the origins of maritime economies on the Peruvian coast. Annals Carnegie Museum, 52, 265–275.CrossRefGoogle Scholar
Richardson, C. A. (2001). Molluscs as archives of environmental changes. Oceanogr. Mar. Biol. Ann. Rev., 39, 103–164.Google Scholar
Roark, B., et al. (2005). Radiocarbon-based ages and growth rates of bamboo corals from the Gulf of Alaska, Geophys. Res. Lett., 32, L04606. https://doi.org/10.1029/2004GL021919.CrossRefGoogle Scholar
Rodriguez-Ramirez, A., et al. (2014). Coral luminescence identifies the Pacific Decadal Oscillation as a primary driver of river runoff variability impacting the southern Great Barrier Reef. PLoS One, 9(1), e84305. https://doi.org/10.1371/journal.pone.0084305.CrossRefGoogle ScholarPubMed
Rogers, S. S., et al. (2014). Coastal change and beach ridges along the northwest coast of Peru: Image and GIS Analysis of the Chira, Piura, and Colán Beach-Ridge Plains. J. Coast. Res., 20(4), 1102–1125.Google Scholar
Rohling, E. J., et al. (2008). High rates of sea-level rise during the last interglacial period. Nature Geosci., 1, 38–42.CrossRefGoogle Scholar
Rollins, H. B., et al. (1986). The birth of El Niño: Geoarchaeological evidence and implications. Geoarchaeology, 1, 1, 3–15.CrossRefGoogle Scholar
Rollins, H. B., et al. (1987). Growth increment and stable-isotope analysis of marine bivalves: Implications for the geoarchaeological record of El Niño. Geoarchaeology, 2, 181–187.CrossRefGoogle Scholar
Roman-Gonzalez, A., et al. (2017). A sclerochronological archive for Antarctic coastal waters based on the marine bivalve Yoldia eightsi (Jay, 1839) from the South Orkney Islands. Holocene, 27(2), 271–281.CrossRefGoogle Scholar
Roman, M., et al. (2021). A multi-decadal geochemical record from Rano Aroi (Easter Island/ Rapa Nui): Implications for the environment, climate and humans during the last two millennia. Quat. Sci. Rev., 268(2021), 107115.CrossRefGoogle Scholar
Russon, T., et al. (2013). Inter-annual tropical Pacific climate variability in an isotope-enabled CGCM: Implications for interpreting coral stable oxygen isotope records of ENSO. Climate of the Past, 9(4), 1543–1557.CrossRefGoogle Scholar
Sachs, J. P., et al. (2009). Southward movement of the Pacific intertropical convergence zone AD 1400–1850. Nat. Geosci., 2, 519–525.CrossRefGoogle Scholar
Sachs, J. P. and Mhyrvold, C. L. (2011). A shifting band of rain. Scientific American, 304(3), 60–65.CrossRefGoogle ScholarPubMed
Sachs, J. P., et al. (2021). Last millennium hydroclimate in the central equatorial North Pacific (5°N, 160°W). Quat. Sci. Rev., 259, 106906. https://doi.org/10.1016/j.quascirev.2021.106906.CrossRefGoogle Scholar
Sachse, D., et al. (2004). Hydrogen isotope ratios of recent lacustrine sedimentary n-alkanes record modern climate variability, Geochim. Cosmochim. Acta, 68, 4877–4889.CrossRefGoogle Scholar
Sachse, D., et al. (2012). Molecular paleohydrology: Interpreting the hydrogen-isotopic composition of lipid biomarkers from photosynthesizing organisms. Annu. Rev. Earth Planet Sci., 40, 221–249.CrossRefGoogle Scholar
Sadler, J., et al. (2012). Reconstructing past upwelling intensity and the seasonal dynamics of primary productivity along the Peruvian coastline from mollusk shell stable isotope. Geochem. Geophys. Geosyst., 13, Q01015. https://doi.org/10.1029/2011GC003595.CrossRefGoogle Scholar
Sadler, J., et al. (2014). Geochemistry-based coral palaeoclimate studies and the potential of ‘non-traditional’ (non-massive Porites) corals: Recent developments and future progression. Earth-Sci. Rev., 139, 291–316.CrossRefGoogle Scholar
Sadler, J., et al. (2016). Acropora interbranch skeleton Sr/Ca ratios: Evaluation of a potential new high-resolution paleothermometer. Paleoceanogr., 31, 505–517.CrossRefGoogle Scholar
Saenger, C., et al. (2012). Carbonate clumped isotope variability in shallow water corals: Temperature dependence and growth- related vital effects. Geochim. Cosmochim. Acta, 99, 224–242.CrossRefGoogle Scholar
Saenger, C. and Wang, Z. (2014). Magnesium isotope fractionation in biogenic and abiogenic carbonates: Implications for paleoenvironmental proxies. Quat. Sci. Rev., 90, 1–21.CrossRefGoogle Scholar
Saha, N., et al. (2016). Coral skeletal geochemistry as a monitor of inshore water quality. Sci. Total Environ., 566–567, 652–684. https://doi.org/10.1016/j.scitotenv.2016.05.066.Google ScholarPubMed
Saha, N., et al. (2018). Seasonal to decadal scale influence of environmental drivers on Ba/Ca and Y/Ca in coral aragonite from the southern Great Barrier Reef. Sci. Total Environ., 639, 1099–1109. https://doi.org/10.1016/j.scitotenv.2018.05.156.CrossRefGoogle Scholar
Sandweiss, D. H. (1986) The beach ridges at Santa, Peru: El Niño, uplift, and prehistory. Geoarcheology, 1, 17–28.CrossRefGoogle Scholar
Sandweiss, D. H. (2003). Terminal Pleistocene through Mid-Holocene archaeological sites as paleoclimatic archives for the Peruvian coast. Palaeogeogr. Palaeoclimatol. Palaeoecol., 194(1–3), 23–40.CrossRefGoogle Scholar
Sandweiss, D. H., et al. (1983). Landscape alteration and prehistoric human occupation on the North Coast of Peru. Ann. Carnegie Mus., 52, 277–298.CrossRefGoogle Scholar
Sandweiss, D. H., and Richardson, J. B. (1996). Geoarcheological evidence from Peru for a 5000 years B.P. onset of El Niño. Science, 273, 1531–1533.CrossRefGoogle Scholar
Sandweiss, D. H., et al. (1996). Geoarcheological evidence from Peru for a 5000 years B.P. onset of El Niño. Science, 273, 1531–1533.CrossRefGoogle Scholar
Sandweiss, D. H., et al. (2001). Variation in Holocene El Niño frequencies: Climate records and cultural consequences in ancient Peru. Geol., 29(7), 603–606.2.0.CO;2>CrossRefGoogle Scholar
Sarnthein, M., et al. (1981). Glacial and interglacial wind regimes over the eastern subtropical Atlantic and northwest Africa. Nature, 293, 193–196.CrossRefGoogle Scholar
Sayani, H. R., et al. (2011). Effects of diagenesis on paleoclimate reconstructions from modern and young fossil corals. Geochim. Cosmochim. Acta, 75(21), 6361–6373.CrossRefGoogle Scholar
Sayani, H. R., et al. (2019). Intercolony δ18O and Sr/Ca variability among Porites spp. corals at Palmyra Atoll: Toward more robust coral-based estimates of climate. Geochem. Geophys. Geosyst., 20(11), 5270–5284. https://doi.org/10.1029/2019GC008420.CrossRefGoogle Scholar
Sayani, H. R., et al. (2021). Reproducibility of coral Mn/Ca-based wind reconstructions at Kiritimati Island and Butaritari Atoll. Geochem. Geophys. Geosyst., 22, e2020GC009398. https://doi.org/10.1029/2020GC009398.CrossRefGoogle Scholar
Schefuß, E., et al. (2005). Climatic controls on central African hydrology during the past 20,000 years. Nature, 437, 1003–1006. https://doi.org/10.1038/nature03945.CrossRefGoogle ScholarPubMed
Schollaen, K., et al. (2013). Multiple tree-ring chronologies (ring width, δ13C and δ18O) reveal dry and rainy season signals of rainfall in Indonesia. Quat. Sci. Rev., 73, 170–181.CrossRefGoogle Scholar
Schone, B. R. (2013). Arctica islandica (Bivalvia): A unique paleoenvironmental archive of the northern North Atlantic Ocean. Glob. Planet. Change, 111, 199–225. https://doi.org/10.1016/j.gloplacha.2013.09.013.CrossRefGoogle Scholar
Schone, B. R. and Gillikin, D. P. (2013). Unraveling environmental histories from skeletal diaries – Advances in sclerochronology. Palaeogeogr. Palaeoclimatol. Palaeoecol., 373, 1–5.CrossRefGoogle Scholar
Schöne, B. R., et al. (2013). Crystal fabrics and element impurities (Sr/Ca, Mg/Ca, and Ba/Ca) in shells of Arctica islandica – Implications for paleoclimate reconstructions. Palaeogeogr. Palaeoclimatol. Palaeoecol., 373, 50–59.CrossRefGoogle Scholar
Schrag, D. P. (1999). Rapid analysis of high-precision Sr/Ca ratios in corals and other marine carbonates. Paleoceanogr., 14, 97–102.CrossRefGoogle Scholar
Scoffin, T. P. (1993). The geological effects of hurricanes on coral reefs and the interpretation of storm deposits. Coral Reefs, 12, 203–221.CrossRefGoogle Scholar
Scoffin, T. P., et al. (1978). The nature and significance of microatolls. Philos. Trans. R. Soc. B., Biological Sciences, 284(999), 99–122.Google Scholar
Sear, D. A., et al. (2020). Human settlement of East Polynesia earlier, incremental, and coincident with prolonged South Pacific drought. Proc. Nat. Acad. Sci. USA, 117(16), 8813–8819. www.pnas.org/cgi/doi/10.1073/pnas.1920975117.CrossRefGoogle Scholar
Shen, G. T., et al. (1987) Cadmium in corals as a tracer of historical upwelling and industrial fallout. Nature, 328, 794–796.CrossRefGoogle Scholar
Shen, G. T., et al. (1991). Paleochemistry of manganese in corals from the Galapagos Islands. Coral Reefs, 10(2), 91–100.CrossRefGoogle Scholar
Shen, G. T., et al. (1992). A chemical indicator of trade wind reversal in corals from the western tropical Pacific. J. Geophys. Res., 97, 12698–12697.Google Scholar
Shen, C.-C., et al. (1996). The calibration of D[Sr/Ca] versus sea surface temperature relationship for Porites corals. Geochim. Cosmochim. Acta, 60, 3849–3858.CrossRefGoogle Scholar
Sherwood, O. A., et al. (2009). Multi-century time-series of 15N and 14C in bamboo corals from deep Tasmanian seamounts: Evidence for stable oceanographic conditions. Mar. Ecol. Prog. Ser., 397, 209–218, https://doi.org/10.3354/meps08166.CrossRefGoogle Scholar
Shinn, E. A., et al. (2000). African dust and the demise of Caribbean coral reefs. Geophys. Res. Lett., 27, 3029–3032.CrossRefGoogle Scholar
Smith, S. V. and Buddemeier, R. W. (1992). Global change and coral reef ecosystems. Ann. Rev. Ecol. Syst., 23(1), 89–118.CrossRefGoogle Scholar
Smithers, S. (2011). Microatoll. In Hopley, D. (eds.), Encyclopedia of Modern Coral Reefs. Encyclopedia of Earth Sciences Series. Springer, Dordrecht. https://doi-org.simsrad.net.ocs.mq.edu.au/10.1007/978-90-481-2639-2_111.Google Scholar
Smithers, S. G. and Woodroffe, C. D. (2000). Microatolls as sea- level indicators on a mid-ocean atoll. Mar. Geol., 168, 61–78.CrossRefGoogle Scholar
Smithers, S. G. and Woodroffe, C. D. (2001). Coral microatolls and 20th century sea level in the eastern Indian Ocean. Earth Planet. Sci. Lett., 191, 173–184.CrossRefGoogle Scholar
Smittenberg, R. H., et al. (2011). Compound-specific D/ H ratios of the marine lakes of Palau as proxies for West Pacific Warm Pool hydrologic variability. Quat. Sci. Rev., 30, 921–933.CrossRefGoogle Scholar
Spencer, T., et al. (1997). Reconstructing sea level change from coral microatolls, Tongareva (Penrhyn) Atoll, northern Cook Islands. In Lessios, H. A. and Macintyre, I. G. (eds.), Proc. 8th International Coral Reef Symposium. Smithsonian Tropical Research Institute, Panama, pp. 489–494.Google Scholar
Stahle, D. W., et al. (1999). Management implications of annual growth rings in Pterocarpus angolensis from Zimbabwe. For. Ecol. Manag., 124, 217–229.CrossRefGoogle Scholar
Stevenson, C. M., et al. (2015). Variation in Rapa Nui (Easter Island) land use indicates production and population peaks prior to European contact. Proc. Nat. Acad. Sci. USA, 112(4), 1025–1030. www.pnas.org/cgi/doi/10.1073/pnas.1420712112.CrossRefGoogle ScholarPubMed
Stevenson, S., et al. (2018). Twentieth century seawater δ18O dynamics and implications for coral-based climate reconstruction. Paleoceanogr. Paleoclimatol., 33, 606–625, https://doi.org/10.1029/2017PA003304.CrossRefGoogle Scholar
Stoddart, D. R. and Scoffin, T. P. (1979). Microatolls: Review of form, origin and terminology. Atoll Res. Bull., 224, 1–17.CrossRefGoogle Scholar
Stuiver, M., et al. (1981). Isotopic indicators of age/growth in tropical trees. In Bohrmann, F. H. and Berlyn, G. (eds.), Age and Growth Rate of Tropical Trees: New Directions for Research. Yale University Press, New Haven, pp. 75–82.Google Scholar
Su, L., et al. (2019). Ningaloo Niño/Niña and their regional climate impacts as recorded by corals along the coast of Western Australia. Palaeogeogr. Palaeoclimatol. Palaeoecol., 535, 109368.CrossRefGoogle Scholar
Takahashi, K. and Martinez, A. G. (2019). The very strong coastal El Niño in 1925 in the far-eastern Pacific. Clim. Dyn., 52, 7389–7415. https://doi.org/10.1007/s00382-017-3702-1.CrossRefGoogle Scholar
Thiede, J. (1979). Wind regimes over the late Quaternary southwest Pacific Ocean. Geology, 7, 259–262.2.0.CO;2>CrossRefGoogle Scholar
Thompson, D. M. (2022). Environmental records from coral skeletons: A decade of novel insights and innovation. WIRES Clim. Change, 13(1), e745. https://doi.org/10.1002/wcc.745.CrossRefGoogle Scholar
Thompson, D. M., et al. (2011). Comparison of observed and simulated tropical climate trends using a forward model of coral δ18O. Geophys. Res. Lett., 38(14), L14706. https://doi.org/10.1029/2011GL048224.Google Scholar
Thompson, D. M., et al. (2013). Coral-model comparison highlighting the role of salinity in long-term trends. PAGES News, 21(2), 60–61.CrossRefGoogle Scholar
Thompson, D. M., et al. (2015). Early twentieth-century warming linked to tropical Pacific wind strength. Nat. Geosci., 8, 117–120. https://doi.org/10.1038/NGEO2321.CrossRefGoogle Scholar
Thompson, W. G., et al. (2003). An open-system model for U-series age determinations of fossil corals. Earth Planet. Sci. Lett., 210, 1–2, 365–381. https://doi.org/10.1016/S0012-821X(03)00121-3.CrossRefGoogle Scholar
Thresher, R., et al. (2004). Oceanic evidence of climate change in southern Australia over the last three centuries. Geophys. Res. Lett., 31, L07212. https://doi.org/10.1029/2003GL018869, 2004.CrossRefGoogle Scholar
Tierney, J. E., et al. (2010). Coordinated hydrological regimes in the Indo-Pacific region during the past two millennia. Paleoceanogr., 25, PA1102.CrossRefGoogle Scholar
Tudhope, A. W., et al. (2001). Variability in El Niño-Southern Oscillation through a glacial-interglacial cycle. Science, 291, 1511–1517.CrossRefGoogle ScholarPubMed
Urey, H. C. (1947). The thermodynamic properties of isotopic substances. J. Chem. Soc., 1, 562–581.Google Scholar
Urey, H. C. (1948). Oxygen isotopes in nature and in the laboratory. Science, 108, 489–496.CrossRefGoogle ScholarPubMed
Vecchi, G. A. and Harrison, D. E. (2000). Tropical Pacific sea surface temperature anomalies, El Niño, and equatorial westerly wind events. J. Clim., 13, 1814–1830.2.0.CO;2>CrossRefGoogle Scholar
Veron, J. E. N. (1995). Corals in Space and Time: The Biogeography and Evolution of the Scleractinia. Cornell University Press, New York, 321pp.Google Scholar
Veron, J. E. N., et al. (2016). Corals of the World. www.coralsoftheworld.org.Google Scholar
Villalba, R. (1990). Climatic fluctuations in Northern Patagonia during the last 1000 years as inferred from tree-ring records. Quat. Res., 34, 346–360.CrossRefGoogle Scholar
Villalba, R., et al. (1997). Sea-level pressure variability around Antarctica since A.D. 1750 inferred from subantarctic tree-ring records. Clim. Dyn., 13(6), 375–390.CrossRefGoogle Scholar
Volland, F., et al. (2016). Hydro-climatic variability in southern Ecuador reflected by tree-ring oxygen isotopes. Erdkunde, 70, 69–82. https://doi.org/10.3112/erdkunde.2016.01.05.Google Scholar
von Reumont, J., et al. (2018). Tracking interannual- to multi-decadal-scale climate variability in the Atlantic Warm Pool using Central Caribbean coral data. Paleoceanogr. Paleoclimatol., 33, 395–411. https://doi.org/10.1002/2018PA003321.CrossRefGoogle Scholar
Wahl, E. and Frank, D. (2012). Evidence of environmental change from annually-resolved proxies with particular reference to dendroclimatology and the last millennium. In Matthews, J. A. (ed), The SAGE Handbook of Environmental Change, 1, SAGE Publications, London, pp. 320–344.Google Scholar
Warner, J. P., et al. (2022). Investigating the influence of temperature and seawater δ18O on Donax obesulus (Reeve, 1854) shell δ18O. Chem. Geol., 588, 120638.CrossRefGoogle Scholar
Watanabe, T., et al. (2004). A 60- year isotopic record from a mid-Holocene fossil giant clam (Tridacna gigas) in the Ryukyu Islands: Physiological and paleoclimatic implications. Palaeogeogr. Palaeoclimatol. Palaeoecol., 212, 343–354.CrossRefGoogle Scholar
Watanabe, T. K., et al. (2017). Past summer upwelling events in the Gulf of Oman derived from a coral geochemical record. Sci. Rep., 7(1), 4568. https://doi.org/10.1038/s41598-017-04865-5.CrossRefGoogle ScholarPubMed
Webster, J. M., et al. (2004). Drowning of the –150 m reef off Hawaii: A casualty of global meltwater pulse 1A? Geol., 32, 249–252.CrossRefGoogle Scholar
Webster, J. M., et al. (2007). Numerical modeling of the growth and drowning of Hawaiian coral reefs during the last two glacial cycles (0–250 kyr). Geochem. Geophys. Geosyst., 8, Q03011.CrossRefGoogle Scholar
Webster, J. M., et al. (2009). Coral reef evolution on rapidly subsiding margins. Glob. Planet. Change, 66, 129–148.CrossRefGoogle Scholar
Weil-Accardo, J., et al. (2016). Relative sea-level changes during the last century recorded by coral microatolls in Belloc, Haiti. Glob. Planet. Change, 139, 1–14.CrossRefGoogle Scholar
Weil-Accardo, J., et al. (2020). Relative sea-level changes over the past centuries in the central Ryukyu Arc inferred from coral microatolls. J. Geophys. Res: Solid Earth, 125, e2019JB018466. https://doi.org/10.1029/2019JB018466.CrossRefGoogle Scholar
Woodroffe, C. D. and McLean, R. F. (1990). Microatolls and recent sea level change on coral atolls. Nature, 344, 531–534.CrossRefGoogle Scholar
Woodroffe, C. D. and Gagan, M. K. (2000). Coral microatolls from the central Pacific record late Holocene El Nino. Geophys. Res. Lett., 27, 1511–1514.CrossRefGoogle Scholar
Woodroffe, C. D., et al. (2000). Holocene reef growth in Torres Strait. Mar. Geol., 170, 331–346.CrossRefGoogle Scholar
Woodroffe, C. D., et al. (2003). Mid-late Holocene El Niño variability in the equatorial Pacific from coral microatolls. Geophys. Res. Lett., 30, 7, 1358. https://doi.org/10.1029/2002GL015868.CrossRefGoogle Scholar
Woodroffe, C. D., et al. (2012). Mid-Pacific microatolls record sea-level stability over the past 5000 yr. Geology, 40, 951–954.CrossRefGoogle Scholar
Worbes, M. (2002) One hundred years of tree-ring research in the tropics – A brief history and an outlook to future challenges. Dendrchronol., 20, 1–2, 217–231.Google Scholar
Wu, H. C., et al. (2013). Oceanographic variability in the South Pacific Convergence Zone region over the last 210 years from multi-site coral Sr/Ca records. Geochem. Geophys. Geosyst., 14(5), 1435–1453. https://doi.org/10.1029/2012GC004293.CrossRefGoogle Scholar
Yan, H., et al. (2015). Dynamics of the intertropical convergence zone over the western Pacific during the Little Ice Age. Nat. Geosci., 8, 315–320.CrossRefGoogle Scholar
Zaw, Z., et al. (2020). Drought reconstruction over the past two centuries in southern Myanmar using teak tree-rings: Linkages to the Pacific and Indian Oceans. Geophys. Res. Lett., 47, e2020GL087627. https://doi.org/10.1029/2020GL087627.CrossRefGoogle Scholar
Zinke, J., et al. (2009). Western Indian Ocean marine and terrestrial records of climate variability: A review and new concepts on land–ocean interactions since AD 1660. Int. J. Earth Sci. (Geol Rundsch), 98, 115–133. https://doi.org/10.1007/s00531-008-0365-5.CrossRefGoogle Scholar
Zinke, J., et al. (2014). Corals record long-term Leeuwin Current variability during Ningaloo Niño/Niña since 1795. Nat. Commun., 5, 1, 3607. https://doi.org/10.1038/ncomms4607.CrossRefGoogle ScholarPubMed
Zinke, J., et al. (2015). Madagascar corals track sea surface temperature variability in the Agulhas Current core region over the past 334 years. Sci. Rep., 4(1), 4393. https://doi.org/10.1038/srep04393.Google Scholar
Ahlmann, H. W. (1948). Glaciological Research on the North Atlantic Coasts. London: Royal Geographical Society.Google Scholar
Anderson, B. M. and Mackintosh, A. N. (2006). Temperature change is the major driver of late-glacial and Holocene glacier fluctuations in New Zealand. Geol., 34(2), 121–134. https://doi.org/10.1130/G22151.1.CrossRefGoogle Scholar
Anderson, B. M., et al. (2008). Response of Franz Josef Glacier Ka Roimata o Hine Hukatere to climate change. Glob. Planet. Change, 63, 23–30. https://doi.org/10.1016/j.gloplacha.2008.04.003.CrossRefGoogle Scholar
Anderson, B. and Mackintosh, A. (2012). Controls on mass balance sensitivity of maritime glaciers in the Southern Alps, New Zealand: The role of debris cover. J. Geophys. Res., 117, F01003. https://doi.org/10.1029/2011JF002064.Google Scholar
Aristarain, A. J. and Delmas, R. J. (1993). Firn-core study from the southern Patagonia ice cap, South America. J. Glaciol., 39(132), 249–254.CrossRefGoogle Scholar
Bahr, D. B., et al. (1998). Response time of glaciers as a function of size and mass balance: 1. Theory. J. Geophys. Res., 103, 9777–9782.Google Scholar
Balco, G. (2011). Contributions and unrealized potential contributions of cosmogenic-nuclide exposure dating to glacier chronology, 1990–2010. Quat. Sci. Rev., 30, 3–27.CrossRefGoogle Scholar
Banerjee, A. and Shankar, R. (2013). On the response of Himalayan glaciers to climate change. J. Glaciol., 59, 480–490. https://doi.org/10.3189/2013JoG12J130.CrossRefGoogle Scholar
Benn, D. I. and Ballantyne, C. K. (1994). Reconstructing the transport history of glacigenic sediments: A new approach based on the co-variance of clast form indices. Sed. Geol., 91, 215–227.CrossRefGoogle Scholar
Benn, D. I., et al. (2005). Reconstruction of equilibrium-line altitudes for tropical and subtropical glaciers. Quat. Int., 138–139, 8–21.Google Scholar
Benn, D. I. and Evans, D. J. A. (2010). Glaciers and Glaciation, 2nd ed. Routledge, New York.Google Scholar
Bolano-Ortiz, T. T., et al. (2019). Assessment of absorbing aerosols on austral spring snow albedo reduction by several basins in the Central Andes of Chile from daily satellite observations (2000–2016) and a case study with the WRF-Chem model. SN Applied Sciences, 1, 1352. https://doi.org/10.1007/s42452-019-1256-z.CrossRefGoogle Scholar
Bolius, D., et al. (2006). A first shallow firn-core record from Glaciar La Ollada, Cerro Mercedario, central Argentine Andes. Ann. Glaciol., 43, 14–22. https://doi.org/10.3189/172756406781812474.CrossRefGoogle Scholar
Boyle, E. A. (1997). Cool tropical temperatures shift the global δ18O–T relationship: An explanation for the ice core δ18O – borehole thermometry conflict? Geophys. Res. Lett., 24, 273–276.CrossRefGoogle Scholar
Bozkurt, D., et al. (2018). Foehn event triggered by an atmospheric river underlies record-setting temperature along continental Antarctica. J. Geophys. Res., Atmos., 123, 3871–3892. https://doi.org/10.1002/2017JD027796.CrossRefGoogle Scholar
Bradley, R. S. (1999). Paleoclimatology: Reconstructing Climates of the Quaternary, 2nd ed. International Geophysics Press, Volume 64, 610pp, Academic Press, San Diego.Google Scholar
Bradley, R. S., et al. (2003). Low latitude ice cores record Pacific sea surface temperatures. Geophys. Res. Lett., 30(4), 1174. https://doi.org/10.1029/2002GL016546.CrossRefGoogle Scholar
Braithwaite, R. J. (1981). On glacier energy balance, ablation and air temperature. J. Glaciol., 27(97), 381–391.CrossRefGoogle Scholar
Braithwaite, R. J. (1985). Calculation of degree days for glacier climate research. Z. Glelscherkd. Glazialgeol., 20, 1–8.Google Scholar
Braithwaite, R. J. and Olesen, O. B. (1985). Ice ablation in West Greenland in relation to air temperature and global radiation. Z. Glelscherkd. Glazialgeol., 20(1984), 155–168.Google Scholar
Braithwaite, R. J. (1995). Positive degree-day factors for ablation on the Greenland ice sheet studied by energy-balance modeling. J. Glaciol., 41(137), 153–160.CrossRefGoogle Scholar
Braithwaite, R. J. (2008). Temperature and precipitation climate at the equilibrium-line altitude of glaciers expressed by the degree-day factor for melting snow. J. Glaciol., 54, 437–444. https://doi.org/10.3189/002214308785836968.CrossRefGoogle Scholar
Braithwaite, R. J. (2011). Degree days. In Singh, V. P., et al. (eds.), Encyclopedia of Snow, Ice and Glaciers, pp. 196–199, Springer, New York. https://doi.org/10.1007/978-90-481-2642-2.Google Scholar
Braithwaite, R. J., et al. (2003). Temperature sensitivity of the mass balance of mountain glaciers and ice caps as a climatological characteristic. Z. Glelscherkd. Glazialgeol., 38(1), 35–61.Google Scholar
Braithwaite, R. J. and Hughes, P. D. (2020). Regional geography of glacier mass balance variability over seven decades 1946–2015. Front. Earth Sci., 8, 302. https://doi.org/10.3389/feart.2020.00302.CrossRefGoogle Scholar
Bravo, C., et al. (2019). Air temperature characteristics, distribution, and impact on modeled ablation for the South Patagonia Icefield. J. Geophys. Res., Atmos., 124, 907–925. https://doi.org/10.1029/2018JD028857.CrossRefGoogle Scholar
Bravo, C., et al. (2019). Assessing snow accumulation patterns and changes on the Patagonian Icefields. Geophys. Res. Lett., 7, 30. https://doi.org/10.3389/fenvs.2019.00030.Google Scholar
Briner, J. P. (2011). Dating glacial landforms. In Singh, V. P., et al. (eds.), Encylcopedia of Snow, Ice and Glaciers. Springer, Dordrecht, 175–186.Google Scholar
Budd, W. F. and Jenssen, D. (1975). Numerical modelling of glacier systems. In Snow and Ice Symposium – Neiges et Glaces. IAHS, Wallingford, pp. 257–291.Google Scholar
Cape, M. R., et al. (2015). Foehn winds link climate-driven warming to ice shelf evolution in Antarctica, J. Geophys. Res. Atmos., 120, 11037–11057. https://doi.org/10.1002/2015JD023465.CrossRefGoogle Scholar
Carrivick, J. L. and Chase, S. E. (2011). Spatial and temporal variability of annual glacier equilibrium line altitudes in the Southern Alps, New Zealand. N. Z. J. Geol. Geophys., 54(4), 415–429. https://doi.org/10.1080/00288306.2011.607463.Google Scholar
Carrivick, J. L. and Tweed, F. S. (2013). Proglacial lakes: Character, behaviour and geological importance. Quat. Sci. Rev., 78, 34–52.CrossRefGoogle Scholar
Casassa, G., et al. (1998). Glaciers in South America. In Haeberli, W., et al. (eds.), Into the Second Century of World-Wide Glacier Monitoring: Prospects and Strategies. A contribution to the International Hydrological Programme (IHP) and the Global Environment Monitoring System (GEMS). UNESCO, Paris, 125–146.Google Scholar
Casassa, G., et al. (2002). The Patagonian Icefields: A Unique Natural Laboratory for Environmental and Climate Change Studies. Springer, New York, 194pp.CrossRefGoogle Scholar
Charlesworth, J. K. (1957). The Quaternary Era. 2 Volumes. Edward Arnold, London.Google Scholar
Chinn, T. J. (1996). New Zealand glacier responses to climate change of the past century, N. Z. J. Geol. Geophys., 39(3), 415–428. https://doi.org/10.1080/00288306.1996.9514723.CrossRefGoogle Scholar
Chinn, T. J., et al. (2005). Use of ELA as a practical method of monitoring glacier response to climate in New Zealand’s Southern Alps. J. Glaciol., 51(172), 85–95. http://dx.doi.org/10.3189/172756505781829593.CrossRefGoogle Scholar
Chinn, T., et al. (2012). Annual ice volume changes 1976–2008 for the New Zealand Southern Alps. Glob. Planet. Change, 92–93, 105–118.Google Scholar
Cogley, J. G., et al. (2014). Remote sensing of glaciers of the Subantarctic islands. Chapter 32. In Kargel, J. S., et al. (eds.), Global Land Ice Measurements from Space. Springer-Verlag, Berlin Heidelberg, pp. 759–780.Google Scholar
Collier, E., et al. (2019). The influence of tropical cyclones on circulation, moisture transport, and snow accumulation at Kilimanjaro during the 2006–2007 season. J. Geophys. Res., Atmos., 124, 6919–6928. https://doi.org/10.1029/2019JD030682.CrossRefGoogle Scholar
Craig, H. (1961). Isotopic variations in meteoric waters. Science, 133, 1702–1703.CrossRefGoogle ScholarPubMed
Cuffey, K. M. and Patterson, W. S. B. (2010). The Physics of Glaciers. Butterworth-Heinemann, Oxford.Google Scholar
Cullen, N. J., et al. (2007). Energy-balance model validation on the top of Kilimanjaro, Tanzania, using eddy covariance data. Ann. Glaciol., 46, 227–233. https://doi.org/10.3189/172756407782871224.CrossRefGoogle Scholar
Cullen, N. J. and Conway, J. P. (2015). A 22 month record of surface meteorology and energy balance from the ablation zone of Brewster Glacier, New Zealand. J. Glaciol., 61(229), 931–946.CrossRefGoogle Scholar
Dansgaard, W. (1964). Stables isotopes in precipitation. Tellus, 16(4), 436–447.Google Scholar
De Angelis, H. (2014). Hypsometry and sensitivity of the mass balance to changes in equilibrium-line altitude: The case of the Southern Patagonia Icefield. J. Glaciol., 60, 14–28.CrossRefGoogle Scholar
De Angelis, M., et al. (2003). Volcanic eruptions recorded in the Illimani ice core (Bolivia): 1918–1998 and Tambora periods. Atmos. Chem. Phys., 3, 1725–1741.CrossRefGoogle Scholar
Donaghue, S. L. (2009). Changes in the morphology, mass balance, and dynamics of Brown Glacier, Heard Island, with comparison to the surrounding sub-Antarctic islands. Unpubl. PhD thesis, University of Tasmania, 233pp.Google Scholar
Drewry, D., 1986. Glacial Geologic Processes. Edward Arnold, London, 276pp.Google Scholar
Dussaillant, I., et al. (2019). Two decades of glacier mass loss along the Andes. Nat. Geosci., 12, 802–808. https://doi.org/10.1038/s41561-019-0432-5.CrossRefGoogle Scholar
Eaves, S. R., et al. (2019). Climate amelioration during the Last Glacial Maximum indicated by a sensitive mountain glacier in New Zealand. Geol., 47, 299–302.CrossRefGoogle Scholar
Egholm, D. L., et al. (2011). Modeling the flow of glaciers in steep terrains: The integrated second-order shallow ice approximation (iSOSIA). J. Geophys. Res., 116, F02012.Google Scholar
Falaschi, D., et al. (2019). Six Decades (1958–2018) of Geodetic Glacier Mass Balance in Monte San Lorenzo, Patagonian Andes. Front. Earth Sci., 7, 326. https://doi.org/10.3389/feart.2019.00326.CrossRefGoogle Scholar
Flint, R. F. (1947). Glacial Geology and the Pleistocene Epoch. John Wiley and Sons, New York, 589pp.Google Scholar
Flores-Rojas, J. L., et al. (2021). On the dynamic mechanisms of intense rainfall events in the central Andes of Peru, Mantaro valley. Atmos. Res., 248, 105188.CrossRefGoogle Scholar
Fonseca, R. and Martin-Torres, J. (2019). High-resolution dynamical downscaling of re-analysis data over the Kerguelen Islands using the WRF model. Theor. Appl. Climatol., 135, 1259–1277. https://doi.org/10.1007/s00704-018-2438-0.CrossRefGoogle Scholar
Furbish, D. J. and Andrews, J. T. (1984). The use of hypsometry to indicate long term stability and response of valley glaciers to changes in mass transfer. J. Glaciol., 30, 199–211.CrossRefGoogle Scholar
Ginot, P., et al. (2006). Glacier mass balance reconstruction by sublimation induced enrichment of chemical species on Cerro Tapado (Chilean Andes). Clim. Past, 2, 21–30.CrossRefGoogle Scholar
Ginot, P., et al. (2010). Influence of the Tungurahua eruption on the ice core records of Chimborazo, Ecuador. Cryosphere, 4(4), 561–568.Google Scholar
Glasser, N. F., et al. (2009). Topographic controls on glacier sediment–landform associations around the temperate North Patagonian Icefield. Quat. Sci. Rev., 28, 2817–2832.CrossRefGoogle Scholar
Goehring, B. M., et al. (2011). The Rhone Glacier was smaller than today for most of the Holocene. Geol., 39, 679–682.CrossRefGoogle Scholar
Golledge, N. R., et al. (2012). Last Glacial Maximum climate in New Zealand inferred from a modelled Southern Alps icefield. Quat. Sci. Rev., 46, 30–45.CrossRefGoogle Scholar
Gundestrup. (2002). Locating a drill site on the Patagonian icefields. In Casassa, G., et al. (eds.), The Patagonian Icefields: A Unique Natural Laboratory for Environmental and Climate Change Studies. Springer, New York, pp. 117–124.CrossRefGoogle Scholar
Hambrey, M. J. (1994). Glacial Environments. UCL Press, London, 296pp.Google Scholar
Hobbs, W. H. (1922). Characteristics of Existing Glaciers. Macmillan, New York, 301pp.Google Scholar
Hock, R. (2003). Temperature index melt modelling in mountain areas. J. Hydrol., 282(1–4), 104–115.CrossRefGoogle Scholar
Hock, R. (2005). Glacier melt: A review of processes and their modelling. Prog. Phys. Geogr., 29(3), 362–391. https://doi.org/10.1191/0309133305pp453ra.CrossRefGoogle Scholar
Hodgson, D. A., et al. (2014). Terrestrial and marine evidence for the extent and timing of glaciation on the sub-Antarctic islands. Quat. Sci. Rev., 100, 137–158. http://dx.doi.org/10.1016/j.quascirev.2013.12.001.CrossRefGoogle Scholar
Hoffmann, G., et al. (2003). Coherent isotope history of Andean ice cores over the last century. Geophys. Res. Lett., 30(4), 1179. https://doi.org/10.1029/2002GL014870.CrossRefGoogle Scholar
Izagirre, E., et al. (2018). Glacial geomorphology of the Marinelli and Pigafetta glaciers, Cordillera Darwin Icefield, southernmost Chile. J. Maps, 14(2), 269–281. https://doi.org/10.1080/17445647.2018.1462264.CrossRefGoogle Scholar
Jara, I. A., et al. (2019). Centennial-scale precipitation anomalies in the southern Altiplano (18°S) suggest an extratropical driver for the South American summer monsoon during the late Holocene. Clim. Past, 15, 1845–1859.CrossRefGoogle Scholar
Johannesson, T., et al. (1989). Timescale for adjustments of glaciers to changes in mass balance. J. Glaciol., 35(121), 355–369.CrossRefGoogle Scholar
Junquas, C., et al. (2018). Understanding the influence of orography on the precipitation diurnal cycle and the associated atmospheric processes in the central Andes. Clim Dyn., 50, 3995–4017. https://doi.org/10.1007/s00382-017-3858-8.CrossRefGoogle Scholar
Kargel, J. S., et al. (eds.). (2014). Global Land Ice Measurements from Space. Springer-Verlag, Berlin Heidelberg, 876pp.CrossRefGoogle Scholar
Kaser, G. (2001). Glacier-climate interaction at low latitudes. J. Glaciol., 47, 195–204.CrossRefGoogle Scholar
Kaser, G. and Osmaston, H. (2002). Tropical Glaciers. International Hydrology Series. Cambridge University Press, Cambridge, 207pp.Google Scholar
Kaser, G., et al. (2010). Is the decline of ice on Kilimanjaro unprecedented in the Holocene? Holocene, 20(7), 1079–1091. https://doi.org/10.1177/0959683610369498.CrossRefGoogle Scholar
Kellerhals, T., et al. (2010). Ammonium concentration in ice cores: A new proxy for regional temperature reconstruction?, J. Geophys. Res., Atmos., 115, 2156–2202.CrossRefGoogle Scholar
Kinnard, C., et al. (2020). Mass balance and climate history of a high-altitude glacier, Desert Andes of Chile. Front. Earth Sci., 8, 40. https://doi.org/10.3389/feart.2020.00040.Google Scholar
Knusel, S., et al. (2003). Dating of two nearby ice cores from the Illimani, Bolivia. J. Geophys. Res., 108(D6), 4181–4191.Google Scholar
Kock, S. T., et al. (2020). Multi-centennial-scale variations of South American summer monsoon intensity in the southern central Andes (24–27°S) during the late Holocene. Geophys. Res. Lett., 47(4), e2019GL084157.CrossRefGoogle Scholar
Kohshima, S., et al. (2007). Estimation of net accumulation rate at a Patagonian glacier by ice core analyses using snow algae. Glob. Planet. Change, 59(2007), 236–244.CrossRefGoogle Scholar
Kuhn, M. (1989). The response of the equilibrium line altitude to climatic fluctuations: Theory and observations. In Oerlemans, J. (ed), Glacier Fluctuations and Climate Change. Kluwer Academic Publishers, Amsterdam, pp. 407–417.Google Scholar
Le Meur, E., et al. (2004). Glacier flow modelling: A comparison of the Shallow Ice Approximation and the full-Stokes solution. Comptes Rendus Physique, 5, 709–722.CrossRefGoogle Scholar
Levy, R. H., et al. (2018). A high-resolution climate record spanning the past 17 000 years recovered from Lake Ohau, South Island, New Zealand. Scientific Drilling (Hokkaido, Japan), 24, 41–50.Google Scholar
Lopez, P., et al. (2010). A regional view of fluctuations in glacier length in southern South America. Glob. Planet. Change, 71, 85–108. https://doi.org/10.1016/j.gloplacha.2009.12.009.CrossRefGoogle Scholar
MacAyeal, D. R. (2019). Revisiting Weertman’s tombstone bed. Ann. Glaciol., 60(80), 21–29. https://doi.org/10.1017/aog.2019.31.CrossRefGoogle Scholar
MacDonell, S., et al. (2013). Meteorological drivers of ablation processes on a cold glacier in the semi-arid Andes of Chile. Cryosphere, 7, 1513–1526. https://doi.org/10.5194/tc-7-1513-2013.CrossRefGoogle Scholar
Mackintosh, A. N., et al. (2017a). Reconstructing climate from glaciers. Ann. Rev. Earth Planet. Sci., 2017. 45, 649–80. https://doi.org/10.1146/annurev-earth-063016-020643.CrossRefGoogle Scholar
Mackintosh, A. N., et al. (2017b). Regional cooling caused recent New Zealand glacier advances in a period of global warming. Nat. Commun., 8, 14202. https://doi.org/10.1038/ncomms14202.CrossRefGoogle Scholar
Mahaney, W. C., et al. (2000a). Late quaternary deglaciation and neoglaciation of the Humboldt massif, northern Venezuela. Z. Geomorphol. Suppl., 122, 209–226.Google Scholar
Mahaney, W. C., et al. (2000b). Stratotype for the Mérida Glaciation at Pueblo Llano in the northern Venezuelan Andes. J. S. Am. Earth Sci., 13, 761–774.CrossRefGoogle Scholar
Malz, P., et al. (2018). Elevation and mass changes of the Southern Patagonia icefield derived from TanDEM-X and SRTM data. Remote Sens., 10, 188. https://doi.org/10.3390/rs10020188.CrossRefGoogle Scholar
Marshall, S. J. and Losic, M. (2011). Temperature lapse rates in glacierized basins. In Singh, V. P., et al. (eds.), Encyclopedia of Snow, Ice and Glaciers. Springer, Dordrecht, pp. 1145–1150. https://doi.org/10.1007/978-90-481-2642-2.Google Scholar
Masiokas, M. H., et al. (2020). A review of the current state and recent changes of the Andean cryosphere. Front. Earth Sci., 8, 99. https://doi.org/10.3389/feart.2020.00099.Google Scholar
Matejka, M., et al. (2021). High-resolution numerical modelling of near surface atmospheric fields in the complex terrain of James Ross Island, Antarctic Peninsula. Atmosphere, 12, 360. https://doi.org/10.3390/atmos12030360.CrossRefGoogle Scholar
Matsuoka, K. and Naruse, R. (1999). Mass balance features derived from a firn core at Hielo Patagonico Norte, South America. Arct. Antarct. Alpine Res., 31, 333–340.CrossRefGoogle Scholar
Mayewski, P. A., et al. (2016). Initial reconnaissance for a South Georgia ice core. J. Glaciol., 62, 54–61.CrossRefGoogle Scholar
Meier, W. J.-H., et al. (2018). An updated multi-temporal glacier inventory for the Patagonian Andes with changes between the Little Ice Age and 2016. Front. Earth Sci., 6, 62. https://doi.org/10.3389/feart.2018.00062.CrossRefGoogle Scholar
Melkonian, A. K., et al. (2013). Satellite-derived volume loss rates and glacier speeds for the Cordillera Darwin Icefield, Chile. Cryosphere, 7, 823–839. https://doi.org/10.5194/tc-7-823-2013.CrossRefGoogle Scholar
Menzies, J., (ed). (1995). Modern Glacial Environments: Processes, Dynamics and Sediments. Glacial Environments Volume 1. Butterworth Heinemann,. Oxford, 621pp.Google Scholar
Menzies, J. and van der Meer, J. (eds.). (2017). Past Glacial Environments, Elsevier, Amsterdam, 858pp.Google Scholar
Minowa, M., et al. (2017). Seasonal variations in ice-front position controlled by frontal ablation at Glaciar Perito Moreno, the Southern Patagonia Icefield. Front. Earth Sci., 5, 1. https://doi.org/10.3389/feart.2017.00001.CrossRefGoogle Scholar
Mölg, T., et al. (2008). Mass balance of a slope glacier on Kilimanjaro and its sensitivity to climate. Int. J. Climatol., 28, 881–892. https://doi.org/10.1002/joc.1589.CrossRefGoogle Scholar
Mölg, T., et al. (2009a). Quantifying climate change in the tropical midtroposphere over East Africa from glacier shrinkage on Kilimanjaro. J. Clim., 22, 4162–4182.CrossRefGoogle Scholar
Mölg, T., et al. (2009b). Temporal precipitation variability versus altitude on a tropical high mountain: Observations and mesoscale atmospheric modeling. Quart. J. Roy. Meteor. Soc., 135(643), 1439–1455.CrossRefGoogle Scholar
Mölg, T., et al. (2009c). Solar radiation, cloudiness and longwave radiation over low-latitude glaciers: Implications for mass balance modeling. J. Glaciol., 55, 292–302.CrossRefGoogle Scholar
Morgenstern, U., et al. (2004). Ice core research in the New Zealand Southern Alps. Geophys. Res. Abs., 6. https://doi:10.1016/j.gca.2006.06.865.Google Scholar
Mourre, L., et al. (2016). Spatio-temporal assessment of WRF, TRMM and in situ precipitation data in a tropical mountain environment (Cordillera Blanca, Peru). Hydrol. Earth Syst. Sci., 20, 125–141.CrossRefGoogle Scholar
Moya-Alvarez, A. S., et al. (2018). Extreme rainfall forecast with the WRF-ARW model in the Central Andes of Peru. Atmosphere, 9, 362. https://doi.org/10.3390/atmos9090362.CrossRefGoogle Scholar
Nesje, A. (1992). Topographical effects on the equilibrium-line altitude on glaciers. GeoJournal, 27(4), 383–391.CrossRefGoogle Scholar
Nye, J. F. (1965). A numerical method of inferring the budget history of a glacier from its advance and retreat. J. Glaciol., 35, 355–369.Google Scholar
Oerlemans, J. (1994). Quantifying global warming from the retreat of glaciers. Science, 264(5156), 243–245.CrossRefGoogle ScholarPubMed
Oerlemans, J. (2001). Glaciers and Climate Change. A.A. Balkema, Rotterdam.Google Scholar
Ohmura, A. (2001). Physical basis for the temperature-based melt-index method. J. Appl. Meteorol., 40(4), 753–761.2.0.CO;2>CrossRefGoogle Scholar
Ohmura, A., et al. (1992). Climate at the equilibrium line of glaciers. J. Glaciol., 38(130), 397–411.CrossRefGoogle Scholar
Osmaston, H. (2005). Estimates of glacier equilibrium line altitudes by the Area x Altitude, the Area x Altitude Balance Ratio and the Area x Altitude Balance Index methods and their validation. Quat. Int., 138–139, 22–31.Google Scholar
Permana, D. S., et al. (2016). Tropical West Pacific moisture dynamics and climate controls on rainfall isotopic ratios in southern Papua, Indonesia. J. Geophys. Res., Atmos., 121, 2222–2245. https://doi.org/10.1002/2015JD023893.Google Scholar
Permana, D. S., et al. (2019). Disappearance of the last tropical glaciers in the Western Pacific Warm Pool (Papua, Indonesia) appears imminent. Proc. Nat. Acad. Sci. USA., 116(52), 26382–26388. https://doi.org/10.1073/pnas.1822037116.CrossRefGoogle ScholarPubMed
Pierrehumbert, R. T. (2011). Principles of Planetary Climate. Cambridge University Press, Cambridge.Google Scholar
Porter, S. C. (1975). Equilibrium-line altitudes of late Quaternary glaciers in the southern Alps, New Zealand. Quat. Res., 5, 27–47.CrossRefGoogle Scholar
Porter, S. C. (2001). Snowline depression in the tropics during the last glaciation. Quat. Sci. Rev., 20, 1067–1091.Google Scholar
Prentice, M. L., et al. (2005). An evaluation of snowline data across New Guinea during the last major glaciation, and area-based glacier snowlines in the Mt. Jaya region of Papua, Indonesia, during the Last Glacial Maximum. Quat. Int., 138–139, 93–117.Google Scholar
Purdie, H., et al. (2011). Inter-annual variability in net accumulation on Tasman Glacier, New Zealand, and its relationship with climate. Glob. Planet. Change., 77, 142–152.CrossRefGoogle Scholar
Ramirez, E., et al. (2003). A new Andean deep ice core from the Illimani (6350 m), Bolivia. Ear. Plan. Sci. Lett., 212(3–4), 337–350.Google Scholar
Raper, S. C. B. and Braithwaite, R. J. (2009). Glacier volume response time and its links to climate and topography based on a conceptual model of glacier hypsometry. Cryosphere, 3, 183–194. www.the-cryosphere.net/3/183/2009/.CrossRefGoogle Scholar
Rimac, A., et al. (2017). Numerical simulations of glacier evolution performed using flow-line models of varying complexity. Geosci. Model Dev. Discuss. https://doi.org/10.5194/gmd-2017-67.Google Scholar
Roe, G. H. (2011). What do glaciers tell us about climate variability and climate change? J. Glaciol., 57(203), 567–578.CrossRefGoogle Scholar
Roe, G. H., et al. (2016). Centennial glacier retreat as categorical evidence of regional climate change. Nat. Geosci., 10, 95–99. https://doi.org/10.1038/NGEO2863.Google Scholar
Rozanski, K., et al. (1993). Isotopic patterns in modern global precipitation. In Swart, P. K., et al. (eds.), Climate Change in Continental Isotopic Records. Am. Geophys. Union, Washington, DC, pp. 1–36.Google Scholar
Sagredo, E. A. and Lowell, T. V. (2012). Climatology of Andean glaciers: A framework to understand glacier response to climate change. Glob. Planet. Change, 8, 101–109. https://doi.org/10.1016/j.gloplacha.2012.02.010.Google Scholar
Sagredo, E. A., et al. (2014). Sensitivities of the equilibrium line altitude to temperature and precipitation changes across the Andes. Quat. Res., 81, 355–366.CrossRefGoogle Scholar
Sagredo, E. A., et al. (2016). Equilibrium line altitudes along the Andes during the Last millennium: Paleoclimatic implications. Holocene, 1–15. https://doi.org/10.1177/0959683616678458.Google Scholar
Schittek, K., et al. (2015). Holocene environmental changes in the highlands of the southern Peruvian Andes (14°S) and their impact on pre-Columbian cultures. Clim. Past, 11, 27–44. https://doi.org/10.5194/cp-11-27-2015.CrossRefGoogle Scholar
Schwikowski, M., et al. (2006). A potential high-elevation ice-core site at Hielo Patagonico Sur. Ann. Glaciol., 43, 8–13.CrossRefGoogle Scholar
Schwikowski, M., et al. (2013). Net accumulation rates derived from ice core stable isotope records of Pío XI glacier, Southern Patagonia Icefield. Cryosphere, 7, 1635–1644. https://doi.org/10.5194/tc-7-1635-2013.CrossRefGoogle Scholar
Shiraiwa, T., et al. (2002). High net accumulation rates at Campo de Hielo Patagonico Sur, South America, revealed by analysis of a 45.97 m long ice core. Ann. Glaciol., 35, 84–90.CrossRefGoogle Scholar
Skamarock, W. C., et al. (2019). A Description of the Advanced Research WRF Version 4. Technical Note 556. National Center for Atmospheric Research, Boulder, CO.Google Scholar
Skvarca, P. and De Angelis, H. (2003). First cloud-free Landsat TM image mosaic of Hielo Patagónico Sur, southwestern Patagonia, South America. (Contribución del Instituto Antártico Argentino 535) Dirección Nacional del Antártico, Buenos Aires.Google Scholar
Somos-Valenzuela, M. and Manquehual-Cheuque, F. (2020). Evaluating multiple WRF configurations and forcing over the Northern Patagonian Icecap (NPI) and Baker River Basin. Atmosphere, 11, 815. https://doi.org/10.3390/atmos11080815.CrossRefGoogle Scholar
Sugden, D. E. and John, B. S. (1976). Glaciers and Landscape: A Geomorphological Approach. Edward Arnold, London, 1986.Google Scholar
Temme, F., et al. (2020). Flow regimes and föhn types characterize the local climate of Southern Patagonia. Atmosphere, 11, 899. https://doi.org/10.3390/atmos11090899.CrossRefGoogle Scholar
Thomas, E. R., et al. (2021). Physical properties of shallow ice cores from Antarctic and sub-Antarctic islands. Cryosphere, 15, 1173–1186. https://doi.org/10.5194/tc-15-1173-2021.CrossRefGoogle Scholar
Thompson, L. G. (2000). Ice core evidence for climate change in the Tropics: Implications for our future. Quat. Sci. Rev., 19, 19–35.CrossRefGoogle Scholar
Thompson, L. G., et al. (1998). A 25000-year tropical climate history from Bolivian ice cores, Science, 282, 1858–1864.CrossRefGoogle ScholarPubMed
Thompson, L. G., et al. (2000). Ice-core palaeoclimate records in tropical South America since the Last Glacial Maximum. J. Quaternary Sci., 15, 377–394.3.0.CO;2-L>CrossRefGoogle Scholar
Thompson, L. G., et al. (2002). Kilimanjaro ice core records: Evidence of Holocene climate change in Tropical Africa. Science, 298, 589–593.CrossRefGoogle ScholarPubMed
Thompson, L. G., et al. (2013). Annually resolved ice core records of Tropical climate variability over the past ~1800 Years. Science, 340, 945–950.CrossRefGoogle ScholarPubMed
Uglietti, C., et al. (2016). Radiocarbon dating of glacier ice: Overview, optimisation, validation and potential. Cryosphere, 10, 3091–3105. https://doi.org/10.5194/tc-10-3091-2016.CrossRefGoogle Scholar
van der Veen, C. J. (2013). Fundamentals of Glacier Dynamics, 2nd ed. CRC Press, Boca Raton, 403pp. https://doi.org/10.1201/b14059.CrossRefGoogle Scholar
Verfaillie, D., et al. (2015). Recent glacier decline in the Kerguelen Islands (49°S, 69°E) derived from modeling, field observations, and satellite data. J. Geophys. Res., 120, 637–654.Google Scholar
Villarroel, C., et al. (2013). Modeling near-surface air temperature and precipitation using WRF with 5-km resolution in the northern Patagonia Icefield: A pilot simulation. Int. J. Geosci., 04(08), 1193–1199. https://doi.org/10.4236/ijg.2013.48113.CrossRefGoogle Scholar
Vimeux, F., et al. (2008). A promising location in Patagonia for paleoclimate and paleoenvironmental reconstructions revealed by a shallow firn core from Monte San Valent́ın (Northern Patagonia Icefield, Chile). J. Geophys. Res., 113, D16118. https://doi.org/10.1029/2007JD009502.Google Scholar
Vimeux, F., et al. (2009). Climate variability during the last 1000 years inferred from Andean ice cores: A review of methodology and recent results. Palaeogeography, Palaeoclimatology, Palaeoecology, 281, 229–241. https://doi.org/10.1016/j.palaeo.2008.03.054.CrossRefGoogle Scholar
Vuille, M. and Ammann, C. (1997). Regional snowfall patterns in the high arid Andes. Clim. Change, 36, 413–423.CrossRefGoogle Scholar
Vuille, M., et al. (2003a). Modeling δ18O in precipitation over the tropical Americas, part I, Interannual variability and climatic controls, J. Geophys. Res., 108. https://doi.org/10.1029/2001JD002038.CrossRefGoogle Scholar
Vuille, M., et al. (2003b). Modeling δ18O in precipitation over the tropical Americas, part II, Simulation of the stable isotope signal in Andean ice cores, J. Geophys. Res., 108. https://doi.org/10.1029/2001JD002039.CrossRefGoogle Scholar
Weertman, J. (1957). On the sliding of glaciers. J. Glaciol., 3(21), 33–38.CrossRefGoogle Scholar
Wijngaard, R. R., et al. (2019). Modeling the response of the Langtang Glacier and the Hintereisferner to a changing climate since the Little Ice Age. Front. Earth Sci., 7, 143. https://doi.org/10.3389/feart.2019.00143.CrossRefGoogle Scholar
Wilson, R., et al. (2018). Glacial lakes of the Central and Patagonian Andes. Glob. Planet. Change., 162, 275–291.Google Scholar
Yamada, T. (1987). Glaciological characteristics revealed by 37.6-m deep ice core drilled at the accumulation area of San Rafael Glacier, the Northern Patagonian Icefield. Bull. Glacier Res., 4, 59–67.Google Scholar
Zalazar, L., et al. (2017). Glaciares de Argentina: resultados preliminares del inventario Nacional de Glaciares. Rev. Glac. Ecosist. Montaña, 2, 13–22.Google Scholar
Zekollari, H., et al. (2020). On the imbalance and response time of glaciers in the European Alps. Geophys. Res. Lett., 47, e2019GL085578. https://doi.org/10.1029/2019GL085578.CrossRefGoogle Scholar
Zemp, M., et al. (2013). Reanalysing glacier mass balance measurement series. Cryosphere, 7, 1227–1245. https://doi.org/10.5194/tc-7-1227-2013.CrossRefGoogle Scholar
Abermann, J., et al. (2014). Albedo variations and the impact of clouds on glaciers in the Chilean semi-arid Andes. J. Glaciol., 60(219), 183–191. https://doi.org/10.3189/2014JoG13J094.CrossRefGoogle Scholar
Abram, N. J., et al. (2013). Acceleration of snow melt in an Antarctic Peninsula ice core during the twentieth century. Nat. Geosci., 6(5), 404–411. https://doi.org/10.1038/NGEO1787.CrossRefGoogle Scholar
Ackerley, D., et al. (2011). Using synoptic type analysis to understand New Zealand climate during the Mid-Holocene. Clim. Past, 7, 1189–1207. https://doi.org/10.5194/cp-7-1189-2011.CrossRefGoogle Scholar
Anderson, B. and Mackintosh, A. (2006). Temperature change is the major driver of late-glacial and Holocene glacier fluctuations in New Zealand. Geology, 34, 2, 121–124. https://doi.org/10.1130/G22151.1.CrossRefGoogle Scholar
Anderson, B. and Mackintosh, A. (2012). Controls on mass balance sensitivity of maritime glaciers in the Southern Alps, New Zealand: The role of debris cover. J. Geophys. Res., 117, F01003. https://doi.org/10.1029/2011JF002064.Google Scholar
Anderson, B. M., et al. (2008). Response of Franz Josef Glacier Ka Roimata o Hine Hukatere to climate change. Global Planet. Change, 63, 23–30. https://doi.org/10.1016/j.gloplacha.2008.04.003.CrossRefGoogle Scholar
Aniya, M., (2013). Holocene glaciations of Hielo Patagonico (Patagonia Icefield), south America: A brief review. Geochem. J., 47(2), 97–105.CrossRefGoogle Scholar
Araneda, A., et al. (2007). Historical records of San Rafael glacier advances (North Patagonian Icefield): Another clue to ‘Little Ice Age’ timing in southern Chile? Holocene, 17, 987–998. https://doi.org/10.1177/0959683607082414.Google Scholar
Aravena, J. C. (2007). Reconstructing climate variability using tree rings and glacier fluctuations in the southern Chilean Andes. Ph.D. thesis, University of Western Ontario, 236.Google Scholar
Aravena, J. C., et al. (2002). Tree-ring growth patterns and temperature reconstruction from nothofagus pumilio (fagaceae) forests at the upper tree line of southern Chilean Patagonia. Rev. Chil. Hist. Nat. 75, 361–376. https://doi.org/10.4067/S0716-078X2002000200008.Google Scholar
Aravena, J.-C. and Luckman, B. H. (2009). Spatio-temporal rainfall patterns in Southern South America. Int. J. Climatol., 29, 2106–2120. https://doi.org/10.1002/joc.1761.CrossRefGoogle Scholar
Arias, P. A., et al. (2021). Hydroclimate of the Andes Part II: Hydroclimate variability and sub- continental patterns. Front. Earth Sci., 8, 505467. https://doi.org/10.3389/feart.2020.505467.CrossRefGoogle Scholar
Bakke, J., et al. (2021). Long-term demise of sub-Antarctic glaciers modulated by the Southern Hemisphere Westerlies. Sci. Rep., 11, 8361. https://doi.org/10.1038/s41598-021-87317-5.CrossRefGoogle ScholarPubMed
Bannister, D. and King, J. (2015). Föhn winds on South Georgia and their impact on regional climate. Weather, 70, 324–329.CrossRefGoogle Scholar
Bannister, D. and King, J. (2019). The characteristics and temporal variability of föhn winds at King Edward Point, South Georgia. Int. J. Climatol., 40, 2778–2794. https://doi.org/10.1002/joc.6366.Google Scholar
Barcaza, G., et al. (2009). Satellite-derived equilibrium lines in Northern Patagonia Icefield, Chile, and their implications to glacier variations. Arct. Antarct. Alp. Res., 41, 2, 174–182. https://doi.org/10.1657/1938-4246-41.2.174.CrossRefGoogle Scholar
Barichivich, J., et al. (2009). Climate signals in high elevation tree- rings from the semiarid Andes of north-central Chile: Responses to regional and large-scale variability. Palaeogeogr. Palaeoclimatol. Palaeoecol., 281, 320–333.CrossRefGoogle Scholar
Barrows, T. J., et al. (2002). The timing of the Last Glacial Maximum in Australia. Quat. Sci. Rev., 21, 159–173.CrossRefGoogle Scholar
Barrows, T. T., et al. (2007). Long term sea surface temperature and climate change in the Australian-New Zealand region. Paleoceanogr., 22, PA2215.CrossRefGoogle Scholar
Baumann, S., et al. (2021). Updated inventory of glacier ice in New Zealand based on 2016 satellite imagery. J. Glaciol., 67(261), 13–26. https://doi.org/10.1017/jog.2020.78.CrossRefGoogle Scholar
Bentley, M. J., et al. (2007). Glacial geomorphology and chronology of deglaciation, South Georgia, sub-Antarctic. Quat. Sci. Rev., 26, 644–677.CrossRefGoogle Scholar
Bentley, M. J., et al. (2011). Rapid deglaciation of Marguerite bay, western Antarctic Peninsula in the early Holocene. Quat. Sci. Rev., 30, 3338–3349.CrossRefGoogle Scholar
Berg, S., et al. (2019). Holocene glacier fluctuations and environmental changes in subantarctic South Georgia inferred from a sediment record from a coastal inlet. Quaternary Research, 91(1), 132–148.CrossRefGoogle Scholar
Berthier, E., et al. (2009). Ice wastage on the Kerguelen Islands (49° S, 69° E) between 1963 and 2006. J. Geophys. Res., 114, F03005. https://doi.org/10.1029/2008JF001192.Google Scholar
Bertrand, S., et al. (2014). Late Holocene covariability of the southern westerlies and sea surface temperature in northern Chilean Patagonia. Quat. Sci. Rev., 105, 195–208.CrossRefGoogle Scholar
Bertrand, S., et al. (2017). Postglacial fluctuations of Cordillera Darwin glaciers (southernmost Patagonia) reconstructed from Almirantazgo fjord sediments. Quat. Sci. Rev., 177, 265–275.CrossRefGoogle Scholar
Bown, F., et al. (2008). Recent glacier variations at the Aconcagua basin, central Chilean Andes. Ann. Glaciol., 48, 43–48.CrossRefGoogle Scholar
Bown, F., et al. (2014). Chapter 28: First glacier inventory and recent glacier variations on Isla Grande de Tierra del Fuego and adjacent islands in Southern Chile. In Kargel, J. S., et al. (eds.), Global Land Ice Measurements from Space. Springer-Praxis, pp. 661–674. https://doi.org/10.1007/978-3-540-79818-7_28.Google Scholar
Bradley, R. S. (1999). Paleoclimatology: Reconstructing Climates of the Quaternary, 2nd ed. International Geophysics Press, Volume 64, 610pp, Academic Press, San Diego.Google Scholar
Bradley, R. S., et al. (2003). Low latitude ice cores record Pacific sea surface temperatures, Geophys. Res. Lett., 30(4), 1174. https://doi.org/10.1029/2002GL016546.CrossRefGoogle Scholar
Bradley, R. S., et al. (2009). Recent changes in freezing level heights in the Tropics with implications for the deglacierization of high mountain regions. Geophys. Res. Lett., 36, L17701. https://doi.org/10.1029/2009GL037712.CrossRefGoogle Scholar
Braun, M. H., et al. (2019). Constraining glacier elevation and mass changes in South America. Nat. Clim. Change, 9, 130–136. https://doi.org/10.1038/s41558-018-0375-7.CrossRefGoogle Scholar
Bravo, C., et al. (2015). Modelled glacier equilibrium line altitudes during the mid-Holocene in the southern mid-latitudes. Clim. Past, 11, 1575–1586. https://doi.org/10.5194/cp-11-1575-2015.CrossRefGoogle Scholar
Carrivick, J. L. and Chase, S. E. (2011). Spatial and temporal variability of annual glacier equilibrium line altitudes in the Southern Alps, New Zealand. NZ J. Geol. Geophys., 54(4), 415–429.Google Scholar
Carrivick, J. L., et al. (2020). Ice thickness and volume changes across the Southern Alps, new Zealand, from the little ice age to present. Sci. Rep., 10, 13392. https://doi.org/10.1038/s41598-020-70276-8.Google ScholarPubMed
Casassa, G., et al. (1997). A century-long recession record of Glaciar O’Higgins, Chilean Patagonia. Ann. Glaciol., 24, 106–110.CrossRefGoogle Scholar
Cavalieri, D. J. and Parkinson, C. L. (2008). Antarctic sea ice variability and trends, 1979–2006. J. Geophys. Res., 113, C07004. https://doi.org/10.1029/2007JC004564.Google Scholar
Charton, J., et al. (2021). A debris-covered glacier at Kerguelen (49°S, 69°E) over the past 15 000 years. Antarct. Sci., 33, 1, 103–115.CrossRefGoogle Scholar
Chinn, T. J. H. and Whitehouse, I. E. (1980). Glacier snow line variations in the Southern Alps, New Zealand. IAHS-AISH Publ. 126, 219–228.Google Scholar
Chinn, T. J. H. (1996). New Zealand glacier responses to climate change of the past century. NZ. J. Geol. Geophys., 39, 415–428.CrossRefGoogle Scholar
Chinn, T. J. H. (2000). Glaciers of Irian Jaya, Indonesia, and New Zealand – Glaciers of New Zealand. In USGS, ed. Satellite Image Atlas of glaciers of the world, USGS.Google Scholar
Chinn, T. J. H. (2001). Distribution of the glacial water resources of New Zealand. J. Hydrol. (NZ), 40(2), 139–187.Google Scholar
Chinn, T. J. H. and Dillon, A. (1987). Observations on a debris-covered polar glacier ‘Whisky Glacier’, James Ross Island, Antarctic Peninsula, Antarctica. J. Glaciol., 33, 1–11.CrossRefGoogle Scholar
Chinn, T. J. H., et al. (2005). Recent glacier advances in Norway and New Zealand: A comparison of their glaciological and meteorological causes. Geogr. Ann. A: Phys. Geogr., 87(1), 141–157.CrossRefGoogle Scholar
Chinn, T. J. H., et al. (2012). Annual ice volume changes 1976–2008 for the New Zealand Southern Alps. Glob. Planet. Change, 92–93, 105–118. https://doi.org/10.1016/j.gloplacha.2012.04.002.Google Scholar
Chinn, T. J., et al. (2014). New Zealand’s glaciers. Chapter 29. In Kargel, J. S., et al. (eds.), Global Land Ice Measurements from Space, Springer Praxis Books, Ó Springer-Verlag, Berlin Heidelberg, pp. 675–715. https://doi.org/10.1007/978-3-540-79818-7_29.Google Scholar
Clapperton, C. M., et al. (1989). Late-glacial and Holocene glacier fluctuations and environmental change on South Georgia, Southern Ocean. Quaternary Research, 31, 210–228.CrossRefGoogle Scholar
Clare, G. R., et al. (2002). Interannual variation in end-of-summer- snowlines of the Southern Alps of New Zealand, in response to changes in Southern Hemisphere atmospheric circulation and sea surface temperature patterns. Int. J. Climatol., 22(1), 121–128.CrossRefGoogle Scholar
Cogley, J. G., et al. (2014). Remote sensing of glaciers of the subantarctic islands. Chapter 32. In Kargel, J. S., et al. (eds.), Global Land Ice Measurements from Space. Springer Praxis Books, 759. Ó Springer-Verlag, Berlin Heidelberg, pp. 759–780. https://doi.org/10.1007/978-3-540-79818-7_32.Google Scholar
Condom, T., et al. (2007). Computation of the space and time evolution of equilibrium-line altitudes on Andean glaciers (10°N–55°S). Global and Planetary Change, 59(2007), 189–202.CrossRefGoogle Scholar
Cook, A. J., et al. (2005). Retreating glacier fronts on the Antarctic Peninsula over the past half-century. Science, 308, 541–544.CrossRefGoogle ScholarPubMed
Cook, A. J., et al. (2010). Glacier retreat on South Georgia and implications for the spread of rats. Antarct. Sci., 22(3), 255–263.CrossRefGoogle Scholar
Cuffey, K. M. and Patterson, W. S. B. (2010). The Physics of Glaciers. Butterworth-Heinemann, Oxford.Google Scholar
Cullen, N. J., et al. (2013). A century of ice retreat on Kilimanjaro: The mapping reloaded. Cryosphere Discussions, 7, 419–431. https://doi.org/10.5194/tc-7-419-2013.Google Scholar
Cullen, N. J., et al. (2019). The influence of weather systems in controlling mass balance in the Southern Alps of New Zealand. J. Geophys. Res. Atmos., 124, 4514–4529. https://doi.org/10.1029/2018JD030052.CrossRefGoogle Scholar
Davies, B. and Glasser, N. (2012). Accelerating shrinkage of Patagonian glaciers from the Little Ice Age (AD 1870) to 2011. J. Glaciol., 58, 1063–1084. https://doi.org/10.3189/2012JoG12J026.CrossRefGoogle Scholar
Davies, B. J., et al. (2014). Modelled glacier response to centennial temperature and precipitation trends on the Antarctic Peninsula. Nat. Clim. Change, 4, 993–998.Google Scholar
Davies, B. J., et al. (2020). The evolution of the Patagonian Ice Sheet from 35 ka to the present day (PATICE). Earth-Sci. Rev., 204, 103152.CrossRefGoogle Scholar
Deline, P., et al. (2024). Mapping of morainic complexes and reconstruction of glacier dynamics north-east of Cook Ice Cap, Kerguelen Archipelago (49°S). Antarct. Sci, 36(2), 75–100.CrossRefGoogle Scholar
Dickens, W. A., et al. (2019). Enhanced glacial discharge from the eastern Antarctic Peninsula since the 1700s associated with a positive Southern Annular Mode. Sci. Rep., 9, 14606. https://doi.org/10.1038/s41598-019-50897-4.CrossRefGoogle ScholarPubMed
Doughty, A. M., et al. (2015). Mismatch of glacier extent and summer insolation in Southern Hemisphere mid-latitudes. Geology, 43(5), 407–410. https://doi.org/10.1130/G36477.1.CrossRefGoogle Scholar
Doughty, A. M., et al. (2017). An exercise in glacier length modeling: Interannual climatic variability alone cannot explain Holocene glacier fluctuations in New Zealand. Earth and Planetary Science Letters, 470(2017), 48–53. http://dx.doi.org/10.1016/j.epsl.2017.04.032.CrossRefGoogle Scholar
Dowling, L., et al. (2021). Local summer insolation and greenhouse gas forcing drove warming and glacier retreat in New Zealand during the Holocene. Quat. Sci. Rev., 266, 107068. https://doi.org/10.1016/j.quascirev.2021.107068.CrossRefGoogle Scholar
Dussaillant, I., et al. (2018). Geodetic Mass Balance of the Northern Patagonian Icefield from 2000 to 2012 Using Two Independent Methods. Front. Earth Sci., 6, 8. https://doi.org/10.3389/feart.2018.00008.Google Scholar
Dussaillant, I., et al. (2019). Two decades of glacier mass loss along the Andes. Nat. Geosci. 12, 802–808. https://doi.org/10.1038/s41561-019-0432-5.Google Scholar
Eaves, S. R., et al. (2016). The Last Glacial Maximum in the central North Island, New Zealand: Palaeoclimate inferences from glacier modelling. Clim. Past, 12, 943–960.CrossRefGoogle Scholar
Eaves, S. R., et al. (2019). Late-glacial and Holocene glacier fluctuations in North Island, New Zealand. Quat. Sci. Rev., 223, 105914. https://doi.org/10.1016/j.quascirev.2019.105914.CrossRefGoogle Scholar
Espinoza, J. C., et al. (2015). Rainfall hotspots over the southern tropical Andes: Spatial distribution, rainfall intensity, and relations with large-scale atmospheric circulation. Water Resour. Res., 51, 3459–3475. https://doi.org/10.1002/2014WR016273.CrossRefGoogle Scholar
Falaschi, D., et al. (2017). Mass changes of Alpine Glaciers at the Eastern Margin of the Northern and Southern Patagonian Icefields between 2000 and 2012. J. Glaciol., 63, 258–272. https://doi.org/10.1017/jog.2016.136.CrossRefGoogle Scholar
Falaschi, D., et al. (2019). New evidence of glacier surges in the Central Andes of Argentina and Chile. Progr. Phys. Geogr., 42(6), 792–825. https://doi.org/10.1177/0309133318803014.Google Scholar
Favier, V., et al. (2004a). Glaciers of the outer and inner tropics: A different behavior but a common response to climatic forcing. Geophys. Res. Lett., 31, L16403. http://dx.doi.org/10.1029/2004GL020654.CrossRefGoogle Scholar
Favier, V., et al. (2004b). One-year measurements of surface heat budget on the ablation zone of Antizana Glacier 15, Ecuadorian Andes. J. Geophys. Res., 109, D18105. https://doi.org/10.1029/2003JD004359.Google Scholar
Favier, V., et al. (2016). Atmospheric drying as the main driver of dramatic glacier wastage in the southern Indian Ocean. Sci. Rep., 6, 32396. https://doi.org/10.1038/srep32396.CrossRefGoogle Scholar
Ferreira, G. W. S. and Reboita, M. S. (2022). A new look into the South America precipitation regimes: Observation and forecast. Atmosphere, 13, 873. https://doi.org/10.3390/atmos13060873.CrossRefGoogle Scholar
Ferri, L., et al. (2020). Ice mass loss in the Central Andes of Argentina between 2000 and 2018 derived from a new glacier inventory and satellite stereo-imagery. Front. Earth Sci., 8, 530997. https://doi.org/10.3389/feart.2020.530997.Google Scholar
Fitzgerald, N. B. and Kirkpatrick, J. B. (2020). Air temperature lapse rates and cloud cover in a hyper-oceanic climate. Antarct. Sci., 32(6), 440–453. https://doi.org/10.1017/S0954102020000309.CrossRefGoogle Scholar
Fitzharris, B. B., et al. (1992). Behaviour of New Zealand glaciers and atmospheric circulation changes over the past 130 years. Holocene, 2(2), 97–106.CrossRefGoogle Scholar
Fitzharris, B. B., et al. (1997). Glacier balance fluctuations and atmospheric circula- tion patterns over the Southern Alps, New Zealand. Int. J. Climatol., 17, 745–763.3.0.CO;2-Y>CrossRefGoogle Scholar
Fitzharris, B. B., et al. (2007). Teleconnections between Andean and New Zealand glaciers. Glob. Planet Change, 59, 159–174. https://doi.org/10.1016/j.gloplacha.2006.11.022.CrossRefGoogle Scholar
Fonseca, R. and Martín-Torres, J. (2019). High-resolution dynamical downscaling of re-analysis data over the Kerguelen Islands using the WRF model. Theoret. Appl. Climatol., 135, 1259–1277. https://doi.org/10.1007/s00704-018-2438-0.CrossRefGoogle Scholar
Frenot, Y., et al. (1997). Dating of some Holocene peat sediments and glacier fluctuations in the Kerguelen Islands. C. R. l’Acad. Sci. Ser. III Sci. Vie, 320, 7, 567–573.Google Scholar
Gabrielli, P. (2014). Deglaciated areas of Kilimanjaro as a source of volcanic trace elements deposited on the ice cap during the late Holocene. Quat. Sci. Rev., 93, 1–10. http://dx.doi.org/10.1016/j.quascirev.2014.03.007.CrossRefGoogle Scholar
Garcia, J.-L., et al. (2020). 14C and 10Be dated Late Holocene fluctuations of Patagonian glaciers in Torres del Paine (Chile, 51°S) and connections to Antarctic climate change. Quat. Sci. Rev., 246, 106541. https://doi.org/10.1016/j.quascirev.2020.106541.CrossRefGoogle Scholar
Garelick, S., et al. (2022). The dynamics of warming during the last deglaciation in high-elevation regions of Eastern Equatorial Africa. Quat. Sci. Rev., 281, 107416.CrossRefGoogle Scholar
Garreaud, R., et al. (2003). The climate of the Altiplano: Observed current conditions and mechanisms of past changes. Palaeogeogr. Palaeoclimatol. Palaeoecol., 194, 5–22.CrossRefGoogle Scholar
Garreaud, R. D. (2009). The Andes climate and weather. Adv. Geosci., 7, 1–9. https://doi.org/10.5194/adgeo-22-3-2009.Google Scholar
Garreaud, R. D., et al. (2009). Present-day South American climate. Palaeogeogr. Palaeoclimatol. Palaeoecol., 281(3–4), 180–195. http://dx.doi.org/10.1016/j.palaeo.2007.10.032.CrossRefGoogle Scholar
Garreaud, R. D., et al. (2017). The 2010–2015 megadrought in central Chile: Impacts on regional hydroclimate and vegetation. Hydrol. Earth Syst. Sci., 21, 6307–6327. https://doi.org/10.5194/hess-21-6307-2017.CrossRefGoogle Scholar
Gillett, N. P., et al. (2006). Regional climate impacts of the Southern Annular Mode. Geophys. Res. Lett., 33, L23704. https://doi.org/10.1029/2006GL027721.Google Scholar
Gonzalez, S., et al. (2018). Atmospheric Patterns over the Antarctic Peninsula. Journal of Climate, 31, 3597–3608. https://doi.org/10.1175/JCLI-D-17-0598.1.CrossRefGoogle Scholar
Goodwin, I. D., et al. (2014). Climate windows for Polynesian Voyaging to New Zealand and Easter Island. Proc. Natl. Acad. Sc., 111(41), 14716–14721. www.pnas.org/cgi/doi/10.1073/pnas.1408918111.CrossRefGoogle ScholarPubMed
Gordon, J. E. and Timmis, R. J. (1992). Glacier fluctuations on South Georgia during the 1970s and early 1980s. Antarct. Sci., 4(2), 215–226.CrossRefGoogle Scholar
Gordon, J. E., et al. (2008). Recent glacier changes and climate trends on South Georgia. Global Planet. Change, 60(1–2), 72–84.CrossRefGoogle Scholar
Graham, A. G., et al. (2017). Major advance of South Georgia glaciers during the Antarctic Cold Reversal following extensive sub-Antarctic glaciation. Nat. Commun., 8, 14798. https://doi.org/10.1038/ncomms14798.CrossRefGoogle ScholarPubMed
Grießinger, J., et al. (2018). Imprints of climate signals in a 204 year δ18O tree-ring record of Nothofagus pumilio from Perito Moreno Glacier, Southern Patagonia (50°S). Front. Earth Sci., 6, 27. https://doi.org/10.3389/feart.2018.00027.CrossRefGoogle Scholar
Grise, K. M., et al. (2018). Regional and seasonal characteristics of the recent expansion of the Tropics. J. Clim,. 31, 6839–6856. https://doi.org/10.1175/JCLI-D-18-0060.1.CrossRefGoogle Scholar
Hall, B. L. (2007). Late-Holocene advance of the Collins ice cap, King George island, South Shetland islands. Holocene, 17, 1253–1258.CrossRefGoogle Scholar
Hall, B. L. (2009). Holocene glacial history of Antarctica and the sub-Antarctic islands. Quat. Sci. Rev., 28, 2213–2230.CrossRefGoogle Scholar
Hall, B. L., et al. (2010). Constant Holocene Southern-Ocean 14C reservoir ages and ice-shelf flow rates. Earth Planet Sci. Lett., 296, 115–123.CrossRefGoogle Scholar
Hall, B. L., et al. (2011). Reduced ice extent on the western Antarctic Peninsula at 700–970 cal. yr B.P. Geology, 38, 635–638. https://doi.org/10.1130/G30932.1.Google Scholar
Hall, B. L., et al. (2019). Holocene glacier fluctuations on the northern flank of Cordillera Darwin, southernmost South America. Quat. Sci. Rev., 222, 105904.CrossRefGoogle Scholar
Hastenrath, S. (1984). The Glaciers of Equatorial East Africa. Reidel, Dordrecht, Boston, Lancaster, p. 353.CrossRefGoogle Scholar
Hastenrath, S. (1991). Climate Dynamics of the Tropics. Kluwer, Dordrecht.Google Scholar
Hastenrath, S. (2001). Variations of East African climate during the past two centuries. Climate Change, 50, 209–217.Google Scholar
Hjort, C., et al. (1998). Holocene deglaciation and climate history of the northern Antarctic Peninsula region: A discussion of correlations between the Southern and Northern hemispheres. Ann. Glaciol., 27, 110–112.CrossRefGoogle Scholar
Hodgson, D. A., et al. (2014). Terrestrial and submarine evidence for the extent and timing of the Last Glacial Maximum and the onset of deglaciation on the maritime- Antarctic and sub-Antarctic islands. Quat. Sci. Rev., 100, 137e158. https://doi.org/10.1016/j.quascirev.2013.12.001.CrossRefGoogle Scholar
Hoelzle, M., et al. (2007). The application of glacier inventory data for estimating past climate change effects on mountain glaciers: A comparison between the European Alps and the Southern Alps of New Zealand. Glob. Planet. Change, 56, 69–82.CrossRefGoogle Scholar
Hoffmann, G., et al. (2003). Coherent isotope history of Andean ice cores over the last century. Geophys. Res. Lett., 30(4), 1179. https://doi.org/10.1029/2002GL014870.CrossRefGoogle Scholar
Holmlund, P. and Fuenzalida, H. (1995). Anomalous glacier responses to 20th century climatic changes in Darwin Cordillera, southern Chile. J. Glaciol., 41, 465–473. https://doi.org/10.3189/S0022143000034808.CrossRefGoogle Scholar
Hooker, B. L. and Fitzharris, B. B. (1999). The correlation between climatic parameters and the retreat and advance of Franz Josef Glacier, New Zealand. Glob. Planet. Change, 22, 39–48.CrossRefGoogle Scholar
Hope, P. K., et al. (2006). Shifts in the synoptic systems influencing south west Western Australia. Clim. Dyn., 26, 751–764.Google Scholar
Hurrell, J. W. and van Loon, H. (1994). A modulation of the atmospheric annual cycle in the Southern Hemisphere. Tellus, 46A, 325–338.Google Scholar
Ingolfsson, O. and Hjort, C. (2002). Glacial history of the Antarctic Peninsula since the Last Glacial Maximum–a synthesis. Polar Research, 21(2), 227–234. https://doi.org/10.3402/polar.v21i2.6482.Google Scholar
Izagirre, E., et al. (2018). Glacial geomorphology of the Marinelli and Pigafetta glaciers, Cordillera Darwin Icefield, southernmost Chile. J. Maps, 14(2), 269–281. https://doi.org/10.1080/17445647.2018.1462264.CrossRefGoogle Scholar
Jackson, M. S., et al. (2020). Glacial fluctuations in tropical Africa during the last glacial termination and implications for tropical climate following the Last Glacial Maximum. Quat. Sci. Rev., 243, 106455. https://doi.org/10.1016/j.quascirev.2020.106455.CrossRefGoogle Scholar
Jara, I. A., (2019). Centennial-scale precipitation anomalies in the southern Altiplano (18◦ S) suggest an extratropical driver for the South American summer monsoon during the late Holocene. Clim. Past, 15, 1845–1859. https://doi.org/10.5194/cp-15-1845-2019.CrossRefGoogle Scholar
Jomelli, V., et al. (2007). Assessment study of lichenometric methods for dating surface. Geomorphol., 86, 131–143.CrossRefGoogle Scholar
Jomelli, V., et al. (2009). Fluctuations of Glaciers in the tropical andes over the last millennium and paleoclimatic implications: A review. Palaeogeogr. Palaeoclimatol. Palaeoecol., 281, 269–282.CrossRefGoogle Scholar
Jomelli, V., et al. (2011). Irregular tropical glacier retreat over the Holocene epoch driven by progressive warming. Nature, 474, 196–199. http://dx.doi.org/10.1038/nature10150.CrossRefGoogle ScholarPubMed
Jomelli, V., et al. (2014). A major advance of tropical Andean glaciers during the Antarctic Cold Reversal. Nature, 513, 224–228. http://dx.doi.org/10.1038/nature13546.CrossRefGoogle Scholar
Jomelli, V., et al. (2017). Sub-Antarctic glacier extensions in the Kerguelen region (49◦ S, Indian Ocean) over the past 24,000 years constrained by 36Cl moraine dating. Quat. Sci. Rev., 162, 128–144. https://doi.org/10.1016/j.quascirev.2017.03.010,2017.CrossRefGoogle Scholar
Jomelli, V., et al. (2018). Glacier extent in sub-Antarctic Kerguelen archipelago from MIS 3 period: Evidence from 36Cl dating. Quat. Sci. Rev., 183, 110–123. https://doi.org/10.1016/j.quascirev.2018.01.008,2018.CrossRefGoogle Scholar
Jordan, T. E., et al. (2019). Isotopic characteristics and paleo-climate implications of the extreme precipitation event of March 2015 in northern Chile. Andean Geol., 46(1), 1–31. https://doi.org/10.5027/andgeoV46n1-3087.Google Scholar
Jourdain, N. C., et al. (2017). Ocean circulation and sea-ice thinning induced by melting ice shelves in the Amundsen Sea. J. Geophys. Res. Oceans, 122, 2550–2573. https://doi.org/10.1002/2016JC012509.CrossRefGoogle Scholar
Kaplan, M. R., et al. (2013). The anatomy of ‘long-term’ warming since 15 kyr ago in New Zealand based on net glacier snowline rise. Geology, 41, 887–890.CrossRefGoogle Scholar
Kaplan, M. R., et al. (2016). Patagonian and southern South Atlantic view of Holocene climate. Quat. Sci. Rev., 141, 112–125. https://doi.org/10.1016/j.quascirev.2016.03.014.CrossRefGoogle Scholar
Kaplan, M. R., et al. (2020). Holocene glacier behavior around the northern Antarctic Peninsula and possible causes. Earth Planet Sci. Lett., 534, 116077.CrossRefGoogle Scholar
Kaser, G. (2001). Glacier-climate interaction at low latitudes. J. Glaciol., 47(157), 195–204.CrossRefGoogle Scholar
Kaser, G. and Osmaston, H. (2002). Tropical Glaciers. International Hydrology Series. Cambridge University Press, Cambridge, 207pp.Google Scholar
Kaser, G., et al. (2004). Modern glacier retreat on Kilimanjaro as evidence of climate change: Observations and facts. I. J. Climatol., 24, 329–339. https://doi.org/10.1002/joc.1008.Google Scholar
Kaser, G., et al. (2010). Is the decline of ice on Kilimanjaro unprecedented in the Holocene? Holocene, 20(7), 1079–1091. https://doi.org/10.1177/0959683610369498.CrossRefGoogle Scholar
Kidson, J. W. (2000). An analysis of New Zealand synoptic types and their use in defining weather regimes. Int. J. Climatol., 20, 299–316.3.0.CO;2-B>CrossRefGoogle Scholar
Kiefer, J. and Karamperidou, C. (2019). High-resolution modeling of ENSO-induced precipitation in the tropical Andes: Implications for proxy interpretation. Paleoceanogr. Paleoclimatol., 34, 217–236. https://doi.org/10.1029/2018PA003423.CrossRefGoogle Scholar
Kilian, R. and Lamy, F. (2012). A review of Glacial and Holocene paleoclimate records from southernmost Patagonia (49–55°S). Quat. Sci. Rev., 53, 1–23. https://doi.org/10.1016/j.quascirev.2012.07.017.CrossRefGoogle Scholar
Kinnard, C., et al. (2020). Mass balance and climate history of a high-altitude glacier, Desert Andes of Chile. Front. Earth Sci., 8, 40. https://doi.org/10.3389/feart.2020.00040.CrossRefGoogle Scholar
Klein, A. G. and Kincaid, J. L. (2006). Retreat of glaciers on Puncak Jaya, Irian Jaya, Indonesia, determined from 2000 and 2002 IKONOS satellite images. J. Glaciol., 52, 65–79. https://doi.org/10.3189/172756506781828818.CrossRefGoogle Scholar
Kock, S. T., et al. (2020). Multi-centennial-scale variations of South American summer monsoon intensity in the southern central Andes (24–27°S) during the late Holocene. Geophys. Res. Lett., 47, e2019GL084157. https://doi.org/10.1029/2019GL084157.CrossRefGoogle Scholar
Kropač, E., et al. (2021). A detailed, multi-scale assessment of an atmospheric river event and its impact on extreme glacier melt in the Southern Alps of New Zealand. J. Geophys. Res. Atmos., 126, e2020JD034217. https://doi.org/10.1029/2020JD034217.CrossRefGoogle Scholar
Langhamer, L., et al. (2018). Lagrangian detection of moisture sources for the Southern Patagonia Icefield (1979–2017). Front. Earth Sci., 6, 219. https://doi.org/10.3389/feart.2018.00219.CrossRefGoogle Scholar
Le Quesne, C., et al. (2009). Long-term glacier variations in the Central Andes of Argentina and Chile, inferred from historical records and tree-ring reconstructed precipitation. Palaeogeogr. Palaeoclimatol. Palaeoecol., 281, 334–344.CrossRefGoogle Scholar
Li, C., et al. (2020). Holocene dynamics of the southern westerly winds over the Indian Ocean inferred from a peat dust deposition record. Quat. Sci. Rev., 231, 106169. https://doi.org/10.1016/j.quascirev.2020.106169.CrossRefGoogle Scholar
Li, X., et al. (2015). Rossby waves mediate impacts of tropical oceans on West Antarctic atmospheric circulation in Austral winter. J. Clim., 28, 8151–8164. https://doi.org/10.1175/JCLI-D-15-0113.1.CrossRefGoogle Scholar
Little, K., et al. (2019). The role of atmospheric rivers in controlling extreme ablation and snowfall events on Brewster Glacier, Southern Alps, New Zealand. Geophys. Res. Lett., 46, 2761–2771. https://doi.org/10.1029/2018GL081669.CrossRefGoogle Scholar
Lliboutry, L. (1954). The origin of penitents. J. Glaciol., 2(15), 331–338.Google Scholar
Lliboutry, L. (1956). Nieves y Glaciares de Chile, Fundamentos de Glaciología. Universidad de Chile, Santiago, 471pp. [in Spanish].Google Scholar
Lliboutry, L. (1998). Glaciers of the Dry Andes. In Williams, R. S. J. and Ferrigno, J. G. (eds.), Satellite Image Atlas of Glaciers of the World – South America. United States Geological Survey Professional Paper 1386-I.Google Scholar
Lodolo, E., et al. (2020). The submerged footprint of Perito Moreno glacier. Sci. Rep., 10, 16437. https://doi.org/10.1038/s41598-020-73410-8.CrossRefGoogle ScholarPubMed
Lopez, P., et al. (2010). A regional view of fluctuations in glacier length in southern South America. Glob. Planet. Change, 71, 85–108. https://doi.org/10.1016/j.gloplacha.2009.12.009.CrossRefGoogle Scholar
Lorrey, A. M., et al. (2014). The Little Ice Age climate of New Zealand reconstructed from Southern Alps cirque glaciers: A synoptic type approach. Clim. Dyn., 42, 3039–3060. https://doi.org/10.1007/s00382-013-1876-8.CrossRefGoogle Scholar
MacDonnell, S., et al. (2013). Meteorological drivers of ablation processes on a cold glacier in the semi-arid Andes of Chile. Cryosphere, 7, 1513–1526. https://doi.org/10.5194/tc-7-1513-2013.CrossRefGoogle Scholar
Mackintosh, A. N., et al. (2017a). Reconstructing climate from glaciers. Annu. Rev. Earth Planet. Sci., 45, 649–680. https://doi.org/10.1146/annurev-earth-063016-020643.CrossRefGoogle Scholar
Mackintosh, A. N., et al. (2017b). Regional cooling caused recent New Zealand glacier advances in a period of global warming. Nat. Commun., 8, 14202. https://doi.org/10.1038/ncomms14202.CrossRefGoogle Scholar
Masiokas, M. H., et al. (2009). Glacier fluctuations in extratropical South America during the past 1000 years. Palaeogeogr. Palaeoclimatol. Palaeoecol., 281, 242–268. https://doi.org/10.1016/j.palaeo.2009.08.006.Google Scholar
Masiokas, M. H., et al. (2012). Snowpack variations since AD 1150 in the Andes of Chile and Argentina (30°–37°S) inferred from rainfall, tree-ring and documentary records. J. Geophys. Res., 117, D05112. https://doi.org/10.1029/2011JD016748.Google Scholar
Masiokas, M. H., et al. (2016). Snowpack variations in the central Andes of Argentina and Chile, 1951–2005: Large-scale atmospheric influences and implications for water resources in the region. J. Clim., 19, 6334–6352.Google Scholar
Masiokas, M. H., et al. (2020). A review of the current state and recent changes of the Andean cryosphere. Front. Earth Sci., 8, 99. https://doi.org/10.3389/feart.2020.00099.CrossRefGoogle Scholar
McAlpine, J. R., et al. (1983). Climate of Papua New Guinea. CSIRO in Association with ANU. Press, Canberra, Aust.Google Scholar
McKinzey, K. M., et al. (2004). A revised Little Ice Age chronology of the Franz Josef Glacier, Westland, New Zealand. J. Roy. Soc. NZ., 34(4), 381–394.CrossRefGoogle Scholar
Meier, W. J. H., et al. (2018). An updated multi-temporal glacier inventory for the Patagonian Andes with changes between the little ice age and 2016. Front. Earth Sci., 6, 8. https://doi.org/10.3389/feart.2018.00062.CrossRefGoogle Scholar
Meier, W. J.-H., et al. (2019). Late Holocene Glacial Fluctuations of Schiaparelli Glacier at Monte Sarmiento Massif, Tierra del Fuego (54◦24′ S). Geosciences, 9, 340. https://doi.org/10.3390/geosciences9080340.CrossRefGoogle Scholar
Melkonian, A. K., et al. (2013). Satellite-derived volume loss rates and glacier speeds for the Cordillera Darwin Icefield, Chile. Cryosphere, 7, 823–839. https://doi.org/10.5194/tc-7-823-2013.CrossRefGoogle Scholar
Mölg, T., et al. (2003). Solar radiation-maintained glacier recession on Kilimanjaro drawn from combined ice–radiation geometry modeling. J. Geophys. Res., 108, 4731.Google Scholar
Mölg, T., et al. (2008). Mass balance of a slope glacier on Kilimanjaro and its sensitivity to climate. Int. J. Climatol., 28, 881–892. https://doi.org/10.1002/joc.1589.CrossRefGoogle Scholar
Mölg, T., et al. (2009). Quantifying climate change in the tropical midtroposphere over East Africa from glacier shrinkage on Kilimanjaro. J. Clim., 22, 4162–4182.CrossRefGoogle Scholar
Mölg, T., et al. (2010). Glacier loss on Kilimanjaro is an exceptional case. Proc. Nat. Acad. Sci. USA., 107, 17. www.pnas.org/cgi/doi/10.1073/pnas.0913780107.CrossRefGoogle Scholar
Moragues, S., et al. (2018). Surface velocities of Upsala glacier, Southern Patagonian Andes, estimated using cross-correlation satellite imagery: 2013–2014 period. Andean Geology, 45(1), 87–103. https://doi.org/10.5027/andgeoV45n1-3034.Google Scholar
Morales, M. S., et al. (2012). Precipitation changes in the South American Altiplano since 1300 AD reconstructed by tree-rings. Clim. Past, 8, 653–666. https://doi.org/10.5194/cp-8-653-2012.CrossRefGoogle Scholar
Moreno, P. I., et al. (2018). Onset and evolution of southern annular mode-like changes at centennial timescale. Sci. Rep., 8, 3458. https://doi.org/10.1038/s41598-018-21836-6.CrossRefGoogle ScholarPubMed
Morioka, Y., et al. (2014). Role of tropical SST variability on the formation of subtropical dipoles. J. Clim., 27, 4486–4507. https://doi.org/10.1175/JCLI-D-13-00506.1.CrossRefGoogle Scholar
Nimick, D. A., et al. (2016). Latest Pleistocene and Holocene glacial events in the Colonia Valley, northern Patagonia Icefield, southern Chile. J. Quat. Sci., 31, 551–564. https://doi.org/10.1002/jqs.2847.CrossRefGoogle Scholar
Oerlemans, J. (2001). Glaciers and Climate Change. A.A. Balkema, Rotterdam.Google Scholar
Oppedal, L. T., et al. (2018). Cirque glacier on South Georgia shows centennial variability over the last 7000 years. Front. Earth Sci., 6, 2. https://doi.org/10.3389/feart.2018.00002.CrossRefGoogle Scholar
Pabón-Caicedo, J. D., et al. (2020). Observed and projected hydroclimate changes in the Andes. Front. Earth Sci., 8, 61. https://doi.org/10.3389/feart.2020.00061.CrossRefGoogle Scholar
Permana, D. S., et al. (2019). Disappearance of the last tropical glaciers in the Western Pacific Warm Pool (Papua, Indonesia) appears imminent. Proc. Nat. Acad. Sci. USA., 116(52), 26382–26388.CrossRefGoogle ScholarPubMed
Prentice, M. L. and Glidden, S. (2010). Glacier crippling and the rise of the snowline in the western New Guinea (Papua Province, Indonesia) from 1972 to 2000. In Haberle, S., et al. (eds.), Altered Ecologies: Fire, Climate and Human Influence on Terrestrial Landscapes. ANU Press, Canberra, Australia, pp. 457–471.Google Scholar
Prentice, M. L. and Hope, G. S (2015). The Paleoglacial record of paleotemperature at 600 hPa on Puncak Jayawijaya during the Little Ice Age. Unpublished manuscript.Google Scholar
Prince, H. D., et al. (2021). A climatology of atmospheric rivers in New Zealand. J. Clim., 34(11), 4383–4402. https://doi.org/10.1175/JCLI-D-20-0664.1.CrossRefGoogle Scholar
Purdie, H., et al. (2011a). Inter-annual variability in net accumulation on Tasman Glacier, New Zealand, and its relationship with climate. Glob. Planet. Chang., 77, 142–152.CrossRefGoogle Scholar
Purdie, H., et al. (2011b). Synoptic influences on snow accumulation on glaciers East and West of a topographic divide: Southern Alps, New Zealand. Arct. Antarct. Alp. Res., 43(1), 82–94.CrossRefGoogle Scholar
Purdie, H., et al. (2014). Franz Josef and Fox Glaciers, New Zealand: Historic length records. Glob. Planet. Change, 121, 41–52.CrossRefGoogle Scholar
Purdie, H., et al. (2021). Morphological changes to the terminus of a maritime glacier during advance and retreat phases: Fox Glacier/Te Moeka o Tuawe, New Zealand. Geogr. Ann. A: Phys. Geogr., 103(2), 167–185. https://doi.org/10.1080/04353676.2020.1840179.CrossRefGoogle Scholar
Putnam, A. E., et al. (2010). Glacier advance in southern middle latitudes during the Antarctic Cold Reversal. Nat. Geosci., 3, 700–704.CrossRefGoogle Scholar
Putnam, A. E., et al. (2012). Regional climate control of glaciers in New Zealand and Europe during the pre- industrial Holocene. Nat Geosci., 5, 627–630. https://doi.org/10.1038/ngeo1548.CrossRefGoogle Scholar
Rabatel, A., et al. (2005). Dating of Little Ice Age glacier fluctuations in the tropical Andes: Charquini glaciers, Bolvia, 16°S. C. R. Geosci., 337, 1311–1322. https://doi.org/10.1016/j.crte.2005.07.009, 2005a.CrossRefGoogle Scholar
Rabatel, A., et al. (2008). A chronology of the Little Ice Age in the tropical Andes of Bolivia (16° S) and its implications for climate reconstruction. Quat. Res., 70, 198–212. https://doi.org/10.1016/j.yqres.2008.02.012, 2008a.CrossRefGoogle Scholar
Rabatel, A., et al. (2013). Current state of glaciers in the tropical Andes: A multi-century perspective on glacier evolution and climate change. Cryosphere, 7(1), 81–102.CrossRefGoogle Scholar
Rainsley, E., et al. (2019). Pleistocene glacial history of the New Zealand subantarctic islands. Clim. Past, 15(2), 423–448. https://doi.org/10.5194/cp-15-423-2019.CrossRefGoogle Scholar
Rau, F. and Braun, M. (2002). The regional distribution of the dry-snow zone on the Antarctic Peninsula north of 70° S. Ann. Glaciol., 34, 95–100.CrossRefGoogle Scholar
Reynhout, S. A., et al. (2019). Holocene glacier fluctuations in Patagonia are modulated by summer insolation intensity and paced by Southern Annular Mode-like variability. Quat. Sci. Rev., 220, 178–187. https://doi.org/10.1016/j.quascirev.2019.05.029.CrossRefGoogle Scholar
Reynhout, S. R., et al. (2022). Holocene glacier history of northeastern Cordillera Darwin, southernmost South America (55°S). Quat. Res., 105, 166–181. https://doi.org/10.1017/qua.2021.45.CrossRefGoogle Scholar
Rivera, A., et al. (2012a). Glacier Jorge Montt dynamics derived from photos obtained by fixed cameras and satellite image feature tracking. Ann. Glaciol., 53(60), 147–155.CrossRefGoogle Scholar
Rivera, A., et al. (2012b). Little Ice Age advance and retreat of Glaciar Jorge Montt, Chilean Patagonia, recorded in maps, air photographs and dendrochronology. Clim. Past, 8, 403–414. https://doi.org/10.5194/cp-8-403-2012.CrossRefGoogle Scholar
Rodbell, D. T., et al. (2009). Glaciation in the Andes during the Late glacial and Holocene. Quat. Sci. Rev., 28, 2165–2212.CrossRefGoogle Scholar
Rojas-Murrillo, K., et al. (2020). ENSO and PDO related interannual variability in the north and east-central part of the Bolivian Altiplano in South America. Int. J. Climatol., 42, 2413–2439. https://doi.org/10.1002/joc.7374.Google Scholar
Rott, H., et al. (1996). Rapid collapse of Northern Larsen Ice Shelf, Antarctica. Science, 271(5250), 788–792.CrossRefGoogle Scholar
Rouault, M., et al. (2005). Climate variability at Marion Island, Southern Ocean, since 1960, J. Geophys. Res., 110, C05007. https://doi.org/10.1029/2004JC002492.Google Scholar
Ruddell, A. R. (1995). Recent glacier and climate change in the New Zealand Alps. Ph.D. thesis, University of Melbourne, Melbourne, Victoria, Australia.Google Scholar
Rudolph, E. M., et al. (2020). Early glacial maximum and deglaciation at sub-Antarctic Marion Island from cosmogenic 36Cl exposure dating. Quat. Sci. Rev., 231, 106208. https://doi.org/10.1016/j.quascirev.2020.106208.CrossRefGoogle Scholar
Ruiz, L., et al. (2012). Fluctuations of Glaciar Esperanza Norte in the north Patagonian Andes of Argentina during the past 400 yr. Clim. Past, 8, 1079–1090. https://doi.org/10.5194/cp-8-1079-2012.CrossRefGoogle Scholar
Ruiz, L., et al. (2018). Recent geodetic mass balance of Monte Tronador glaciers, northern Patagonian Andes. Cryosphere, 11, 619–634, 2017. https://doi.org/10.5194/tc-11-619-2017.Google Scholar
Russell, J., et al. (2009). Paleolimnological records of recent glacier recession in the Rwenzori Mountains, Uganda–D.R. Congo. J. Paleolimnol., 41, 253–271. https://doi.org/10.1007/s10933-008-9224-4.CrossRefGoogle Scholar
Saavedra, F., et al. (2020). Atmospheric rivers contribution to the snow accumulation over the Southern Andes (26.5 S–37.5 S). Front. Earth Sci., 261. https://doi.org/10.3389/feart.2020.00261.CrossRefGoogle Scholar
Sagredo, E. A. and Lowell, T. V. (2012). Climatology of Andean glaciers: A framework to understand glacier response to climate change. Glob. Planet. Change, 8, 101–109. https://doi.org/10.1016/j.gloplacha.2012.02.010.Google Scholar
Sagredo, E. A., et al. (2014). Sensitivity of the equilibrium line altitude across the Andes. Quat. Res., 81, 355–366.CrossRefGoogle Scholar
Sagredo, E. A., et al. (2016). Equilibrium line altitudes along the Andes during the Last millennium: Paleoclimatic implications. Holocene, 27, 1019–1033. https://doi.org/10.1177/0959683616678458.Google Scholar
Salinger, M. J., et al. (2019). Atmospheric circulation and ice volume changes for the small and medium glaciers of New Zealand’s Southern Alps mountain range 1977–2018. Int. J. Climatol., 39, 4274–4287. https://doi.org/10.1002/joc.6072.CrossRefGoogle Scholar
Sauter, T. (2020). Revisiting extreme precipitation amounts over southern South America and implications for the Patagonian Icefields. Hydrol. Earth Syst. Sci., 24, 2003–2016. https://doi.org/10.5194/hess-24-2003-2020.CrossRefGoogle Scholar
Schaefer, J. M., et al. (2009). High-frequency Holocene glacier fluctuations in New Zealand differ from the northern signature. Science, 324, 622–625. https://doi.org/10.1126/science.1169312.CrossRefGoogle ScholarPubMed
Schaefer, J. M., et al. (2015). The Southern glacial maximum 65,000 years ago and its unfinished termination. Quat. Sci. Rev., 114, 52–60. http://dx.doi.org/10.1016/j.quascirev.2015.02.009.CrossRefGoogle Scholar
Schaefer, M., et al. (2015). Quantifying mass balance processes on the Southern Patagonia Icefield. Cryosphere, 9(1), 25–35. https://doi.org/10.5194/tc-9-25-2015.CrossRefGoogle Scholar
Schneider, C., et al. (2020). Editorial: Climate impacts on glaciers and biosphere in Fuego-Patagonia. Front. Earth Sci., 8, 91. https://doi.org/10.3389/feart.2020.00091.CrossRefGoogle Scholar
Simms, A. R., et al. (2021). Evidence for a ‘Little Ice Age’ glacial advance within the Antarctic Peninsula: Examples from glacially-overrun raised beaches. Quat. Sci. Rev., 271, 107195. https://doi.org/10.1016/j.quascirev.2021.107195.CrossRefGoogle Scholar
Sinclair, K. E. and MacDonnell, S. (2016). Seasonal evolution of penitente glaciochemistry at Tapado Glacier, Northern Chile. Hydrol. Process., 30, 176–186. https://doi.org/10.1002/hyp.10531.CrossRefGoogle Scholar
Solomina, O. N., et al. (2015). Holocene glacier fluctuations. Quat. Sci. Rev., 111, 9–34. https://doi.org/10.1016/j.quascirev.2014.11.018.CrossRefGoogle Scholar
Srur, A. M., et al. (2018). Climate and Nothofagus pumilio establishment at upper treelines in the Patagonian Andes. Front. Earth Sci., 6, 57. https://doi.org/10.3389/feart.2018.00057.CrossRefGoogle Scholar
Stansell, N. D., et al. (2015). Late Glacial and Holocene glacier fluctuations at Nevado Huaguruncho in the Eastern Cordillera of the Peruvian Andes. Geology, 43, 747–750. https://doi.org/10.1130/G36735.1.CrossRefGoogle Scholar
Stansell, N. D., et al. (2017). Tropical ocean-atmospheric forcing of Late Glacial and Holocene glacier fluctuations in the Cordillera Blanca, Peru. Geophys. Res. Lett., 44, 4176–4185.CrossRefGoogle Scholar
Strand, P. D., et al. (2019). Millennial-scale pulsebeat of glaciation in the Southern Alps of New Zealand. Quat. Sci. Rev., 220, 165–177. https://doi.org/10.1016/j.quascirev.2019.07.022.CrossRefGoogle Scholar
Strelin, J., et al. (2008). Holocene glaciations in the Ema Glacier Valley, Monte Sarmiento Massif, Tierra del Fuego. Palaeogeogr. Palaeoclimatol. Palaeoecol., 260, 299–314.CrossRefGoogle Scholar
Strother, S. L., et al. (2015). Changes in Holocene climate and the intensity of Southern Hemisphere Westerly Winds based on a high-resolution palynological record from sub-Antarctic South Georgia. Holocene, 25, 263–279. https://doi.org/10.1177/0959683614557576.CrossRefGoogle Scholar
Sumner, P. D., et al. (2004). Climate change melts Marion Island’s snow and ice. S. Afr. J. Sci., 100, 395–398.Google Scholar
Taylor, R. G., et al. (2006). Recent glacial recession in the Rwenzori Mountains of East Africa due to rising air temperature. Geophys. Res. Lett., 33, L10402.Google Scholar
Thielke, A. and Mölg, T. (2019). Observed and simulated Indian Ocean Dipole activity since the mid-19th century and its relation to East African short rains. Int. J. Climatol., 39, 4467–4478. https://doi.org/10.1002/joc.6085.CrossRefGoogle Scholar
Thomas, E. R., et al. (2008). A doubling in snow accumulation in the western Antarctic Peninsula since 1850. Geophys. Res. Lett., 35, L01706. https://doi.org/10.1029/2007GL032529.CrossRefGoogle Scholar
Thomas, Z. A., et al. (2018). Evidence for increased expression of the Amundsen Sea Low over the South Atlantic during the late Holocene. Clim. Past, 14, 1727–1738. https://doi.org/10.5194/cp-14-1727-2018.CrossRefGoogle Scholar
Thompson, L. G., et al. (2009). Glacier loss on Kilimanjaro continues unabated. Proc. Natl. Acad. Sci. USA, 106, 19770–19775.CrossRefGoogle ScholarPubMed
Thompson, L. G., et al. (2013). Annually resolved ice core records of Tropical climate variability over the past 1800 Years. Science, 340, 945–950.CrossRefGoogle ScholarPubMed
Thost, D. E. and Truffer, M. (2008). Glacier recession on Heard Island, southern Indian Ocean. Arct., Antarct. Alp. Res., 40(1), 199–214.CrossRefGoogle Scholar
Tielidze, L.G., et al. (2025). Glacier inventories reveal an acceleration of Heard Island glacier loss over recent decades. EGUsphere [preprint], https://doi.org/10.5194/egusphere-2024-3811CrossRefGoogle Scholar
Timbal, B. and Drosdowsky, W. (2012). The relationship between the decline of Southern Australian rainfall and the strengthening of the subtropical ridge. Int. J. Climatol., 33(4), 1021–1034. https://doi.org/10.1002/joc.3492.Google Scholar
Turner, J. and Pendlebury, S. (eds.), (2004). The International Antarctic Weather Forecasting Handbook. British Antarctic Survey, Cambridge, 663pp.Google Scholar
Turney, C., et al. (2007). Redating the advance of the New Zealand Franz Josef Glacier during the Last Termination: Evidence for asynchronous climate change. Quat. Sci. Rev., 26, 3037–3042. https://doi.org/10.1016/j.quascirev.2007.09.014.CrossRefGoogle Scholar
Tyson, P. D., et al. (1997). Circulation changes and teleconnections between glacial advances on the west coast of New Zealand and extended spells of drought years in South Africa. Int. J. Climatol., 17, 1499–1512.3.0.CO;2-O>CrossRefGoogle Scholar
Ummenhofer, C. C., et al. (2009). Effect of tropical and subtropical Indian Ocean dipoles in driving precipitation around Indian Ocean rim countries. Unpublished abstract.Google Scholar
van der Bilt, W. G. M., et al. (2017). Late Holocene glacier reconstruction reveals retreat behind present limits and two-stage Little Ice Age on subantarctic South Georgia. J. Quat. Sci., 32, 888–901.CrossRefGoogle Scholar
Van der Bilt, W. G. M., et al. (2022). Stable Southern Hemisphere westerly winds throughout the Holocene until intensification in the last two millennia. Commun. Earth Environ., 3, 186. https://doi.org/10.1038/s43247-022-00512-8.CrossRefGoogle Scholar
Vaughan, D. G. and Doake, C. S. M. (1996). Recent atmospheric warming and retreat of ice shelves on the Antarctic Peninsula. Nature, 379, 328–33.CrossRefGoogle Scholar
Veettil, B. K., et al. (2014). Combined influence of PDO and ENSO on northern Andean glaciers: A case study on the Cotopaxi ice-covered volcano, Ecuador. Clim. Dyn., 43(12), 3439–3448.CrossRefGoogle Scholar
Veettil, B. K., et al. (2016). Influence of ENSO and PDO on mountain glaciers in the outer tropics: Case studies in Bolivia. Theor. Appl. Climatol., 125, 757e768. http://dx.doi.org/10.1007/s00704-015-1545-4.CrossRefGoogle Scholar
Veettil, B. K., et al. (2017). Glacier monitoring and glacier-climate interactions on the tropical Andes: A review. J. Sth. Am. Earth Sci., 77, 1–29. https://doi.org/10.1016/j.jsames.2017.04.009.Google Scholar
Verfaillie, D., et al. (2015). Recent glacier decline in the Kerguelen Islands (49°S, 69°E) derived from modeling, field observations, and satellite data. J. Geophys. Res. Earth Surf., 120, 637–654. https://doi.org/10.1002/2014JF003329.CrossRefGoogle Scholar
Viale, M., et al. (2019). Contrasting climates at both sides of the Andes in Argentina and Chile. Front. Environ. Sci., 7, 69. https://doi.org/10.3389/fenvs.2019.00069.CrossRefGoogle Scholar
Villalba, R., et al. (2003). Large-scale temperature changes across the southern Andes: 20th-century variations in the context of the past 400 years. Clim. Change, 59, 177–232. https://doi.org/10.1023/A:1024452701153.CrossRefGoogle Scholar
Vimeux, F., et al. (2009). Climate variability during the last 1000 years inferred from Andean ice cores: A review of methodology and recent results. Palaeogeogr. Palaeocl., 281, 229–241. https://doi.org/10.1016/j.palaeo.2008.03.054, 2009.CrossRefGoogle Scholar
Vuille, M. (1999). Atmospheric circulation over the Bolivian altiplano during dry and wet periods and extreme phases of the Southern Oscillation. Int. J. Climatol. 19, 1579–1600.3.0.CO;2-N>CrossRefGoogle Scholar
Vuille, M. and Ammann, C. (1997). Regional snowfall patterns in the high arid Andes. Clim. Chang., 36(3–4), 413–423.CrossRefGoogle Scholar
Vuille, M. and Bradley, R. S. (2000). Mean annual temperature trends and their vertical structure in the tropical Andes. Geophys. Res. Lett., 27, 3885–3888.CrossRefGoogle Scholar
Vuille, M., et al. (2000a). Interannual climate variability in the Central Andes and its relation to tropical Pacific and Atlantic forcing. J. Geophys. Res., 105, 12447–12460.Google Scholar
Vuille, M., et al. (2000b). Climatic variability in the Andes of Ecuador and its relation to tropical Pacific and Atlantic sea surface temperature anomalies. J. Clim., 13, 2520–2535.2.0.CO;2>CrossRefGoogle Scholar
Vuille, M. and Keimig, F. (2004). Interannual variability of summertime convective cloudiness and precipitation in the Central Andes derived from ISCCP-B3 data. J. Clim., 17, 3334–3348.2.0.CO;2>CrossRefGoogle Scholar
Vuille, M., et al. (2008). Climate change and tropical Andean glaciers – Past, present and future. Earth Sci. Rev., 89, 79–96. http://dx.doi.org/10.1016/j.earscirev.2008.04.002.CrossRefGoogle Scholar
Vuille, M., et al. (2018). Rapid decline of snow and ice in the tropical Andes – Impacts, uncertainties and challenges ahead. Earth-Sci. Rev., 176, 195–213. http://dx.doi.org/10.1016/j.earscirev.2017.09.019.CrossRefGoogle Scholar
Waliser, D. and Guan, B. (2017). Extreme winds and precipitation during landfall of atmospheric rivers. Nat. Geosci., 10, 179–183.CrossRefGoogle Scholar
Warren, C. R., et al. (1997). Greatest Holocene advance of Glaciar Pio XI, Chilean Patagonia: Possible causes. Ann. Glaciol., 21, 311–316.Google Scholar
Webster, P. J., et al. (1999). Coupled ocean-atmosphere dynamics in the Indian Ocean during 1997–98. Nature, 401, 356–360.CrossRefGoogle ScholarPubMed
Weidemann, S. S., et al. (2018). A 17-year record of meteorological observations across the Gran Campo Nevado Ice Cap in Southern Patagonia, Chile, related to synoptic weather types and climate modes. Front. Earth Sci., 6, 53. https://doi.org/10.3389/feart.2018.00053.CrossRefGoogle Scholar
Weidemann, S. S., et al. (2020). Recent climatic mass balance of the Schiaparelli Glacier at the Monte Sarmiento Massif and reconstruction of Little Ice Age climate by simulating steady-state glacier conditions. Geosciences, 10, 272. https://doi.org/10.3390/geosciences10070272.CrossRefGoogle Scholar
Zazulie, N., et al. (2017). Spatio-temporal mapping of glacier fluctuations in the subtropical Central Andes: Case studies of Alto Del Plomo and Volcan Maipo. Remote Sens. Appl.: Soc. Environ., 8, 140–147.Google Scholar
Zech, R., et al. (2008). Timing of the late Quaternary glaciation in the Andes from ~15 to 40° S. J. Quat, Sci., 23, 635–647. ISSN 0267-8179.CrossRefGoogle Scholar
Zwier, M., et al. (2021). Pollen evidence of variations in Holocene climate and Southern Hemisphere westerly wind strength on sub-Antarctic South Georgia. The Holocene, 1–12. https://doi.org/10.1177/09596836211060495.Google Scholar
Aggarwal, P. K., et al. (2016). Proportions of convective and stratiform precipitation revealed in water isotope ratios. Nat. Geosci., 9, 624–628. https://doi.org/10.1038/NGEO2739.CrossRefGoogle Scholar
Agosta, C., et al. (2019). Estimation of the Antarctic surface mass balance using the regional climate model MAR (1979–2015) and identification of dominant processes. The Cryosphere, 13, 281–296. https://doi.org/10.5194/tc-13-281-2019.CrossRefGoogle Scholar
Akers, P. D., et al. (2022). Sunlight-driven nitrate loss records Antarctic surface mass balance. Nat. Commun., 13, 4274. https://doi.org/10.1038/s41467-022-31855-7.CrossRefGoogle ScholarPubMed
Alexander, B., et al. (2003). East Antarctic ice core sulfur isotope measurements over a complete glacial-interglacial cycle. J. Geophys. Res. Atmos., 108(24), 4786. https://doi.org/10.1029/2003JD003513.CrossRefGoogle Scholar
Alley, R. B. (2000). The Two-Mile Time Machine: Ice Cores, Abrupt Climate Change, and Our Future – Updated Edition (REV-Revised). Princeton University Press. https://doi.org/10.2307/j.ctt6wq1f8.Google Scholar
Amory, C. (2020). Drifting-snow statistics from multiple-year autonomous measurements in Adélie Land, East Antarctica. Cryosphere, 14, 1713–1725. https://doi.org/10.5194/tc-14-1713-2020.CrossRefGoogle Scholar
Arthur, J. F., et al. (2020). Recent understanding of Antarctic supraglacial lakes using satellite remote sensing. Prog. Phys. Geogr., 44(6), 837–869. https://doi.org/10.1177/0309133320916114.CrossRefGoogle Scholar
Bagnold, R. A. (1941). The Physics of Blown Sand and Desert Dunes. Methuen, London, 265pp.Google Scholar
Ball, F. K. (1956). The theory of strong katabatic winds. Aust. J. Phys., 9, 373–386.CrossRefGoogle Scholar
Ball, F. K. (1957). The katabatic winds of Adélie Land and King George V Land. Tellus, 9, 201–208.Google Scholar
Ball, F. K. (1960). Winds on the ice slopes of Antarctica. Antarctic Meteorology: Proceedings of the Symposium, Melbourne 1959, Pergamon, 9–16.Google Scholar
Barkan, E. and Luz, B. (2007). Diffusivity fractionations of H216O/H217O and H216O/H218O in air and their implications for isotope hydrology. Rapid Commun. Mass. Spectrom., 21, 2999–3005. https://doi.org/10.1002/rcm.3180.CrossRefGoogle Scholar
Bauska, T. K., et al. (2016). Carbon isotopes characterize rapid changes in atmospheric carbon dioxide during the last deglaciation. Proc. Natl. Acad. Sci. USA, 113(13), 3465–3470. www.pnas.org/cgi/doi/10.1073/pnas.1513868113.CrossRefGoogle ScholarPubMed
Bell, R. E., et al. (2018). Antarctic surface hydrology and impacts on ice-sheet mass balance. Nat. Clim. Change, 8, 1044–1052.CrossRefGoogle Scholar
Bender, M. L., et al. (2008). The contemporary degassing rate of 40Ar from the solid Earth. Proc. Natl. Acad. Sci. USA, 105(24), 8232–8237.CrossRefGoogle ScholarPubMed
Berner, W., et al. (1980). Information on the CO2 cycle from ice core studies. Radiocarbon, 22, 227–235.CrossRefGoogle Scholar
Bintanja, R. (1999). On the glaciological, meteorological and climatological significance of Antarctic blue ice areas. Rev. Geophys., 37(3), 337–359.CrossRefGoogle Scholar
Bromwich, D. H. (1988). Snowfall in high southern latitudes. Rev. Geophys., 26(1), 149–168. https://doi.org/10.1029/RG026i001p00149.CrossRefGoogle Scholar
Bromwich, D. H. (1989). Satellite analyses of Antarctic katabatic wind behavior. Bull. Am. Meteorol. Soc., 70, 738–749.2.0.CO;2>CrossRefGoogle Scholar
Bromwich, D. H., et al. (1992). Satellite observations of katabatic-wind propagation for great distances across the Ross Ice Shelf. Mon. Weather. Rev., 120, 1940–1949.2.0.CO;2>CrossRefGoogle Scholar
Bromwich, D. H., et al. (1993). Hemispheric atmospheric variations and oceanographic impacts associated with katabatic surges across the Ross Ice Shelf, Antarctica. J. Geophys. Res., 98(D7), 13045–13062.Google Scholar
Bromwich, D. H., et al. (2013). Central West Antarctica among the most rapidly warming regions on Earth. Nat. Geosci., 6, 139–145.CrossRefGoogle Scholar
Budd, W. F. (1966a). The drifting of nonuniform snow particles. In Rubin, M. J. (ed.), Studies in Antarctic Meteorology, pp. 59–70. https://doi.org/10.1029/AR009p0059, 1966.Google Scholar
Budd, W. F. (1966b). Glaciological studies in the region of Wilkes, eastern Antarctica, 1961. Australian National Antarctic Research Expeditions, ANARE Scientific Reports, Series A (IV) Glaciology, Publication No. 88. Antarctic Division, Melbourne. 152pp.Google Scholar
Budd, W. F. (1969). The Dynamics of Ice Masses. Australian National Antarctic Research Expeditions, ANARE Scientific Reports, Series A (IV) Glaciology, Publication No. 108. Antarctic Division, Melbourne. 216pp.Google Scholar
Buizert, C., et al. (2014). Radiometric 81Kr dating identifies 120,000-year-old ice at Taylor Glacier, Antarctica. Proc. Natl. Acad. Sci. USA, 111(19), 6876–6881. www.pnas.org/cgi/doi/10.1073/pnas.1320329111.CrossRefGoogle Scholar
Burgener, L., et al. (2013). An observed negative trend in West Antarctic accumulation rates from 1975 to 2010: Evidence from new observed and simulated records. J. Geophys. Res., Atmos., 118, 4205–4216.CrossRefGoogle Scholar
Casado, M., et al. (2016). Continuous measurements of isotopic composition of water vapour on the East Antarctic Plateau. Atmos. Chem. Phys., 16, 8521–8538. https://doi.org/10.5194/acp-16-8521-2016.CrossRefGoogle Scholar
Ciais, P. and Jouzel, J. (1994). Deuterium and oxygen 18 in precipitation: Isotopic model, including mixed cloud processes. J. Geophys. Res., 99, 16793–16803.Google Scholar
Ciais, P., et al. (1995). The origin of present-day Antarctic precipitation from surface snow deuterium excess data. J. Geophys. Res., 100, 18917–18927.Google Scholar
Corr, D., et al. (2022). An inventory of supraglacial lakes and channels across the West Antarctic Ice Sheet. Earth Syst. Sci. Data, 14, 209–228. https://doi.org/10.5194/essd-14-209-2022.CrossRefGoogle Scholar
Craig, H. (1961). Isotopic variations in meteoric waters. Science, 133, 1702–1703.CrossRefGoogle ScholarPubMed
Craig, H., et al. (1963). Isotopic exchange effects in the evaporation of water. Journ. Geophys. Res., 68, 5079–5087.Google Scholar
Dansgaard, W. (1964). Stable isotopes in precipitation. Tellus, 16, 436–468. https://doi.org/10.3402/tellusa.v16i4.8993.Google Scholar
De Conto, R. M. and Pollard, D. (2016). Contribution of Antarctica to past and future sea-level rise. Nature, 531, 591–597. https://doi.org/10.1038/nature17145.Google ScholarPubMed
Delmas, R. J., et al. (1980). Polar ice evidence that atmospheric CO2 20,000 yr BP was 50% of present. Nature, 284, 155–157.CrossRefGoogle Scholar
Delmotte, M., et al. (2000). A seasonal deuterium excess signal at Law Dome, coastal Eastern Antarctica: A southern ocean signature. J. Geophys. Res., 105, 7187–7197.Google Scholar
Diaz, M. A., et al. (2020). Stable isotopes of nitrate, sulfate, and carbonate in soils from the Transantarctic Mountains, Antarctica: A record of atmospheric deposition and chemical weathering. Front. Earth Sci., 8, 341. https://doi.org/10.3389/feart.2020.00341.CrossRefGoogle Scholar
Domensino, B. (2010). Using synoptic climate pattern typing to resolve snow accumulation and glaciochemical variability at Mill Island, East Antarctica. Department of Environment and Geography, Macquarie University, Honours thesis.Google Scholar
Ekaykin, A. (2003). Meteorological regime of central Antarctic and its role in the formation of isotope composition of snow thickness. Ph.D. thesis, University Grenoble 1-Joseph Fourier, Grenoble, France.Google Scholar
Fahnestock, M. A., et al. (2000). Snow megadune fields on the East Antarctic Plateau: Extreme atmosphere-ice interaction. Geophys. Res. Lett., 27(20), 3719–3722.CrossRefGoogle Scholar
Favier, V., et al. (2017). Antarctica-regional climate and surface mass budget. Curr. Clim. Change Rep., 3, 303–315. https://doi.org/10.1007/s40641-017-0072-z.CrossRefGoogle Scholar
Fernandoy, F., et al. (2018). New insights into the use of stable water isotopes at the northern Antarctic Peninsula as a tool for regional climate studies. Cryosphere, 12, 1069–1090. https://doi.org/10.5194/tc-12-1069-2018.CrossRefGoogle Scholar
Filhol, S. and Sturm, M. (2015). Snow bedforms: A review, new data, and a formation model. J. Geophys. Res., Earth Surf., 120, 1645–1669. https://doi.org/10.1002/2015JF003529.CrossRefGoogle Scholar
Fleming, W. L. S. (1940). Relic glacial forms on the western seaboard of Graham Land. Geogr. J., 96(2), 93–100.CrossRefGoogle Scholar
Franz, P. and Röckmann, T. (2005). High-precision isotope measurements of H16O, H17O, H18O, and the ∆17O-anomaly of water vapor in the southern lowermost stratosphere. Atmos. Chem. Phys., 5, 2949–2959. https://doi.org/10.5194/acp-5-2949-2005.CrossRefGoogle Scholar
Fraser, A. D., et al. (2023). Antarctic landfast sea ice: A review of its physics, biogeochemistry and ecology. Rev. Geophys., 61, e2022RG000770. https://doi.org/10.1029/2022RG000770.CrossRefGoogle Scholar
Frezzotti, M., et al. (2002a). Snow megadunes in Antarctica, sedimentary structure and genesis. J. Geophys. Res., 107(D18), 4344. https://doi.org/10.1029/2001JD000673.Google Scholar
Frezzotti, M., et al. (2002b). Snow dunes and glazed surface in Antarctica: New field and remote sensing data. Ann. Glaciol., 34, 81–88, 2002.CrossRefGoogle Scholar
Fudge, T. J., et al. (2016).Variable relationship between accumulation and temperature in West Antarctica for the past 31,000 years. Geophys. Res. Lett., 43, 3795–3803.CrossRefGoogle Scholar
Fujiwara, K. and Yasoicho, E. (1971). Preliminary report of glaciological studies. In Murayama, M. (ed.), Report of the Japanese Traverse, Syowa-South Pole 1968–1969. JARE Scientific Reports, Special Issue, No. 2, Polar Research Centre, National Science Museum, Tokyo.Google Scholar
Fyke, J., et al. (2017). Basin-scale heterogeneity in Antarctic precipitation and its impact on surface mass variability. Cryosphere, 11, 2595–2609. https://doi.org/10.5194/tc-11-2595-2017.CrossRefGoogle Scholar
Gallée, H., et al. (2005). Simulation of the net snow accumulation along the Wilkes Land transect, Antarctica, with a regional climate model. Ann. Glaciol., 41, 17–22.CrossRefGoogle Scholar
Gallée, H., et al. (2013). Transport of snow by the wind: A comparison between observations in Adélie Land, Antarctica, and simulations made with the regional climate model MAR. Boundary-Layer Meteorol., 146(1), 133–147. https://doi.org/10.1007/s10546-012-9764-z.CrossRefGoogle Scholar
Genthon, C., et al. (2003). Interannual Antarctic tropospheric circulation and precipitation variability. Clim. Dyn., 21, 289–307. https://doi.org/10.1007/s00382-003-0329-1.CrossRefGoogle Scholar
Giovinetto, M. B. and Bentley, C. R. (1985). Surface balance in ice drainage systems of Antarctica. Antarct. J. U.S., 20(4), 6–13.Google Scholar
Goodwin, I. D. (1988). Ice sheet topography and surface characteristics in eastern Wilkes Land, East Antarctica. ANARE Research Notes, 64, 100pp.Google Scholar
Goodwin, I. D. (1990). Snow accumulation and surface topography in the katabatic zone of eastern Wilkes Land, Antarctica. Antarct. Sci., 2(3), 235–242.CrossRefGoogle Scholar
Goodwin, I. D. (1991). Snow-accumulation variability from seasonal surface observations and firn-core stratigraphy, eastern Wilkes Land, Antarctica. J. Glaciol., 37(127), 383–387.CrossRefGoogle Scholar
Goodwin, I. D. (1996). A mid to late Holocene readvance of the Law Dome ice margin, Budd Coast, East Antarctica. Antarct. Sci., 8(4), 395–406.CrossRefGoogle Scholar
Goodwin, I., et al. (2003). Snow accumulation variability in Wilkes Land, East Antarctica, and the relationship to atmospheric ridging in the 130°–170°E region since 1930. J. Geophys. Res., 108(D21), 4673. https://doi.org/10.1029/2002JD002995.Google Scholar
Goodwin, I. D., et al. (2004). Mid latitude winter climate variability in the south Indian and south-west Pacific regions since 1300 AD. Clim. Dyn., 22, 783–794. https://doi.org/10.1007/S00382-004-0403-3.CrossRefGoogle Scholar
Goodwin, I. D., et al. (2014). A reconstruction of extratropical Indo-Pacific sea-level pressure patterns during the Medieval Climate Anomaly. Clim. Dyn., 43(5–6), 1197–1219. https://doi.org/10.1007/s00382-013-1899-1.CrossRefGoogle Scholar
Gore, D. B. and Leishman, M. R. (2020). Tafoni show postglacial and modern wind azimuths that are similar at Bunger Hills. Antarct. Sci., 32. https://doi.org/10.1017/S095410201900035X.CrossRefGoogle Scholar
Goursaud, S., et al. (2018). Water stable isotope spatio-temporal variability in Antarctica in 1960–2013: Observations and simulations from the ECHAM5-wiso atmospheric general circulation model. Clim. Past, 14, 923–946.CrossRefGoogle Scholar
Goyal, R., et al. (2021). Historical and projected changes in the Southern Hemisphere surface westerlies. Geophys. Res. Lett., 48, e2020GL090849. https://doi.org/10.1029/2020GL090849.CrossRefGoogle Scholar
Heinemann, G., et al. (2019). A satellite-based climatology of wind-induced surface temperature anomalies for the Antarctic. Remote Sens., 11, 1539. https://doi.org/10.3390/rs11131539.CrossRefGoogle Scholar
Herbert, W. W. (1963). In Amundsen’s tracks on the Axel Heiberg Glacier. Geogr. J., 129(4), 397–410.CrossRefGoogle Scholar
Higgins, J. A., et al. (2015). Atmospheric composition 1 million years ago from blue ice in the Allan Hills, Antarctica. Proc. Natl. Acad. Sci. USA, 112(22), 6887–6891. https://doi.org/10.1073/pnas.1420232112.CrossRefGoogle ScholarPubMed
Hoefs, J. (2021). Stable Isotope Geochemistry, 9th ed. Springer, Berlin/Heidelberg, p. 503.CrossRefGoogle Scholar
Hoffmann, K., et al. (2018). Stable water isotopes and accumulation rates in the Union Glacier region, West Antarctica over the last 35 years. The Cryosphere Discuss. https://doi.org/10.5194/tc-2018-161.CrossRefGoogle Scholar
Hollin, J. T. and Cameron, R. L. (1961). I.G.Y. glaciological work at Wilkes Station, Antarctica. J. Glaciol., 3(29), 833–843.CrossRefGoogle Scholar
Holtedahl, O. (1929). On the geology and physiography of some Antarctic and sub-Antarctic islands. Scientific Results of the Norwegian Antarctic Expeditions 1927-1928 and 1928–1929. Det Norske Videnskaps-Akademi I Oslo, 172 pp.Google Scholar
Johnsen, S. J., et al. (2000). Diffusion of stable isotopes in polar firn and ice. the isotope effect in firn diffusion. In Hondoh, T. (ed.), Physics of Ice Core Records. Hokkaido University Press, Sapporo, pp. 121–140.Google Scholar
Jones, M. E., et al. (2019). Sixty years of widespread warming in the southern middle and high latitudes (1957–2016). J. Clim., 32(20), 6875–6898. https://doi.org/10.1175/JCLI-D-18-0565.1.CrossRefGoogle Scholar
Jouzel, J., et al. (1982). Deuterium excess in an East Antarctic ice core suggests higher relative humidity at the oceanic surface during the last glacial maximum. Nature, 299, 688–691.CrossRefGoogle Scholar
Jouzel, J. and Merlivat, L. (1984). Deuterium and oxygen18 in precipitation: Modeling of the isotopic effects during snow formation. J. Geophys. Res., 89, 11749–11759. https://doi.org/10.1029/jd089id07p11749.Google Scholar
Jouzel, J., et al. (1987). Vostok ice core, A continuous isotope temperature record over the last climatic cycle (160,000 years). Nature, 329, 403–408.CrossRefGoogle Scholar
Jouzel, J. J., et al. (2007). Orbital and millennial Antarctic climate variability over the past 800,000 Years. Science, 317, 793. https://doi.org/10.1126/science.1141038.CrossRefGoogle ScholarPubMed
Jouzel, J., et al. (2013). Water isotopes as tools to document oceanic sources of precipitation. Water Resour. Res., 49, 7469–7486. https://doi.org/10.1002/2013WR013508.CrossRefGoogle Scholar
Kehrl, L., et al. (2018). Evaluating the duration and continuity of potential climate records from the Allan Hills Blue Ice Area, East Antarctica. Geophys. Res. Lett., 45, 4096–4104. https://doi.org/10.1029/2018GL077511.CrossRefGoogle Scholar
Kidson, E. (1928). British Antarctic Expedition 1907–1909. Meteorology. Reports of the Scientific Investigations. H.J. Green. Melbourne, Australia. 188pp.Google Scholar
Kidson, E. (1946). Australian Antarctic Expedition 1911–14, Meteorology. Discussion of observations at Adelie Land, Queen Mary Land and Macquarie Island. Scientific Reports Series B, Vol. VI. 121pp.Google Scholar
King, A. C. F., et al. (2019). Organic compounds in a sub-Antarctic ice core: A potential suite of sea ice markers. Geophys. Res. Lett., 46, 9930–9939. https://doi.org/10.1029/2019GL084249.CrossRefGoogle Scholar
King, J. C. and Turner, J. (1997). Antarctic Meteorology and Climatology. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Kingslake, J., et al. (2017). Widespread movement of meltwater onto and across Antarctic ice shelves. Nature, 544, 349–352. https://doi.org/10.1038/nature22049.CrossRefGoogle ScholarPubMed
Korotkikh, E. V., et al. (2011). The last interglacial as represented in the glaciochemical record from Mount Moulton Blue Ice Area, West Antarctica. Quat. Sci. Rev., 30, 1940–1947. https://doi.org/10.1016/j.quascirev.2011.04.020.CrossRefGoogle Scholar
Kurita, N. (2011). Origin of Arctic water vapor during the ice-growth season. Geophys. Res. Lett., 38, L02709. https://doi.org/10.1029/2010GL046064.CrossRefGoogle Scholar
Kuttel, M., et al. (2012). Seasonal climate information preserved in West Antarctic ice core water isotopes: Relationships to temperature, large-scale circulation, and sea ice. Clim. Dyn., 39, 1841–1857. https://doi.org/10.1007/s00382-012-1460-7.CrossRefGoogle Scholar
Landais, A., et al. (2008). Record of δ18O and 17O-excess in ice from Vostok Antarctica during the last 150,000 years. Geophys. Res. Lett., 35, L02709. https://doi.org/10.1029/2007GL032096.Google Scholar
Landais, A., et al. (2021). Interglacial Antarctic–Southern Ocean climate decoupling due to moisture source area shifts. Nat. Geosc., 14, 918–923.CrossRefGoogle Scholar
Lenaerts, J. T. M., et al. (2012). Impact of model resolution on simulated wind, drifting snow and surface mass balance in Terre Adélie, East Antarctica. J. Glaciol., 48(211), 821–829. https://doi.org/10.3189/2012JoG12J020.Google Scholar
Lenaerts, J. T. M., et al. (2019). Observing and modeling ice sheet surface mass balance. Rev. Geophys., 57, 376–420. https://doi.org/10.1029/2018RG000622.CrossRefGoogle ScholarPubMed
Leonard, K. C., et al. (2011). Drifting snow threshold measurements near McMurdo station, Antarctica: A sensor comparison study. Cold Reg. Sci. Technol., 70, 71–80. https://doi.org/10.1016/j.coldregions.2011.08.001.Google Scholar
Lin, Y., et al. (2013). Oxygen isotope anomaly observed in water vapor from Alert, Canada and the implication for the stratosphere. Proc. Natl. Acad. Sci. USA, 110, 15608–15613.Google ScholarPubMed
Lisiecki, L. E. and Raymo, M. E. (2005). A Pliocene–Pleistocene stack of 57 globally distributed benthic δ18O records. Paleoceanogr., 20, PA1003.Google Scholar
Lister, H. (1960). Glaciology. 1. Solid precipitation and drift snow. Trans-Antarctic Expedition 1955–1958. Scientific Reports 5., 51pp.Google Scholar
Loewe, F. (1956). Etudies de glaciology en Terre Adelie 1951–1952. Expedition Polaires Francaise, IX, Paris.Google Scholar
Lorius, C. (1969). A physical and chemical study of the coastal ice sampled from a core drilling in Antarctica. IASC Pub., 79, 141–148.Google Scholar
Lorius, C. and Merlivat, L. (1977). Distribution of mean surface stable isotope values in East Antarctica: Observed changes with depth in a coastal area. Isotopes and Impurities in Snow and Ice: Proc. of the Grenoble Symp. August/September 1975, Vienna, Austria, IAHS, 125–137.Google Scholar
Madigan, C. T. (1929). Australian Antarctic Expedition 1911–14. Tabulated and reduced records of the Cape Denison Station, Adelie Land. Scient. Reports. Ser. B 1V. NSW Govt Printer. 287pp.Google Scholar
Markle, B. R., et al. (2012). Synoptic variability in the Ross Sea region, Antarctica, as seen from back-trajectory modeling and ice core analysis. J. Geophys. Res., 117, D02113. https://doi.org/10.1029/2011JD016437.Google Scholar
Marshall, G. J. (2003). Trends in the Southern Annular Mode from observations and reanalyses. J. Clim., 16(24), 4134–4143.2.0.CO;2>CrossRefGoogle Scholar
Marshall, G. J., et al. (2017). The signature of Southern Hemisphere atmospheric circulation patterns in Antarctic precipitation. Geophys. Res. Lett., 44, 580–611. https://doi.org/10.1002/2017GL075998.CrossRefGoogle ScholarPubMed
Massom, R. A., et al. (2004). Precipitation over the interior East Antarctic Ice Sheet related to midlatitude blocking-high activity. J. Clim., 17, 1914–1928.2.0.CO;2>CrossRefGoogle Scholar
Masson-Delmotte, V., et al. (2008). A review of Antarctic surface snow isotopic composition: Observations, atmospheric circulation and isotopic modeling. J. Clim., 21, 3359–3387.CrossRefGoogle Scholar
Mather, K. B. and Miller, G. S. (1966). Wind drainage off the high plateau of eastern Antarctica. Nature, 209, 281–284.CrossRefGoogle Scholar
Mather, K. B. and Miller, G. S. (1967). The problem of the katabatic wind on the coast of Terre Adélie. Polar Rec., 13, 425–432.CrossRefGoogle Scholar
Mayewski, P. A. (2001). An Ice Core Time Machine, An American Museum of Natural History Book, The New Press, New York.Google Scholar
Mayewski, P. A. and Goodwin, I. D. (1997). International Trans-Antarctic Scientific Expedition (ITASE), ‘200 Years of Past Antarctic Climate and Environmental Change,’ Science and Implementation Plan, 1996, PAGES Workshop Rep., Ser. 97–1, 48pp.Google Scholar
McMorrow, A., et al. (2002). Features of meteorological events preserved in a high-resolution Law Dome (East Antarctica) snow pit. Ann. Glaciol., 35, 463–470.CrossRefGoogle Scholar
Merlivat, L. and Jouzel, J. (1979). Global climatic interpretation of the deuterium- oxygen18 relationship for precipitation. J. Geophys. Res., 84(C8), 5029–5033. https://doi.org/10.1029/jc084ic08p05029.Google Scholar
Monaghan, A. J., et al. (2006). Insignificant change in Antarctic snowfall since the International Geophysical Year. Science, 313(5788), 827–831. https://doi.org/10.1126/science.1128243.CrossRefGoogle ScholarPubMed
Moore, J. C., et al. (2006). Interpreting ancient ice in a shallow ice core from the South Yamato (Antarctica) blue ice area using flow modeling and compositional matching to deep ice cores. J. Geophys. Res., 111, D16302. https://doi.org/10.1029/2005JD006343.Google Scholar
Morgan, V. I. (1982). Antarctic Ice Sheet surface oxygen isotope values. J. Glaciol., 18(99), 315–323.Google Scholar
Morgan, V. I. (1985). An oxygen isotope–climate record from the Law Dome. Climatic Change, 7, 415–426. 0165-0009/85.15.CrossRefGoogle Scholar
Mottram, R., et al. (2021). What is the surface mass balance of Antarctica? An intercomparison of regional climate model estimates. Cryosphere, 15, 3751–3784. https://doi.org/10.5194/tc-15-3751-2021.CrossRefGoogle Scholar
Noone, D. and Simmonds, I. (2002). Annular variations in moisture transport mechanisms and the abundance of δ18O in Antarctic snow. J. Geophys. Res., 107(D24), 4742. https://doi.org/10.1029/2002JD002262.Google Scholar
Oliva, M., et al. (2017). Recent regional climate cooling on the Antarctic Peninsula and associated impacts on the cryosphere. Sci. Total Environ., 580, 210–223. https://doi.org/10.1016/J.SCITOTENV.2016.12.030.CrossRefGoogle ScholarPubMed
Palm, S. P., et al. (2011). Satellite remote sensing of blowing snow properties over Antarctica. J. Geophys. Res., 116, D16123. https://doi.org/10.1029/2011JD015828.Google Scholar
Palm, S. P., et al. (2018). Toward a satellite-derived climatology of blowing snow over Antarctica. J. Geophys. Res.: Atmos., 123, 10301–10313. https://doi.org/10.1029/2018JD028632.CrossRefGoogle Scholar
Parish, T. R. (1988). Surface winds over the Antarctic continent: A review. Rev. Geophys., 26, 169–180.CrossRefGoogle Scholar
Parish, T. R. and Bromwich, D. H. (1987). The surface windfield over the Antarctic ice sheets. Nature, 328, 51–54.Google Scholar
Parish, T. R. and Cassano, J. J. (2003). The role of katabatic winds on the Antarctic surface wind regime. Mon. Weather Rev., 131, 317–333.2.0.CO;2>CrossRefGoogle Scholar
Parish, T. R. and Bromwich, D. H. (2007). Reexamination of the near-surface airflow over the Antarctic Continent and implications on atmospheric circulations at high southern latitudes. Mon. Weather Rev., 135, 1961–1973.CrossRefGoogle Scholar
Peel, D. A. (1992). Ice core evidence from the Antarctic Peninsula region. In Bradley, R. S. and Jones, P. D. (eds.), Climate since A.D. 1500. Routledge, London and New York, pp. 549–571.Google Scholar
Petit, J. R., et al. (1991). Deuterium excess in recent Antarctic snow. J. Geophys. Res., 96, 5113–5122.Google Scholar
Petit, J. R., et al. (1999). Climate and atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica. Nature, 399(6735), 429–436.CrossRefGoogle Scholar
Pol, K., et al. (2014). Climate variability features of the last interglacial in the East Antarctic EPICA Dome C ice core. Geophys. Res. Lett., 41, 4004–4012. https://doi.org/10.1002/2014GL059561.CrossRefGoogle Scholar
Pruett, L. E., et al. (2004). Sulfur isotopic measurements from a West Antarctic ice core: Implications for sulfate source and transport. Ann. Glaciol., 39, 161–168.CrossRefGoogle Scholar
Raisbeck, G. M., et al. (2007). Direct north-south synchronization of abrupt climate change record in ice cores using Beryllium 10. Clim. Past, 3, 541–547. https://doi.org/10.5194/cp-3-541-2007.CrossRefGoogle Scholar
Risi, C., et al. (2010). Understanding the 17O excess glacial-interglacial variations in Vostok precipitation. J. Geophys. Res., 115, D10112. https://doi.org/10.1029/2008JD011535.Google Scholar
Roberts, J. L., et al. (2013). Borehole temperatures reveal a changed energy budget at Mill Island, East Antarctica, over recent decades. Cryosphere, 7, 263–273. https://doi.org/10.5194/tc-7-263-2013.CrossRefGoogle Scholar
Robin, G. de Q. (1977). Ice cores and climatic change. Phil. Trans. Roy. Soc., B, 280, 143–168.Google Scholar
Robin, G. de Q. (ed). (1983). The Climatic Record in Polar Ice Sheets. Cambridge University Press, Cambridge, 212pp.Google Scholar
Rymill, J. R. (1938). British Graham Land Expedition, 1934–37. Geogr. J., 91(4), 297–312.Google Scholar
Scambos, T., et al. (2000). The link between climate warming and break-up of ice shelves in the Antarctic Peninsula. J. Glaciol., 46, 516–530.CrossRefGoogle Scholar
Scarchilli, C., et al. (2011). Snow precipitation at four ice core sites in East Antarctica: Provenance, seasonality and blocking factors. Clim. Dyn., 37, 2107–2125. https://doi.org/10.1007/s00382-010-0946-4.CrossRefGoogle Scholar
Schlosser, E., et al. (2004). The influence of precipitation origin on the δ18O-T relationship at Neumayer Station, Ekströmisen, Antarctica. Ann. Glaciol., 39, 41–48.CrossRefGoogle Scholar
Schlosser, E., et al. (2008). The precipitation regime of Dronning Maud Land, Antarctica, derived from AMPS (Antarctic Mesoscale Prediction System) Archive Data. J. Geophys. Res., 113, D24108. https://doi.org/10.1029/2008JD009968.Google Scholar
Schlosser, E., et al. (2017). The influence of the synoptic regime on stable water isotopes in precipitation at Dome C, East Antarctica. Cryosphere, 11, 2345–2361. https://doi.org/10.5194/tc-11-2345-2017.CrossRefGoogle Scholar
Schoenemann, S. W., et al. (2014). Triple water-isotopologue record from WAIS Divide, Antarctica: Controls on glacial-interglacial changes in 17O excess of precipitation. J. Geophys. Res. Atmos., 119, 2014JD021770. https://doi.org/10.1002/2014JD021770.CrossRefGoogle Scholar
Servettaz, A. P. M., et al. (2020). Snowfall and water stable isotope variability in East Antarctica controlled by warm synoptic events. J. Geophys. Res. Atmos., 125, e2020JD032863. https://doi.org/10.1029/2020JD032863.CrossRefGoogle Scholar
Shackleton, N. J. (2000). The 100,000-year ice-age cycle identified and found to lag temperature, carbon dioxide, and orbital eccentricity. Science, 289, 1897–1902.CrossRefGoogle ScholarPubMed
Shindell, D. T. and Schmidt, G. A. (2004). Southern Hemisphere climate response to ozone changes and greenhouse gas increases. Geophys. Res. Lett., 31, L18209. https://doi.org/10.1029/2004GL020724.CrossRefGoogle Scholar
Sime, L. C., et al. (2009). Evidence for warmer interglacials in East Antarctic ice cores. Nature, 462, 342–345.CrossRefGoogle ScholarPubMed
Simpson, G. C. (1919). British Antarctic Expedition 1910–1913, Meteorology, volume 1, Discussion. Harrison and Sons, London. 326pp.Google Scholar
Simpson, G. C. (1919). British Antarctic Expedition 1910–1913, Meteorology, volume 2, Weather Maps and Pressure Curves. Harrison and Sons, London.Google Scholar
Sinclair, V. A. and Dacre, H. F. (2019). Which extratropical cyclones contribute most to the transport of moisture in the Southern Hemisphere? J. Geophys. Res. Atmos., 124(5), 2525–2545. https://doi.org/10.1029/2018JD028766.CrossRefGoogle Scholar
Sinisalo, A. and Moore, J. C. (2010). Antarctic blue ice areas – Towards extracting palaeoclimate information. Antarct. Sci., 22(02), 99–115. https://doi.org/10.1017/S0954102009990691.CrossRefGoogle Scholar
Sodemann, H. and Stohl, A. (2009). Asymmetries in the moisture origin of Antarctic precipitation. Geophys. Res. Lett., 36, L22803. https://doi.org/10.1029/2009GL040242.CrossRefGoogle Scholar
Spaulding, N. E., et al. (2012). Ice motion and mass balance at the Allan Hills blue-ice area, Antarctica, with implications for paleoclimate reconstructions. J. Glaciol., 58(208), 399–406. https://doi.org/10.3189/2012JoG11J176.CrossRefGoogle Scholar
Spaulding, N. E., et al. (2013). Climate archives from 90 to 250 ka in horizontal and vertical ice cores from the Allan Hills Blue Ice Area, Antarctica. Quat. Res., 80(03), 562–574. https://doi.org/10.1016/j.yqres.2013.07.004.CrossRefGoogle Scholar
Steig, E. J., et al. (2014). Calibrated high-precision 17O-excess measurements using cavity ring-down spectroscopy with laser-current-tuned cavity resonance. Atmos. Meas. Tech., 7, 2421–2435. https://doi.org/10.5194/amt-7-2421-2014.CrossRefGoogle Scholar
Steig, E. J., et al. (2021). Continuous-Flow Analysis of δ17O, δ18O, and δD of H2O on an ice core from the South Pole. Front. Earth Sci., 9, 640292. https://doi.org/10.3389/feart.2021.640292.CrossRefGoogle Scholar
Stenni, B., et al. (2016). Three-year monitoring of stable isotopes of precipitation at Concordia Station, East Antarctica. Cryosphere, 10, 2415–2428. https://doi.org/10.5194/tc-10-2415-2016.CrossRefGoogle Scholar
Stokes, C. R., et al. (2019). Widespread distribution of supraglacial lakes around the margin of the east Antarctic ice Sheet. Sci. Rep., 9, 13823. https://doi.org/10.1038/s41598-019-50343-5.CrossRefGoogle ScholarPubMed
Swithinbank, C. (1988). Antarctica. U.S. Geol. Surv. Prof. Pap., 1386-B.Google Scholar
Tetzner, D., et al. (2021). A refined method to analyze insoluble particulate matter in ice cores, and its application to diatom sampling in the Antarctic Peninsula. Front. Earth Sci., 9 617043. https://doi.org/10.3389/feart.2021.617043.CrossRefGoogle Scholar
Thompson, D. W. J. and Barnes, E. A. (2014). Periodic variability in the large-scale Southern Hemisphere atmospheric circulation. Science, 343, 641–645.CrossRefGoogle ScholarPubMed
Thompson, D. W. J. and Woodworth, J. D. (2014). Barotropic and baroclinic annular variability in the Southern Hemisphere. J. Atmos. Sci., 71, 1480–1493. https://doi.org/10.1175/JAS-D-13-0185.1.CrossRefGoogle Scholar
Toggweiler, J. R. (1999). Variation of atmospheric CO2 by ventilation of the ocean’s deepest water. Paleoceanogr., 14, 571–588.CrossRefGoogle Scholar
Toggweiler, J. R. and Russell, J. (2008). Ocean circulation in a warming climate. Nature, 451, 286–288. https://doi.org/10.1038/nature06590.CrossRefGoogle Scholar
Touzeau, A., et al. (2016). Acquisition of isotopic composition for surface snow in East Antarctica and the links to climatic parameters. Cryosphere, 10, 837–852. https://doi.org/10.5194/tc-10-837-2016.Google Scholar
Turner, J., et al. (2005). A positive trend in western Antarctic Peninsula precipitation over the last 50 years reflecting regional and Antarctic-wide atmospheric circulation changes. Ann. Glaciol., 41, 85–91. https://doi.org/10.3189/172756405781813177.CrossRefGoogle Scholar
Turner, J., et al. (2019). The dominant role of extreme precipitation events in Antarctic snowfall variability. Geophys. Res. Lett., 46(6), 3502–3511. https://doi.org/10.1029/2018GL081517.CrossRefGoogle Scholar
Turney, C., et al. (2013). Late Pleistocene and early Holocene Change in the Weddell Sea: A new highly-resolved climate record from Patriot Hills, Ellsworth Mountains, West Antarctica. J. Quatern. Sci., 28(7), 697–704. https://doi.org/10.1002/jqs.2668.CrossRefGoogle Scholar
Uemura, R., et al. (2004). An observation-based method for reconstructing ocean surface changes using a 340,000-year deuterium excess record from the Dome Fuji ice core, Antarctica. Geophys. Res. Lett., 31, L13216. https://doi.org/10.1029/2004GL019954.CrossRefGoogle Scholar
Uemura, R., et al. (2008). Evidence of deuterium excess in water vapor as an indicator of ocean surface conditions. J. Geophys. Res., 113, D19114. https://doi.org/10.1029/2008JD010209.Google Scholar
Uemura, R., et al. (2012). Ranges of moisture-source temperature estimated from Antarctic ice cores stable isotope records over glacial–interglacial cycles. Clim. Past, 8, 1109–1125. https://doi.org/10.5194/cp-8-1109-2012.CrossRefGoogle Scholar
Uemura, R., et al. (2016). Sulfur isotopic composition of surface snow along a latitudinal transect in East Antarctica. Geophys. Res. Lett., 43, 5878–5885. https://doi.org/10.1002/2016GL069482.CrossRefGoogle Scholar
Uotila, P., et al. (2011). Relationships between Antarctic cyclones and surface conditions as derived from high-resolution numerical weather prediction data. J. Geophys. Res., 116, D7. https://doi.org/10.1029/2010JD015358.Google Scholar
Vance, T. R., et al. (2013). A millennial proxy record of ENSO and Eastern Australian rainfall from the Law Dome Ice Core, East Antarctica. J. Clim., 26, 710–725.CrossRefGoogle Scholar
Vance, T. R., et al. (2015). Interdecadal Pacific variability and eastern Australian megadroughts over the last millennium. Geophys. Res. Lett., 42, 129–137.CrossRefGoogle Scholar
Vance, T. R., et al. (2016). Optimal site selection for a high-resolution ice core record in East Antarctica. Clim. Past, 12, 595–610. www.clim-past.net/12/595/2016/.CrossRefGoogle Scholar
Van de Berg, W. J., et al. (2006). Reassessment of the Antarctic surface mass balance using calibrated output of a regional atmospheric climate model. J. Geophys. Res., 111, D11, D11104.10.1029/ 2005JD006495.Google Scholar
van den Broeke, M. R., et al. (2003). Factors controlling the near-surface wind field in Antarctica. Mon. Weather Rev., 131(4), 733–743.2.0.CO;2>CrossRefGoogle Scholar
van den Broeke, M. and van Lipzig, N. (2004). Changes in Antarctic temperature, wind and precipitation in response to the Antarctic Oscillation. Ann. Glaciol., 39, 19–27. https://doi.org/10.3189/172756404781814654.CrossRefGoogle Scholar
van Ommen, T. D. and Morgan, V. (1997). Calibrating the ice core paleothermometer using seasonality. J. Geophys. Res., 102(D8), 9351–9257.Google Scholar
van Ommen, T. D. and Morgan, V. (2010). Snowfall increase in coastal East Antarctica linked with southwest Western Australian drought. Nat. Geosci., 3, 267.CrossRefGoogle Scholar
Vaughan, D. G., et al. (1999). Reassessment of surface mass balance in Antarctica. J. Climate, 12, 933–946.2.0.CO;2>CrossRefGoogle Scholar
Vignon, É., et al. (2021). Present and future of rainfall in Antarctica. Geophys. Res. Lett., 48, e2020GL092281. https://doi.org/10.1029/2020GL092281.Google Scholar
Vimeux, F., et al. (1999). Glacial-interglacial changes in the ocean surface conditions in the Southern Hemisphere. Nature, 398, 410–413.CrossRefGoogle Scholar
Vimeux, F., et al. (2001a). Holocene hydrological cycle changes in the Southern Hemisphere documented in East Antarctic deuterium excess records. Clim. Dyn., 17, 503–513. https://doi.org/10.1007/pl00007928.CrossRefGoogle Scholar
Vimeux, F., et al. (2001b). A 420,000 year deuterium-excess record from East Antarctica: Information on past changes in the origin of precipitation at Vostok. J. Geophys. Res., 106, 31863–31873.Google Scholar
Vimeux, F., et al. (2002). New insights into Southern Hemisphere temperature changes from Vostok ice cores using deuterium excess correction. Earth Planet. Sci. Lett., 203, 829–843.CrossRefGoogle Scholar
Wang, Y., et al. (2009). A new spatial distribution map of δ18O in Antarctic surface snow, Geophys. Res. Lett., 36, L06501. https://doi.org/10.1029/2008GL036939.CrossRefGoogle Scholar
Wendler, G., et al. (1997). On the extraordinary katabatic winds of Adelie Land. J. Geophys. Res., 102(D4), 4463–4474.Google Scholar
Werner, M., et al. (2018). Reconciling glacial Antarctic water stable isotopes with ice sheet topography and the isotopic paleothermometer. Nat. Commun., 9, 3537. https://doi.org/10.1038/s41467-018-05430-y.CrossRefGoogle ScholarPubMed
Winkler, R., et al. (2012). Deglaciation records of 17O-excess in East Antarctica: Reliable reconstruction of oceanic normalized relative humidity from coastal sites. Clim. Past, 8, 1–16. https://doi.org/10.5194/cp-8-1-2012.CrossRefGoogle Scholar
Winkler, R., et al. (2013). Interannual variation of water isotopologues at Vostok indicates a contribution from stratospheric water vapor. Proc. Natl. Acad. Sci. USA, 110, 17674–17679.CrossRefGoogle ScholarPubMed
Winter, K., et al. (2016). Assessing the continuity of the blue ice climate record at Patriot Hills, Horseshoe Valley, West Antarctica. Geophys. Res. Lett., 43, 2019–2026. https://doi.org/10.1002/2015GL066476.Google Scholar
Winton, V. H. L., et al. (2020). Deposition, recycling, and archival of nitrate stable isotopes between the air–snow interface: Comparison between Dronning Maud Land and Dome C, Antarctica. Atmos. Chem. Phys., 20, 5861–5885. https://doi.org/10.5194/acp-20-5861-2020.CrossRefGoogle Scholar
Wolff, E. W. (1995). Nitrate in polar ice. In Delmas, R. J. (ed), Ice Core Studies of Global Biogeochemical Cycles, Springer Berlin, Heidelberg, pp. 195–224.Google Scholar
Wolff, E. W. (2013). Ice sheets and nitrogen. Phil. Trans. R. Soc. B, 368, 20130127. http://dx.doi.org/10.1098/rstb.2013.0127.CrossRefGoogle ScholarPubMed
Wolff, E. W., et al. (2003). An ice core indicator of Antarctic sea ice production? Geophys. Res. Lett., 30(22), 2158. https://doi.org/10.1029/2003GL018454.CrossRefGoogle Scholar
Wolff, E. W., et al. (2006). Southern Ocean sea-ice extent, productivity and iron flux over the past eight glacial cycles. Nature, 440, 491–496. https://doi.org/10.1038/nature04614.CrossRefGoogle ScholarPubMed
Wolff, E. W., et al. (2008). The interpretation of spikes and trends in concentration of nitrate in polar ice cores, based on evidence from snow and atmospheric measurements. Atmos. Chem. Phys., 8, 5627–5634. https://doi.org/10.5194/acp-8-5627-2008.CrossRefGoogle Scholar
Wolff, E. W., et al. (2009). Changes in environment over the last 800,000 years from chemical analysis of the EPICA Dome C ice core. Quat. Sci. Rev., https://doi.org/10.1016/j.quascirev.2009.06.013.Google Scholar
Wright, C. S. and Priestley, R. E. (1922). British (Terra Nova) Antarctic Expedition 1910–1913. Glaciology. Harrison and Sons, London, 582pp.Google Scholar
Young, N. W. and Gibson, J. A. E. (2013). A century of change in the Shackleton and West ice shelves, East Antarctica. Unpublished manuscript, Australian Antarctic Division.Google Scholar
Zahn, A., et al. (2006). Modelling the budget of middle atmospheric water vapour isotopes, Atmos. Chem. Phys., 6, 2073–2090. https://doi.org/10.5194/acp-6-2073-2006.CrossRefGoogle Scholar
Zwally, H. J., et al. (1998). Areal distribution of the oxygen-isotope ratio in Antarctica: Comparison of results based on field and remotely sensed data. Ann. Glaciol., 27, 583–590.CrossRefGoogle Scholar
Aarons, S. M., et al. (2016). The impact of glacier retreat from the Ross Sea on local climate: Characterization of mineral dust in the Taylor Dome ice core, East Antarctica. Earth. Planet. Sci. Lett., 444, 34–44. https://doi.org/10.1016/j.epsl.2016.03.035.Google Scholar
Aarons, S. M., et al. (2017). Dust composition changes from Taylor Glacier (East Antarctica) during the last glacial-interglacial transition: A multi-proxy approach. Quat. Sci. Rev., 162, 60–71.CrossRefGoogle Scholar
Aarons, S. M., et al. (2019). Dust transport to the Taylor Glacier, Antarctica, during the last interglacial. Geophys. Res. Lett., 46, 2261–2270. https://doi.org/10.1029/2018GL081887.CrossRefGoogle Scholar
Abram, N. J., et al. (2007). Ice core records as sea ice proxies: An evaluation from the Weddell Sea region of Antarctica. J. Geophys. Res., 112, D15101. https://doi.org/10.1029/2006JD008139.Google Scholar
Abram, N. J., et al. (2008). The preservation of methanesulphonic acid in frozen ice-core samples. J. Glaciol., 54(187), 680–684.CrossRefGoogle Scholar
Abram, N. J., et al. (2010). Ice core evidence for a 20th century decline of sea ice in the Bellingshausen Sea, Antarctica. J. Geophys. Res., 115, D23101. https://doi.org/10.1029/2010JD014644.Google Scholar
Abram, N. J., et al. (2011). Environmental signals in a highly resolved ice core from James Ross Island, Antarctica. J. Geophys. Res., 116, D20116. https://doi.org/10.1029/2011JD016147.Google Scholar
Abram, N. J., et al. (2013). A review of sea ice proxy information from polar ice cores. Quat. Sci. Rev., 79, 168–183.CrossRefGoogle Scholar
Albani, S., et al. (2012). Interpreting last glacial to Holocene dust changes at Talos Dome (East Antarctica): Implications for atmospheric variations from regional to hemispheric scales. Climate of the Past, 8, 741–750.CrossRefGoogle Scholar
Alloway, B. V., et al. (2007). Towards a climate event stratigraphy for New Zealand over the past 30 000 years (NZ-INTIMATE project). J. Quat. Sci., 22(1), 9–35.CrossRefGoogle Scholar
Armand, L., et al. (2005). The biogeography of major diatom taxa in Southern Ocean sediments. 1. Sea ice related species. Palaeogeogr. Palaeoclimatol. Palaeoecol., 223, 93–126.CrossRefGoogle Scholar
Armand, L. K. and Leventer, A. (2010). Palaeo sea ice distribution and reconstruction derived from the geological record. In Thomas, D. N. and Dieckmann, G. S. (eds.), Sea Ice. Wiley-Blackwell, Oxford, pp. 469–530.Google Scholar
Baccolo, G., et al. (2018). Regionalization of the atmospheric dust cycle on the periphery of the East Ant- arctic ice sheet since the Last Glacial Maximum. Geochem. Geophys. Geosyst., 19, 3540–3554.CrossRefGoogle Scholar
Basile, I., et al. (1997). Patagonian origin of glacial dust deposited in East Antarctica (Vostok and Dome C) during glacial stages 2, 4 and 6. Earth. Planet. Sci. Lett., 146(3–4), 573–589. https://doi.org/10.1016/s0012-821x(96)00255-5.CrossRefGoogle Scholar
Beers, T. M., et al. (2006). 1150 year long ice core record of the Ross Sea Polynya, Antarctica. Unpublished document. Climate Change Institute, University of Maine. arXiv:2006.01093.Google Scholar
Benassai, S., et al. (2005). Sea-spray deposition in Antarctic coastal and plateau areas from ITASE traverses. Ann. Glaciol., 41(1), 32–40. https://doi.org/10.3189/172756405781813285.CrossRefGoogle Scholar
Bertler, N., et al. (2005). Snow chemistry across Antarctica. Ann. Glaciol., 41(1), 167–179. https://doi.org/10.3189/172756405781813320.CrossRefGoogle Scholar
Bertler, N. A. N., et al. (2018). The Ross Sea Dipole – Temperature, snow accumulation and sea ice variability in the Ross Sea region, Antarctica, over the past 2700 years. Clim. Past, 14, 193–214. https://doi.org/10.5194/cp-14-193-2018.Google Scholar
Blakowski, M. A., et al. (2016). Sr-Nd-Hf isotope characterization of dust source areas in Victoria Land and the McMurdo Sound sector of Antarctica. Quat. Sci. Rev., 141, 26–37. https://doi.org/10.1016/j.quascirev.2016.03.023.CrossRefGoogle Scholar
Bluth, G. J. S., et al. (1992). Global tracking of the SO2 clouds from the June 1991 Mount Pinatubo eruptions. Geophys. Res. Lett., 19, 151–154. https://doi.org/10.1029/91GL02792.CrossRefGoogle Scholar
Bory, A. E., et al. (2010). Multiple sources supply eolian mineral dust to the Atlantic sector of coastal Antarctica: Evidence from recent snow layers at the top of Berkner Island ice sheet. Earth. Planet. Sci. Lett., 291, 138–148. http://dx.doi.org/10.1016/j.epsl.2010.01.006.CrossRefGoogle Scholar
Bory, A. M., et al. (2003). Regional variability of ice core dust composition and provenance in Greenland. Geochem. Geophys. Geosyst., 4(12), 1107.CrossRefGoogle Scholar
Boutron, C. F. and Patterson, C. C. (1987). Relative levels of natural and anthropogenic lead in recent Antarctic snow. J. Geophys. Res., 92, 8454–8464. https://doi.org/10.1029/JD092iD07p08454.Google Scholar
Brévière, E., et al. (2015). Surface ocean-lower atmosphere study: Scientific synthesis and contribution to Earth system science. Anthropocene, 11. http://dx.doi.org/10.1016/j.ancene.2015.11.001.CrossRefGoogle Scholar
Brévière, E. and the SOLAS Scientific Steering Committee (eds.) (2016). SOLAS 2015–2025: Science Plan and Organisation. SOLAS International Project Office, GEOMAR Helmholtz Centre for Ocean Research Kiel, Kiel, 76pp.Google Scholar
Bromwich, D. H., et al. (1993). Hemispheric atmospheric variations and oceanographic impacts associated with katabatic surges across the Ross Ice Shelf, Antarctica. J. Geophys. Res., 98(D7), 13045–13062.Google Scholar
Brook, E. J. and Buizert, C. (2018). Antarctic and global climate history viewed from ice cores. Nature, 558, 200–208. https://doi.org/10.1038/s41586-018-0172-5.CrossRefGoogle ScholarPubMed
Brooks, S. D. and Thornton, D. C. O. (2018). Marine Aerosols and Clouds. Annu. Rev. Mar. Sci., 10, 289–313.CrossRefGoogle ScholarPubMed
Chambers, S. D., et al. (2018). Characterizing atmospheric transport pathways to Antarctica and the remote Southern Ocean using Radon-222. Front. Earth Sci., 6, 190. https://doi.org/10.3389/feart.2018.00190.CrossRefGoogle Scholar
Christ, A. J. (2015). Late Holocene glacial advance and ice shelf growth in Barilari Bay, Graham Land, west Antarctic Peninsula. GSA Bulletin, 1–19. https://doi.org/10.1130/B31035.1.CrossRefGoogle Scholar
Cole-Dai, J., et al. (1997). Annually resolved Southern Hemisphere volcanic history from two Antarctic ice cores. J. Geophys. Res., 102(D14), 16761–16771. https://doi.org/10.1029/97jd01394.Google Scholar
Cole-Dai, J. and Mosley-Thompson, E. (1999). The Pinatubo eruption in South Pole snow and its potential value to ice-core paleovolcanic records. Ann. Glaciol., 29, 99–105.CrossRefGoogle Scholar
Cole-Dai, J., et al. (2021). Comprehensive record of volcanic eruptions in the Holocene (11,000 years) from the WAIS Divide, Antarctica ice core. J. Geophys. Res. Atmos., 126, e2020JD032855. https://doi.org/10.1029/2020JD032855.CrossRefGoogle Scholar
Criscitiello, A. S., et al. (2013). Ice sheet record of recent sea-ice behavior and polynya variability in the Amundsen Sea, West Antarctica. J. Geophys. Res., Oceans, 118, 118–130. https://doi.org/10.1029/2012JC008077.CrossRefGoogle Scholar
Criscitiello, A. S., et al. (2014). Tropical Pacific influence on the source and transport of marine aerosols to West Antarctica. J. Clim., 27, 1343–1363.CrossRefGoogle Scholar
Crosta, X., et al. (2007). Holocene long-and short-term climate changes off Adélie Land, East Antarctica. Geochem. Geophys. Geosyst., 8, 1–15.CrossRefGoogle Scholar
Crosta, X., et al. (2008). Sea ice seasonality during the Holocene, Adélie Land, East Antarctica. Mar. Micropaleontol., 66, 222–232.CrossRefGoogle Scholar
Cunningham, J. and Waddington, E. D. (1993). Air flow and dry deposition of non sea-salt sulfate in polar firn: Paleoclimatic implications. Atmos. Environ., Part A, 27, 2943–2956.CrossRefGoogle Scholar
Curran, M. A. J., et al. (1998). Seasonal characteristics of the major ions in the high-accumulation Dome Summit South ice core, Law Dome, Antarctica. Ann. Glaciol., 27, 385–390.CrossRefGoogle Scholar
Curran, M. A. J. and Jones, G. B. (2000). Dimethylsulfide in the Southern Ocean: Seasonality and flux. J. Geophys. Res., 105, 20451–20459. https://doi.org/10.1029/2000JD900176.Google Scholar
Curran, M. A. J., et al. (2003). Ice core evidence for Antarctic sea ice decline since the 1950s. Science, 302, 1203–1206. https://doi.org/10.1126/science.1087888.CrossRefGoogle ScholarPubMed
Dall’Osto, M., et al. (2017). Antarctic sea ice region as a source of biogenic organic nitrogen in aerosols. Sci. Rep., 7, 6047. https://doi.org/10.1038/s41598-017-06188-x.CrossRefGoogle ScholarPubMed
De Deckker, P. (2019). An evaluation of Australia as a major source of dust. Earth-Sci. Rev., 194, 536–567.CrossRefGoogle Scholar
De Deckker, P. (2020). Airborne dust traffic from Australia in modern and Late Quaternary times. Glob. Planet. Change, 184, 103056.CrossRefGoogle Scholar
De Deckker, P., et al. (2010). Lead isotopic evidence for an Australian source of aeolian dust to Antarctica at times over the last 170,000 years. Palaeogeogr. Palaeoclimatol. Palaeoecol., 285(3–4), 205–223.Google Scholar
Delmas, R., et al. (1992). 1000 years of explosive volcanism recorded at the South Pole. Tellus, 44, 335–350.Google Scholar
Delmotte, B., et al. (2007). Late Quaternary interglacials in East Antarctica from ice core dust records. Chapter 6. In Siroko, F. et al. (eds.), The Climate of the Past Interglacials. Developments in Quaternary Sciences, Volume 7, Elsevier, Amsterdam, pp. 53–73.Google Scholar
Delmonte, B., et al. (2008). Aeolian dust in East Antarctica (EPICA-Dome C and Vostok): Provenance during glacial ages over the last 800 kyr. Geophys. Res. Lett., 35(7), L07703.CrossRefGoogle Scholar
Delmonte, B., et al. (2010). Aeolian dust in the Talos Dome ice core (East Antarctica, Pacific/Ross Sea sector): Victoria Land versus remote sources over the last two climate cycles. J. Quat. Sci., 25(8), 1327–1337.CrossRefGoogle Scholar
Delmonte, B., et al. (2013). Modern and Holocene aeolian dust variability from Talos Dome (Northern Victoria Land) to the interior of the Antarctic ice sheet. Quat. Sci. Rev., 64, 76–89.CrossRefGoogle Scholar
Delmonte, B., et al. (2017). Causes of dust size variability in central East Antarctica (Dome B): Atmospheric transport from expanded South American sources during Marine Isotope Stage 2. Quat. Sci. Rev., 168, 55–68. http://dx.doi.org/10.1016/j.quascirev.2017.05.009.CrossRefGoogle Scholar
Delmotte, B., et al. (2020). Holocene dust in East Antarctica: Provenance and variability in time and space. Holocene, 30(4), 546–558.Google Scholar
DePaolo, D. J. (1981). A neodymium and strontium isotopic study of the Mesozoic calc-alkaline granitic batholiths of the Sierra Nevada and Peninsular Ranges, California. J. Geophys. Res., 86(B11), 10470–10488. https://doi.org/10.1029/JB086iB11p10470.Google Scholar
Dixon, D. A., et al. (2012). An ice core proxy for northerly air mass incursions (NAMI) into West Antarctica. Int. J. Climatol., 32, 1455–1465.CrossRefGoogle Scholar
Domack, E., et al. (2001). Chronology of the Palmer Deep site, Antarctic Peninsula: A Holocene palaeoenvironmental reference for the circum-Antarctic. Holocene, 11, 1–9. https://doi.org/10.1191/095968301673881493.CrossRefGoogle Scholar
Domack, E., et al. (2003). Marine sedimentary record of natural environmental variability and recent warming in the Antarctic Peninsula. Antarct. Res. Ser., 79, 205–222.CrossRefGoogle Scholar
Domack, E., et al. (2005). Stability of the Larsen B ice shelf on the Antarctic Peninsula during the Holocene epoch. Nature, 436(7051), 681–685. https://doi.org/10.1038/nature03908.CrossRefGoogle Scholar
Draxler, R. R. (1992). Hybrid Single-Particle Lagrangian Integrated Trajectories (HY-SPLIT): Version 3.0 – User’s guide and model description. Air Resources Laboratory Tech. Memo. ERL ARL-195, 84pp. [Available online at www.arl.noaa.gov/documents/reports/ARL%20TM-195.pdf.]Google Scholar
Draxler, R. R. and Hess, G. D. (1998). An overview of the HYSPLIT_4 modeling system for trajectories, dispersion, and deposition. Aust. Meteorol. Mag., 47, 295–308.Google Scholar
Draxler, R. R. and Rolph, G. D. (2010). HYSPLIT (HYbrid Single-Particle Lagrangian Integrated Trajectory) Model. Access via NOAA ARL READY Website (http://ready.arl.noaa.gov/HYSPLIT.php).Google Scholar
Du, Z., et al. (2018). Identification of multiple natural and anthropogenic sources of dust in snow from Zhongshan Station to Dome A, East Antarctica. J. Glaciol., 64, 855–865.CrossRefGoogle Scholar
Dunbar, N. W., et al. (2003). Tephra layers in the Siple Dome and Taylor Dome ice cores, Antarctica: Sources and correlations. J. Geophys. Res., 108(B8), 2374. https://doi.org/10.1029/2002JB002056.Google Scholar
Eglinton, T. I. and Eglinton, G. (2008). Molecular proxies for paleoclimatology. Earth. Planet. Sci. Lett., 275(1–2), 1–16.CrossRefGoogle Scholar
Felix, J. D. and Elliott, E. M. (2013). The agricultural history of human-nitrogen interactions as recorded in ice core δ15N-NO3. Geophys. Res. Lett., 40(8), 1642–1646.CrossRefGoogle Scholar
Fischer, H., et al. (2007). Reconstruction of millennial changes in dust emission, transport and regional sea ice coverage using the deep EPICA ice cores from the Atlantic and Indian Ocean sector of Antarctica. Earth. Planet. Sci. Lett., 260(1), 340–354.CrossRefGoogle Scholar
Foster, A. F. M., et al. (2006). Covariation of sea ice and methanesulphonic acid in Wilhelm II Land, East Antarctica. Ann. Glaciol., 44, 429–432.CrossRefGoogle Scholar
Gabric, A., et al. (2018). The nexus between sea ice and polar emissions of marine biogenic aerosols. Bull. Amer. Meteorol. Soc., 99, 61–81. https://doi.org/10.1175/BAMS-D-16-0254.1.CrossRefGoogle Scholar
Gabrielli, P., et al. (2010). A major glaciale interglacial change in aeolian dust composition as inferred from Rare Earth Elements in Antarctic ice. Quat. Sci. Rev., 29, 265–273.CrossRefGoogle Scholar
Gabrielli, P., et al. (2014). Deglaciated areas of Kilimanjaro as a source of volcanic trace elements deposited on the ice cap during the late Holocene. Quat. Sci. Rev., 93, 1–10. http://dx.doi.org/10.1016/j.quascirev.2014.03.007.CrossRefGoogle Scholar
Gaiero, D. M., et al. (2004). The signature of river and windborne materials exported from Patagonia to the southern latitudes: A view from REEs and implications for paleoclimatic interpretations. Earth. Planet. Sci. Lett., 219(3–4), 357–376.CrossRefGoogle Scholar
Gaiero, D. M., et al. (2007). A uniform isotopic and chemical signature of dust exported from Patagonia: Rock sources and occurrence in southern environments. Chem. Geol., 238(1–2), 107–120.CrossRefGoogle Scholar
Gaiero, D. M., et al. (2013). Ground/satellite observations and atmospheric modeling of dust storms originating in the high Puna-Altiplano deserts (South America): Implications for the interpretation of paleo-climatic archives. J. Geophys. Res. Atmos., 118(9), 3817–3831.CrossRefGoogle Scholar
Gao, C., et al. (2007). Atmospheric volcanic loading derived from bipolar ice cores: Accounting for the spatial distribution of volcanic deposition. J. Geophys. Res., 112, D09109. https://doi.org/10.1029/2006JD007461.Google Scholar
Gao, C., et al. (2008). Volcanic forcing of climate over the past 1500 years: An improved ice core-based index for climate models. J. Geophys. Res., 113, D2311. https://doi.org/10.1029/2008jd010239.Google Scholar
Gasso, S., et al. (2010). A combined observational and modeling approach to study modern dust transport from the Patagonia desert to East Antarctica. Atmos. Chem. Phys., 10, 8287–8303. https://doi.org/10.5194/acp-10-8287-2010.CrossRefGoogle Scholar
Gaudichet, A., et al. (1992). Comments on the origin of dust in East Antarctica for present and ice age conditions. J. Atmos. Chem., 14, 129–142. https://doi.org/10.1007/BF00115229.Google Scholar
Gautier, E., et al. (2019). 2600-years of stratospheric volcanism through sulfate isotopes. Nature Commun., 10, 466. https://doi.org/10.1038/s41467-019-08357-0.Google ScholarPubMed
Gili, S., et al. (2016). Provenance of dust to Antarctica: A lead isotopic perspective. Geophys. Res. Lett., 43, 2291–2298. https://doi.org/10.1002/2016GL068244.CrossRefGoogle Scholar
Gili, S., et al. (2017). Glacial/interglacial changes of Southern Hemisphere wind circulation from the geochemistry of South American dust. Earth Planet. Sci. Lett., 469, 98–109.CrossRefGoogle Scholar
Gili, S., et al. (2022). South African dust contribution to the high southern latitudes and East Antarctica during interglacial stages. Commun., Earth Environ., 3, 129. https://doi.org/10.1038/s43247-022-00464-z.CrossRefGoogle Scholar
Gingele, F., et al. (2007). Late Pleistocene and Holocene climate of SE Australia reconstructed from dust and river loads deposited offshore the River Murray Mouth. Earth Planet. Sci., Lett., 255(3), 257–272. https://doi.org/10.1016/j.epsl.2006.12.019.CrossRefGoogle Scholar
Goodwin, I., et al. (2003). Snow accumulation variability in Wilkes Land, East Antarctica, and the relationship to atmospheric ridging in the 130°–170°E region since 1930. J. Geophys. Res., 108(D21), 4673. https://doi.org/10.1029/2002JD002995.Google Scholar
Goodwin, I. D., et al. (2004). Mid latitude winter climate variability in the south Indian and south-west Pacific regions since 1300 AD. Clim. Dyn., 22, 783–794. https://doi.org/10.1007/S00382-004-0403-3.CrossRefGoogle Scholar
Goodwin, I. D., et al. (2014). A reconstruction of extratropical Indo-Pacific sea-level pressure patterns during the Medieval Climate Anomaly. Clim. Dyn., 43(5–6), 1197–1219. https://doi.org/10.1007/s00382-013-1899-1.CrossRefGoogle Scholar
Grousset, F., et al. (1992). Antarctic (Dome C) ice-core dust at 18 ky BP: Isotopic constraints on origins. Earth Planet. Sci. Lett., 111, 175–182.CrossRefGoogle Scholar
Hara, K., et al. (2004). Chemistry of sea-salt particles and inorganic halogen species in Antarctic regions: Compositional differences between coastal and inland stations. J. Geophys. Res. Atmos., 109, D20. https://doi.org/10.1029/2004JD004713.CrossRefGoogle Scholar
Hara, K., et al. (2018). Important contributions of sea-salt aerosols to atmospheric bromine cycle in the Antarctic coasts. Sci. Rep., 8, 13852. https://doi.org/10.1038/s41598-018-32287-4.CrossRefGoogle ScholarPubMed
Hara, K., et al. (2020). Atmospheric sea-salt and halogen cycles in the Antarctic. Environ. Sci.: Processes Impacts, 22, 2003.Google ScholarPubMed
Harder, S. L., et al. (1996). Filtering of air through snow as a mechanism for aerosol deposition to the Antarctic ice sheet. J. Geophys. Res., 101, 18729–18743.Google Scholar
Hastings, M. G., et al. (2005). Glacial/interglacial changes in the isotopes of nitrate from the Greenland Ice Sheet Project 2 (GISP2) ice core. Global Biogeochem. Cycles, 19, GB4024.CrossRefGoogle Scholar
Hastings, M. G., et al. (2009). Anthropogenic impacts on nitrogen isotopes of ice-core nitrate. Science, 324(5932), 1288–1288. https://doi.org/10.1126/science.1170510.CrossRefGoogle ScholarPubMed
Heintzenberg, J., et al. (2023). Spatio-temporal distributions of the natural non-sea-salt aerosol over the Southern Ocean and Coastal Antarctica and its potential source regions. Tellus B: Chem. Phys. Meteorol., 75(1), 47–64. https://doi.org/10.16993/tellusb.1869.CrossRefGoogle Scholar
Hesse, P. P. (1994). The record of continental dust from Australia in Tasman Sea sediments. Quat. Sci. Rev., 13, 257–272.CrossRefGoogle Scholar
Hobbs, W., et al. (2016), Century-scale perspectives on observed and simulated Southern Ocean sea ice trends from proxy reconstructions. J. Geophys. Res. Oceans, 121, 7804–7818. https://doi.org/10.1002/2016JC012111.CrossRefGoogle Scholar
Iizuka, Y., et al. (2012). The rates of sea salt sulfatization in the atmosphere and surface snow of inland Antarctica. J. Geophys. Res. Atmos., 117(D4). https://doi.org/10.1029/2011JD016378.CrossRefGoogle Scholar
Iizuka, Y., et al. (2016). Spatial distributions of soluble salts in surface snow of East Antarctica. Tellus B: Chem. Phys. Meteorol., 68(1), 29285. https://doi.org/10.3402/tellusb.v68.29285.CrossRefGoogle Scholar
Jiang, S., et al. (2019). Nitrate preservation in snow at Dome A, East Antarctica from ice core concentration and isotope records. Atmos. Environ., 213, 405–412.CrossRefGoogle Scholar
Jiayue, H. and Lyatt, J. (2017). Wintertime enhancements of sea salt aerosol in polar regions consistent with a sea ice source from blowing snow. Atmos. Chem. Phys., 17, 3699–3712.Google Scholar
Johnson, K. M. (2021). Sensitivity of Holocene East Antarctic productivity to subdecadal variability set by sea ice. Nat. Geosci., 14, 762768. https://doi.org/10.1038/s41561-021-00816-y.CrossRefGoogle Scholar
Kavan, J., et al. (2018). Aerosol concentrations in relationship to local atmospheric conditions on James Ross Island, Antarctica. Front. Earth Sci., 6, 207. https://doi.org/10.3389/feart.2018.00207.CrossRefGoogle Scholar
King, A. C. F., et al. (2019). Organic compounds in a sub-Antarctic ice core: A potential suite of sea ice markers. Geophys. Res. Lett., 46, 9930–9939. https://doi.org/10.1029/2019GL084249.CrossRefGoogle Scholar
King, A. C. F. and Tetzner, D. R. (2021). Exploring novel ice-core proxies for paleoclimate reconstruction in the sub-Antarctic. PAGES Mag., 291. https://doi.org/10.22498/pages.29.1.36.Google Scholar
Koffman, B. G., et al. (2014). Centennial-scale variability of the Southern Hemisphere westerly wind belt in the eastern Pacific over the past two millennia. Clim. Past, 10, 1125–1144.CrossRefGoogle Scholar
Koffman, B. G., et al. (2017). Rapid transport of ash and sulfate from the 2011 Puyehue-Cordón Caulle (Chile) eruption to West Antarctica. J. Geophys. Res. Atmos., 122, 8908–8920. https://doi.org/10.1002/2017JD026893.CrossRefGoogle Scholar
Koffman, B. G., et al. (2021a). Late Holocene dust provenance at Siple Dome, Antarctica. Quat. Sci. Rev., 274, 107271. https://doi.org/10.1016/j.quascirev.2021.107271.CrossRefGoogle Scholar
Koffman, B. G., et al. (2021b). New Zealand as a source of mineral dust to the atmosphere and ocean. Quat. Sci. Rev., 251, 106659.CrossRefGoogle Scholar
Kreutz, K. J. and Mayewski, P. A. (1999). Spatial variability of Antarctic surface snow glaciochemistry: Implications for palaeoatmospheric circulation reconstructions. Antarct. Sci., 11(1), 105–118.CrossRefGoogle Scholar
Kreutz, K. J., et al. (2000). Sea-level pressure variability in the Amundsen Sea region inferred from a West Antarctic glaciochemical record. J. Geophys. Res., 105(D3), 4047–4059.Google Scholar
Lamy, F., et al. (2019). Precession modulation of the South Pacific westerly wind belt over the past million years. Proc. Natl. Acad. Sci. USA, 116(47), 23455–23460. https://doi.org/10.1073/pnas.1905847116.CrossRefGoogle ScholarPubMed
Legrand, M. and Kirchner, S. (1988). Polar atmospheric circulation and chemistry of recent (1957–1983) South Polar precipitation. Geophys. Res. Lett., 15, 879–882.CrossRefGoogle Scholar
Legrand, M. and Kirchner, S. (1990). Origins and variations of nitrate in south polar precipitation. J. Geophys. Res., 95, 3493–3507.Google Scholar
Legrand, M., et al. (1992). Spatial and temporal variations of methanesulfonic-acid and non-seasalt sulfate in Antarctic ice. J. Atmos. Chem., 14(1–4), 245–260. http://dx.doi.org/10.1007/bf00115237.CrossRefGoogle Scholar
Legrand, M. and Mayewski, P. (1997). Glaciochemistry of polar ice cores: A review. Rev. Geophys., 35, 219–243.CrossRefGoogle Scholar
Legrand, M., et al. (1998). Ammonium in coastal Antarctic aerosol and snow: Role of polar ocean and penguin emissions. J. Geophys. Res., 103(D9), 11043–11056.Google Scholar
Legrand, M., et al. (2016). Year-round records of sea salt, gaseous, and particulate inorganic bromine in the atmospheric boundary layer at coastal (Dumont d’Urville) and central (Concordia) East Antarctic sites. J. Geophys. Res. Atmos., 121, 997–1023. https://doi.org/10.1002/2015JD024066.CrossRefGoogle Scholar
Leinen, M. and Sarnthein, M., (eds.). (1987). Paleoclimatology and Paleometeorology: Modern and Past Patterns of Global Atmospheric Transport. NATO ASI Series C. Mathematical and Physical Sciences, 282, Kluwer Academic Publishers, Dordrecht, The Netherlands, 909pp.Google Scholar
Leventer, A. (1992). Modern distribution of diatoms in sediments from the Georges V coast, Antarctica. Mar. Micropaleontol., 19, 315–332.CrossRefGoogle Scholar
Leventer, A., et al. (1996). Productivity cycles of 200–300 years in the Antarctic Peninsula region: Under- standing linkages among the sun, atmosphere, oceans, sea ice, and biota. Geol. Soc. Am. Bull., 108(12), 1626–1644.2.3.CO;2>CrossRefGoogle Scholar
Lewis, E. R. and Schwartz, S. E. (2004). Sea Salt Aerosol Production: Mechanisms, Methods, Measurements and Models: A Critical Review. Geophys. Monograph Ser., 152. AGU, Washington, DC, 413pp.CrossRefGoogle Scholar
Li, C., et al. (2020). Holocene dynamics of the southern westerly winds over the Indian Ocean inferred from a peat dust deposition record. Quat. Sci. Rev., 231, 106169. https://doi.org/10.1016/j.quascirev.2020.106169.CrossRefGoogle Scholar
Li, F., et al. (2008). Distribution, transport, and deposition of mineral dust in the Southern Ocean and Antarctica: Contribution of major sources. J. Geophys. Res., 113, D10207. https://doi.org/10.1029/2007JD009190.Google Scholar
Lohmann, U. (2009). Marine boundary layer clouds. In Le Quere, C. and Saltzman, E. S. (eds.), Surface Ocean–Lower Atmosphere Processes. Volume 187, Geophys. Monograph Ser., AGU, pp. 57–68. https://doi.org/10.1029/2008GM000761.Google Scholar
Lovejoy, S. and Lambert, F. (2019). Spiky fluctuations and scaling in high-resolution EPICA ice core dust fluxes. Clim. Past, 15, 1999–2017. https://doi.org/10.5194/cp-15-1999-2019.CrossRefGoogle Scholar
Lunt, D. J. and Valdes, P. J. (2001). Dust transport to Dome Antarctica at the Last Glacial Maximum and present day. Geophys. Res. Lett., 28(2), 295–298.CrossRefGoogle Scholar
Mahalinganathan, K., et al. (2012). Relation between surface topography and sea-salt snow chemistry from Princess Elizabeth Land, East Antarctica. Cryosphere Discuss., 5, 2967–2989. https://doi.org/10.5194/tcd-5-2967-2011.Google Scholar
Mahowald, N., et al. (1999). Dust sources and deposition during the last glacial maximum and current climate: A comparison of model results with paleodata from ice cores and marine sediments. J. Geophys. Res. Atmos., 104, 15895–15916.CrossRefGoogle Scholar
Mahowald, N. M., et al. (2006). Change in atmospheric mineral aerosols in response to climate: Last glacial period, preindustrial, modern, and doubled carbon dioxide climates. J. Geophys. Res. Atmos., 111(D10), D10202.Google Scholar
Marino, F., et al. (2008). Defining the geochemical composition of the EPICA Dome C ice core dust during the last glacial-interglacial cycle. Geochem. Geophys. Geosyst., 9(10), Q10018.CrossRefGoogle Scholar
Marino, F., et al. (2009). Coherent composition of glacial dust on opposite sides of the East Antarctic Plateau inferred from the deep EPICA ice cores. Geophys. Res. Lett., 36, L23703. https://doi.org/10.1029/2009GL040732.CrossRefGoogle Scholar
Markle, B. R., et al. (2018). Concomitant variability in high-latitude aerosols, water isotopes and the hydrologic cycle. Nat. Geosci., 11, 853–859. https://doi.org/10.1038/s41561-018-0210-9.CrossRefGoogle Scholar
Marx, S. K., et al. (2005). Estimates of Australian dust flux into New Zealand: Quantifying the eastern Australian dust plume pathway using trace element calibrated 210Pb as a monitor. Earth. Planet. Sci. Lett., 239, 336–351.CrossRefGoogle Scholar
Marx, S. K., et al. (2009). Long-range dust transport from eastern Australia: A proxy for Holocene aridity and ENSO-induced climate variability. Earth Planet. Sci. Lett., 282, 167–177.CrossRefGoogle Scholar
Marx, S. K., et al. (2018). Palaeo-dust records: A window to understanding past environments. Glob. Planet. Change, 165, 13–43. https://doi.org/10.1016/j.gloplacha.2018.03.001.CrossRefGoogle Scholar
Marx, S. K., et al. (2022). Dust emissions from Kati Thanda-Lake Eyre: A review, Transactions of the Royal Society of South Australia. Earth. Planet. Sci. Lett., 146(1), 168–206. https://doi.org/10.1080/03721426.2022.2054918.Google Scholar
Matsumoto, A. and Hinkley, T. K. (2001). Trace metal suites in Antarctic pre-industrial ice are consistent with emissions from quiescent degassing of volcanoes worldwide. Earth Planet. Sci. Lett., 186, 33–43.CrossRefGoogle Scholar
Mayewski, P. A. and Legrand, M. R. (1990). Recent increase in nitrate concentration of Antarctic snow. Nature, 346(6281), 258–260. https://doi.org/10.1038/346258a0.CrossRefGoogle Scholar
Mayewski, P. A., et al. (1990). An ice-core record of atmospheric response to anthropogenic sulphate and nitrate. Nature, 346(6284), 554–556. https://doi.org/10.1038/346554a0.CrossRefGoogle Scholar
Mayewski, P. A. and Goodwin, I. D. (1997). International Trans-Antarctic Scientific Expedition (ITASE), ‘200 Years of Past Antarctic Climate and Environmental Change,’ Science and Implementation Plan, 1996, PAGES Workshop Rep., Ser. 97-1, 48pp.Google Scholar
McConnell, J. and Edwards, R. (2008). Coal burning leave toxic heavy metal legacy in the Arctic. Proc. Natl. Acad. Sci. USA, 105(34), 12140–12144. https://doi.org/10.1073/pnas.0803564105.CrossRefGoogle ScholarPubMed
McGowan, H. A., et al. (2000). Identifying regional dust transport pathways: Application of kinematic trajectory modeling to a trans-Tasman case. Earth Surf. Process. Landf., 25, 633–647.3.0.CO;2-J>CrossRefGoogle Scholar
Mezgec, K., et al. (2017). Holocene sea ice variability driven by wind and polynya efficiency in the Ross Sea. Nat. Commun., 8, 1334. https://doi.org/10.1038/s41467-017-01455-x.CrossRefGoogle ScholarPubMed
Minikin, A., et al. (1998). Sulfur-containing species (sulfate and methanesulfonate) in coastal Antarctic aerosol and precipitation. J. Geophys. Res: Atmos., 103(D9), 10975–10990. https://doi.org/10.1029/98JD00249.Google Scholar
Mulvaney, R., et al. (1992). The ratio of MSA to non-sea-salt sulphate in Antarctic Peninsula ice cores. Tellus, Ser. B, 44, 295–303.CrossRefGoogle Scholar
Mulvaney, R., et al. (1993). The fractionation of sea salt and acids during transport across an Antarctic ice shelf. Tellus B: Chem. Phys. Meteorol., 45(2), 179–187. https://doi.org/10.3402/tellusb.v45i2.15591.CrossRefGoogle Scholar
Mulvaney, R. and Wolff, E. W. (1994). Spatial variability of the major chemistry of the Antarctic ice sheet. Ann. Glaciol., 20, 440–447.CrossRefGoogle Scholar
Mulvaney, R., et al. (1998). Post depositional change in snowpack nitrate from observation of year-round near-surface snow in coastal Antarctica. J. Geophys. Res., 103(D9), 11021–11031.Google Scholar
Murphy, B. F. and Simmonds, I. (1993). An analysis of strong wind events simulated in a GCM near Casey in the Antarctic. Mon. Weather Rev., 121, 522–534.2.0.CO;2>CrossRefGoogle Scholar
Narcisi, B., et al. (2016). A new Eemian record of Antarctic tephra layers retrieved from the Talos Dome ice core (Northern Victoria Land). Glob. Planet. Change., 137, 69–78. https://doi.org/10.1016/j.gloplacha.2015.12.016.CrossRefGoogle Scholar
Neff, P. D. and Bertler, N. A. N. (2015). Trajectory modeling of modern dust transport to the Southern Ocean and Antarctica. J. Geophys. Res.: Atmos., 120, 9303–9322.CrossRefGoogle Scholar
Nguyen, H. D., et al. (2019). Dust storm event of February 2019 in central and East Coast of Australia and evidence of long-range transport to New Zealand and Antarctica. Atmosphere, 10, 653. https://doi.org/10.3390/atmos10110653.CrossRefGoogle Scholar
Nightingale, P. D. (2009). Air-sea gas exchange. In Le Quere, C. and Saltzman, E. S. (eds.), Surface Ocean–Lower Atmosphere Processes. Volume 187, Geophys. Monograph Ser., AGU, pp. 69–97. https://doi.org/10.1029/2008GM000774.Google Scholar
Obbard, R. W., et al. (2009). Frost flower surface area and chemistry as a function of salinity and temperature. J. Geophys. Res.: Atmos., 114(D20). https://doi.org/10.1029/2009JD012481.Google Scholar
Olivier, S., et al. (2006). Temporal variations of mineral dust, biogenic tracers, and anthropogenic species during the past two centuries from Belukha ice-core, Siberian Altai. J. Geophys. Res., 111, D05309.Google Scholar
Osipov, E. Y., et al. (2019). Recent variability of atmospheric circulation patterns inferred from East Antarctica glaciochemical records. Geochem., 80, 125554. https://doi.org/10.1016/j.chemer.2019.125554.Google Scholar
Osman, M., et al. (2017). Methanesulfonic acid (MSA) migration in polar ice: Data synthesis and theory. Cryosphere, 11, 2439–2462. https://doi.org/10.5194/tc-11-2439-2017.CrossRefGoogle Scholar
Palmer, A. S., et al. (2001). High-precision dating of volcanic events (AD 1301–1995) using ice cores from Law Dome, Antarctica. J. Geophys. Res.: Atmos., 106(D22), 28089–28095. https://doi.org/10.1029/2001JD000330.Google Scholar
Pasteris, D., et al. (2014). Acidity decline in Antarctic ice cores during the Little Ice Age linked to changes in atmospheric nitrate and sea salt concentrations. J. Geophys. Res.: Atmos., 119, 5640–5652. https://doi.org/10.1002/2013JD020377.CrossRefGoogle Scholar
Petherick, L. M., et al. (2009). Reconstructing transport pathways for late Quaternary dust from eastern Australia. Geomorphol., 105, 67–79.CrossRefGoogle Scholar
Petit, J. R., et al. (1981). Ice age aerosol content from East Antarctic ice core samples and past wind strength. Nature, 293(5831), 391–394.CrossRefGoogle Scholar
Petit, J. R., et al. (1999). Climate and atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica. Nature, 399(6735), 429–436.CrossRefGoogle Scholar
Petit, J. R. and Delmonte, B. (2009). A model for large glacial – Interglacial climate-induced changes in dust and sea salt concentrations in deep ice cores (central Antarctica): Palaeoclimatic implications and prospects for refining ice core chronologies. Tellus B: Chem. Phys. Meteorol., 61(5), 768–790.CrossRefGoogle Scholar
Plummer, C. T., et al. (2012). An independently dated 2000-yr volcanic record from Law Dome, East Antarctica, including a new perspective on the dating of the c. 1450s eruption of Kuwae, Vanuatu. Clim. Past, 8, 1567–1590. https://doi.org/10.5194/cpd-8-1567-2012.CrossRefGoogle Scholar
Pomeroy, J. W. and Jones, H. G. (1996). Wind-blown snow: Sublimation, transport and changes to polar snow, in chemical exchange between the atmosphere and polar snow. In Wolff, E. W. and Bales, R. C. (eds.), NATO ASI Ser., Ser. 1, vol. 43. Springer-Verlag, New York, pp. 453–490.Google Scholar
Pook, M. and Cowled, L. (1999). On the detection of weather systems over the Antarctic interior in the FROST analyses. Weather Forecasting, 14, 920–929.2.0.CO;2>CrossRefGoogle Scholar
Potocki, M., et al. (2016). Recent increase in Antarctic Peninsula ice core uranium concentrations. Atmos. Environ., 140, 381–385.CrossRefGoogle Scholar
Predybaylo, E., et al. (2020). El Niño/Southern Oscillation response to low-latitude volcanic eruptions depends on ocean pre-conditions and eruption timing. Commun. Earth Environ., 1, 12. https://doi.org/10.1038/s43247-020-0013-y.CrossRefGoogle Scholar
Rankin, A. M., et al. (2000). Frost flowers as a source of fractionated sea salt aerosol in the polar regions. Geophys. Res. Lett., 27(21), 3469–3472. https://doi.org/10.1029/2000GL011771.CrossRefGoogle Scholar
Rankin, A. M., et al. (2002). Frost flowers: Implications for tropospheric chemistry and ice core interpretation. J. Geophys. Res., 107(D23), 4683. https://doi.org/10.1029/2002JD002492.Google Scholar
Rea, D. K. (1994). The paleoclimatic record provided by eolian deposition in the deep sea: The geologic history of wind. Rev. Geophys., 32(2), 159–195.CrossRefGoogle Scholar
Revel-Rolland, M., et al. (2006). Eastern Australia: A possible source of dust in East Antarctica interglacial ice. Earth Planet. Sci. Lett., 249(1–2), 1–13.Google Scholar
Rhodes, R. H., et al. (2009). Sea ice variability and primary productivity in the Ross Sea, Antarctica, from methylsulphonate snow record. Geophys. Res. Lett., 36, L10704. https://doi.org/10.1029/2009GL037311.CrossRefGoogle Scholar
Robock, A. (2000). Volcanic eruptions and climate. Rev. Geophys., 38(2), 191–219. https://doi.org/10.1029/1998rg000054.CrossRefGoogle Scholar
Robock, A. and Free, M. P. (1995). Ice cores as an index of global volcanism from 1850 to the present. J. Geophys. Res., 100(D6), 11549–11567. https://doi.org/10.1029/95jd00825.Google Scholar
Röthlisberger, R., et al. (2000). Technique for continuous high- resolution analysis of trace substances in firn and ice cores. Environ. Sci. Technol., 34(2), 338–342. https://doi.org/10.1021/es9907055.CrossRefGoogle Scholar
Röthlisberger, R., et al. (2002). Nitrate in Greenland and Antarctic ice cores: A detailed description of post – Depositional processes. Ann. Glaciol., 35, 209–216.CrossRefGoogle Scholar
Röthlisberger, R., et al. (2010). Potential and limitations of marine and ice core sea ice proxies: An example from the Indian Ocean sector. Quat. Sci. Rev., 29, 296–302. https://doi.org/10.1016/j.quascirev.2009.10.005.CrossRefGoogle Scholar
Rubino, M., et al. (2016). Ice-core records of biomass burning. Anthr. Rev., 3(2), 140–162. https://doi.org/10.1177/2053019615605117.Google Scholar
Ruth, U., et al. (2010). A major glacial-interglacial change in aeolian dust composition inferred from Rare Earth Elements in Antarctic ice. Quat. Sci. Rev., 720(29), 265–273.Google Scholar
Saigne, C. and Legrand, M. (1987). Measurements of methanesulphonic acid in Antarctic ice. Nature, 330(6145), 240–242.CrossRefGoogle Scholar
Scarchilli, C., et al. (2011). Snow precipitation at four ice core sites in East Antarctica: Provenance, seasonality and blocking factors. Clim. Dyn., 37, 2107–2125. https://doi.org/10.1007/s00382-010-0946-4.CrossRefGoogle Scholar
Shi, G., et al. (2015). Investigation of post-depositional processing of nitrate in East Antarctic snow: Isotopic constraints on photolytic loss, re-oxidation, and source inputs. Atmos. Chem. Phys., 15, 9435–9453.CrossRefGoogle Scholar
Sinclair, K. E., et al. (2013). Seasonality of airmass pathways to coastal Antarctica: Ramifications for interpreting high-resolution ice core records. J. Clim., 26, 2065–2076. https://doi.org/10.1175/JCLI-D-12-00167.1.CrossRefGoogle Scholar
Sinclair, K. E., et al. (2014). Twentieth century sea-ice trends in the Ross Sea from a high-resolution, coastal ice-core record. Geophys. Res. Lett., 41, 3510–3516. https://doi.org/10.1002/2014GL059821.CrossRefGoogle Scholar
Smik, L., et al. (2016). Distributions of highly branched isoprenoid alkenes and other algal lipids in surface waters from East Antarctica: Further insights for biomarker-based paleo sea-ice reconstruction. Org. Geochem., 95, 71–80.CrossRefGoogle Scholar
Sofen, E. D., et al. (2014). WAIS Divide ice core suggests sustained changes in the atmospheric formation pathways of sulfate and nitrate since the 19th century in the extratropical Southern Hemisphere. Atmos. Chem. Phys., 14(11), 5749–5769. https://doi.org/10.5194/acp-14-5749-2014.CrossRefGoogle Scholar
Spolaor, A., et al. (2014). Seasonality of halogen deposition in polar snow and ice. Atmos. Chem. Phys., 14, 9613–9622.CrossRefGoogle Scholar
Stein, A. F., et al. (2015). NOAA’s HYSPLIT atmospheric transport and dispersion modeling system. Bull. Amer. Meteor. Soc., 96, 2059–2077. http://dx.doi.org/10.1175/BAMS-D-14-00110.1.CrossRefGoogle Scholar
Stenchikov, G. L., et al. (1998). Radiative forcing from the 1991 Mount Pinatubo volcanic eruption. J. Geophys. Res., 103, 13837–13857. https://doi.org/10.1029/98JD00693.Google Scholar
Sudarchikova, N., et al. (2015). Modelling of mineral dust for interglacial and glacial climate conditions with a focus on Antarctica. Clim. Past, 11, 765–779. https://doi.org/10.5194/cp-11-765-2015.CrossRefGoogle Scholar
Svensson, A., et al. (2000). Characterization of late glacial continental dust in the Greenland Ice Core Project ice core. J. Geophys. Res. Atmos., 105, 4637–4656.CrossRefGoogle Scholar
Teinilä, K., et al. (2000). A study of size-segregated aerosol chemistry in the Antarctic atmosphere. J. Geophys. Res. Atmos., 105(D3), 3893–3904. https://doi.org/10.1029/1999JD901033.CrossRefGoogle Scholar
Thiede, J. (1979). Wind regimes over the late Quaternary southwest Pacific Ocean. Geology, 7, 259–262.2.0.CO;2>CrossRefGoogle Scholar
Thomas, E. R. and Abram, N. J. (2016). Ice core reconstruction of sea ice change in the Amundsen-Ross Seas since 1702 A.D. Geophys. Res. Lett., 43, 5309–5317. https://doi.org/10.1002/2016GL068130.CrossRefGoogle Scholar
Thomas, E. R., et al. (2019). Antarctic sea ice proxies from marine and ice core archives suitable for reconstructing sea ice over the past 2000 years. Geosciences, 9, 506. https://doi.org/10.3390/geosciences9120506.CrossRefGoogle Scholar
Toohey, M., et al. (2013). Volcanic sulfate deposition to Greenland and Antarctica: A modeling sensitivity study. J. Geophys. Res. Atmos., 118, 4788–4800. https://doi.org/10.1002/jgrd.50428.CrossRefGoogle Scholar
Traversi, R., et al. (2004). Spatial and temporal distribution of environmental markers from Coastal to Plateau areas in Antarctica by firn core chemical analysis. Int. J. Environ. Anal. Chem., 84(6–7), 457–470. https://doi.org/10.1080/03067310310001640393.CrossRefGoogle Scholar
Turner, J., et al. (2001). An extreme wind event at Casey station, Antarctica. J. Geophys. Res., 106(D7), 7291–7311.Google Scholar
Udisti, R., et al. (2012). Sea spray aerosol in central Antarctica. Present atmospheric behaviour and implications for paleoclimatic reconstructions. Atmos. Environ., 52, 109–120. https://doi.org/10.1016/j.atmosenv.2011.10.018.CrossRefGoogle Scholar
Uemura, R., et al. (2022). Soluble salts in deserts as a source of sulfate aerosols in an Antarctic ice core during the last glacial period. Earth Planet. Sci. Lett., 578, 117299.CrossRefGoogle Scholar
Vallelonga, P., et al. (2005). A 220 kyr record of Pb isotopes at Dome C Antarctica from analyses of the EPICA ice core. Geophys. Res. Lett., 32(1), L01706.CrossRefGoogle Scholar
Vallelonga, P., et al. (2010). Lead isotopic compositions in the EPICA Dome C ice core and Southern Hemisphere potential source areas. Quat. Sci. Rev., 29(1), 247–255.CrossRefGoogle Scholar
Vallelonga, P., et al. (2017). Sea-ice-related halogen enrichment at Law Dome, coastal East Antarctica. Clim. Past, 13, 171–184.CrossRefGoogle Scholar
Vallelonga, P., et al. (2021). Sea-ice reconstructions from bromine and iodine in ice cores. Quat. Sci. Rev., 269, 107133. https://doi.org/10.1016/j.quascirev.2021.107133.CrossRefGoogle Scholar
Venugopal, A. U., et al. (2022). Role of mineral dust in the nitrate preservation during the glacial period: Insights from the RICE ice core. Glob. Planet. Change, 209, 103745. https://doi.org/10.1016/j.gloplacha.2022.103745.CrossRefGoogle Scholar
Vickery, K. and Eckardt, F. (2013). Dust emission controls on the lower Kuiseb River valley, Central Namib. Aeolian Res., 10, 125–133. https://doi.org/10.1016/j.aeolia.2013.02.006.CrossRefGoogle Scholar
Vickery, K. J., et al. (2013). A sub-basin scale dust plume source frequency inventory for southern Africa, 2005–2008. Geophys. Res. Lett., 40(19), 5274–5279.CrossRefGoogle Scholar
Vogt, M. and Liss, P. S. (2009). Dimethylsulfide and climate. In Le Quere, C. and Saltzman, E. S., (eds.), Surface Ocean–Lower Atmosphere Processes. Geophysical Research Series 187. American Geophysical Union, pp. 197–232. https://doi.org/10.1029/2008GM000790.Google Scholar
von Holdt, J. R., et al. (2017). Landsat identifies aeolian dust emission dynamics at the landform scale. Remote Sens. of Environ., 198, 229–243.CrossRefGoogle Scholar
Wagenbach, D. (1996). Coastal Antarctica: Atmospheric chemical composition and atmospheric transport. In Wolff, E. W. and Bales, R. C. (eds.), Chemical Exchange Between the Atmosphere and Polar Snow, Springer, New York, Berlin, Heidelberg, pp. 173–199.Google Scholar
Wagenbach, D., et al. (1998a). Atmospheric near-surface nitrate at coastal Antarctic sites. J. Geophys. Res., 103(D9), 11007–11020.Google Scholar
Wagenbach, D., et al. (1998b). Sea-salt aerosol in coastal Antarctic regions. J. Geophys. Res. Atmos., 103, 10961–10974.CrossRefGoogle Scholar
Wagnon, P., et al. (1999). Loss of volatile acid species from upper firn layers at Vostok, Antarctica. J. Geophys. Res., 104(D3), 3423–3431.Google Scholar
Wegner, A., et al. (2012). Change in dust variability in the Atlantic sector of Antarctica at the end of the last deglaciation. Clim. Past, 8, 135–147.CrossRefGoogle Scholar
Welch, K., et al. (1993). Methanesulfonic acid in coastal Antarctic snow related to sea-ice extent. Geophys. Res. Lett., 20(6), 443–446.CrossRefGoogle Scholar
Weller, R., et al. (2011). Continuous 25-yr aerosol records at coastal Antarctica – I: Inter-annual variability of ionic compounds and links to climate indices. Tellus B: Chem. Phys. Meteorol., 63(5), 901–919. https://doi.org/10.1111/j.1600-0889.2011.00542.x.Google Scholar
Willis, M. D., et al. (2023). Polar oceans and sea ice in a changing climate. Elem. Sci. Anth., 11(1). https://doi.org/10.1525/elementa.2023.00056.CrossRefGoogle Scholar
Winski, D. A., et al. (2021). Seasonally resolved Holocene sea ice variability inferred from South Pole ice core chemistry. Geophys. Res. Lett., 48, e2020GL091602. https://doi.org/10.1029/2020GL091602.CrossRefGoogle Scholar
Winton, V. H. L., et al. (2014). The contribution of aeolian sand and dust to iron fertilization of phytoplankton blooms in southwestern Ross Sea, Antarctica. Global Biogeochem. Cycles, 28, 423–436. https://doi.org/10.1002/2013GB004574.CrossRefGoogle Scholar
Winton, V. H. L., et al. (2016). The origin of lithogenic sediment in the south-western Ross Sea and implications for iron fertilization. Antarct. Sci., 28(4), 250–260. https://doi.org/10.1017/S095410201600002X.CrossRefGoogle Scholar
Wolff, E. W. (2013). Ice sheets and nitrogen. Phil. Trans. R. Soc. B, 368, 20130127. http://dx.doi.org/10.1098/rstb.2013.0127.CrossRefGoogle ScholarPubMed
Wolff, E. W., et al. (2003). An ice core indicator of Antarctic sea ice production? Geophys. Res. Lett., 30(22), 2158. https://doi.org/10.1029/2003GL018454.CrossRefGoogle Scholar
Wolff, E. W., et al. (2006). Southern Ocean sea-ice extent, productivity and iron flux over the past eight glacial cycles. Nature, 440, 491–496. https://doi.org/10.1038/nature04614.CrossRefGoogle ScholarPubMed
Wu, X., et al. (2018). Long-range transport of volcanic aerosol from the 2010 Merapi tropical eruption to Antarctica. Atmos. Chem. Phys., 18, 15859–15877. https://doi.org/10.5194/acp-18-15859-2018.CrossRefGoogle Scholar
Xiao, C., et al. (2015). An ice-core record of Antarctic sea-ice extent in the Southern Indian Ocean for the past 300 years. Ann. Glaciol., 56, 451–455.CrossRefGoogle Scholar
Xin, Y., et al. (2008). Sea salt aerosol production and bromine release: Role of snow on sea ice. Geophys. Res. Lett., 35, L16815.Google Scholar
Zielinski, G. A., et al. (1994). Record of volcanism since 7000 B.C. from the GISP2 Greenland ice core and implications for the volcano-climate system. Science, 264, 948–952. https://doi.org/10.1126/science.264.5161.948.CrossRefGoogle ScholarPubMed
Anchukaitis, K. J., et al. (2020). An interpreted language implementation of the Vaganov–Shashkin tree-ring proxy system model. Dendrochronologia, 60, 125677. https://doi.org/10.1016/j.dendro.2020.125677.CrossRefGoogle Scholar
Annan, J. D. and Hargreaves, J. C. (2012). Identification of climatic state with limited proxy data. Clim. Past, 8, 1141–1151. https://doi.org/10.5194/cp-8-1141-2012.CrossRefGoogle Scholar
Badgeley, J. A., et al. (2020). Greenland temperature and precipitation over the last 20,000 years using data assimilation. Clim. Past, 16, 1335–1346. https://doi.org/10.5194/cp-2019-164.CrossRefGoogle Scholar
Balmaseda, M. A., et al. (2015). The Ocean Reanalyses Intercomparison Project (ORA-IP). J. Oper. Oceanogr., 8(supp1. 1), s80–s97. https://doi.org/10.1080/1755876X.2015.1022329.Google Scholar
Barkmeijer, J., et al. (2003). Forcing singular vectors and other sensitive model structures. Q. J. Roy. Meteor. Soc. 129(592), 2401–2423.CrossRefGoogle Scholar
Barnett, T. and Preisendorfer, R. (1978). Multifield analog prediction of short-term climate fluctuations using a climate state vector. J. Atmos. Sci., 35, 1771–1787. https://doi.org/10.1175/1520-0469(1978)035<1771:MAPOST>2.0.CO;2.2.0.CO;2>CrossRefGoogle Scholar
Battisti, D. S., et al. (2019). 100 years of progress in understanding the dynamics of coupled atmosphere–ocean variability. Chapter 8. Meteorol. Monogr., 59(1), 8.1–8.57. https://doi.org/10.1175/AMSMONOGRAPHS-D-18-0025.1.CrossRefGoogle Scholar
Ben Daoud, A., et al. (2016). Daily quantitative precipitation forecasts based on the analogue method: Improvements and application to a French large river basin. Atmos. Res., 169, 147–159. https://doi.org/10.1016/j.atmosres.2015.09.015.CrossRefGoogle Scholar
Bhend, J., et al. (2012). An ensemble-based approach to climate reconstructions. Clim. Past, 8(4), 963–976.CrossRefGoogle Scholar
Bjerknes, V. (1904). Das Problem der Wettervorhersage, betrachet vom Stanpunkt der Machanik und der Physik. Meteor Zeits, 21, 1–7.Google Scholar
Bothe, O. and Zorita, E. (2020). Technical note: The analogue method for millennial-scale, spatiotemporal climate reconstructions. Clim. Past Discuss., 20, 1–45. https://doi.org/10.5194/cp-2019-170.Google Scholar
Bourke, W., et al. (1985). Data assimilation. In Saltzman, B. (ed.), Advances in Geophysics Volume 28B, Issues in Atmospheric and Oceanic Modelling. Part B weather dynamics. Academic Press, London, pp. 123–155.Google Scholar
Bouttier, F. and Coutier, P. (2002). Data assimilation concepts and methods, March 1999. Meteorological Training Course Lecture Series, ECMWF. 59pp. (accessed from www.ecmwf.int/sites/default/files/elibrary/2002/16928-data-assimilation-concepts-and-methods.pdf).Google Scholar
Brasseur, P., et al. (eds.). (2006). Ocean Weather Forecasting. Springer, Netherlands, pp. 271–316.Google Scholar
Brönnimann, S., et al. (2013). Transient state estimation in paleoclimatology using data assimilation. PAGES News, 21, 74–75.CrossRefGoogle Scholar
Cardinali, et al. (2013). Observation influence diagnostic of a data assimilation system. Chapter 4. In Park, S. K. and Xu, L. (eds.), Data Assimilation for Atmospheric, Oceanic and Hydrologic Applications (Vol. II). Springer-Verlag, Berlin Heidelberg, pp. 89–110. https://doi.org/10.1007/978-3-642-35088-7.Google Scholar
Chandresakar, A. and Kutty, M. G. (2013). Studies on the impacts of 3D-VAR assimilation of satellite observations on the simulation of monsoon depressions over India. Chapter 26. In Park, S. K. and Xu, L. (eds.), Data Assimilation for Atmospheric, Oceanic and Hydrologic Applications (Vol. II). Springer-Verlag Berlin Heidelberg, pp. 643–705. https://doi.org/10.1007/978-3-642-35088-7.Google Scholar
Clem, K. R., et al. (2020). Record warming at the South Pole during the past three decades. Nat. Clim. Change, 10, 762–770. https://doi.org/10.1038/s41558-020-0815-z.CrossRefGoogle Scholar
Collins, M., et al. (2001). The internal climate variability of HadCM3, a version of the Hadley Centre coupled model without flux adjustments. Clim. Dyn., 17(1), 61–81. https://doi.org/10.1007/s003820000094.CrossRefGoogle Scholar
Comeau, D., et al. (2019). Predicting regional and pan-Arctic sea ice anomalies with kernel analog forecasting. Clim. Dyn., 52, 5507–5525. https://doi.org/10.1007/s00382-018-4459-x.CrossRefGoogle Scholar
Compo, G. P., et al. (2006). Feasibility of a 100 year reanalysis using only surface pressure data. Bull. Amer. Meteor. Soc., 87, 175–190.CrossRefGoogle Scholar
Compo, G. P., et al. (2011). The Twentieth Century Reanalysis Project. Q. J. Roy. Meteorol. Soc., 137, 1–28.CrossRefGoogle Scholar
Croke, J., et al. (2021). A palaeoclimate proxy database for water security planning in Queensland Australia. Sci. Data, 8, 292. https://doi.org/10.1038/s41597-021-01074-8.CrossRefGoogle ScholarPubMed
Dalaiden, Q., et al. (2021). Reconstructing atmospheric circulation and sea-ice extent in the West Antarctic over the past 200 years using data assimilation. Clim. Dyn., 57, 3479–3503. https://doi.org/10.1007/s00382-021-05879-6.CrossRefGoogle Scholar
Dee, D. P., et al. (2011). The ERA-interim reanalysis: Configuration and performance of the data assimilation system. Q. J. Roy. Meteorol. Soc., 137(656), 553–597. https://doi.org/10.1002/qj.828.CrossRefGoogle Scholar
Dee, S., et al. (2015). PRYSM: An open-source framework for PRoxY System Modeling, with applications to oxygen- isotope systems. J. Adv. Model. Earth Syst., 7, 1220–1247. https://doi.org/10.1002/2015MS000447.CrossRefGoogle Scholar
Dee, S., et al. (2016). The utility of proxy system modeling in estimating climate states over the Common Era. J. Adv. Model. Earth Syst., 8, 1164–1179. https://doi.org/10.1002/2016MS000677.CrossRefGoogle Scholar
Dee, S. G., et al. (2017). Improved spectral comparisons of paleoclimate models and observations via proxy system modeling: Implications for multi-decadal variability. Earth Planet. Sci. Lett., 476, 34–46.CrossRefGoogle Scholar
Dee, S. G., et al. (2018). PRYSM v2.0: A proxy system model for lacustrine archives. Paleoceanogr. Paleoclimatol., 33(11), 1250–1269. https://doi.org/10.1029/2018pa003413.CrossRefGoogle Scholar
Dee, S. G., et al. (2021). Hot air, hot lakes, or both? exploring Mid- Holocene African temperatures using proxy system modeling. J. Geophys. Res. Atmos., 126, e2020JD033269. https://doi.org/10.1029/2020JD033269.CrossRefGoogle Scholar
Diaz, H., et al. (2016). A five-century reconstruction of Hawaiian Islands winter rainfall. J. Clim., 29, 5661–5674. https://doi.org/10.1175/JCLI-D-15-0815.1.CrossRefGoogle Scholar
Ding, H., et al. (2018). Skillful climate forecasts of the Tropical Indo-Pacific Ocean using model-analogs. J. Clim., 31, 5437–5459.CrossRefGoogle Scholar
Dirren, S. and Hakim, G. J. (2005). Toward the assimilation of time-averaged observations. Geophys. Res. Lett., 32(4), L04804. https://doi.org/10.1029/2004gl021444.CrossRefGoogle Scholar
Emile-Geay, J., et al. (2019). The Linked Earth Ontology: A modular, extensible representation of open paleoclimate data. Zenodo. http://doi.org/10.5281/zenodo.2577604.CrossRefGoogle Scholar
Evans, M. N., et al. (2013). Applications of proxy system modeling in high resolution paleoclimatology. Quat. Sci. Rev., 76, 16–28.CrossRefGoogle Scholar
Fang, M. and Li, X. (2016). Paleoclimate data assimilation: Its motivation, progress and prospects. Sci. China Earth Sci., 59(9), 1817–1826. https://doi.org/10.1007/s11430-015-5432-6.CrossRefGoogle Scholar
Fang, M. and Li, X. (2019). An artificial neural networks-based tree ring width proxy system model for paleoclimate data assimilation. J. Adv. Model. Earth Syst., 11, 892–904. https://doi.org/10.1029/2018MS001525.CrossRefGoogle Scholar
Felden, J., et al. (2023). PANGAEA – Data Publisher for Earth & Environmental Science. Sci. Data, 10(1), 347. https://doi.org/10.1038/s41597-023-02269-x.CrossRefGoogle ScholarPubMed
Flückiger, S., et al. (2017). Simulating crop yield losses in Switzerland for historical and present Tambora climate scenarios. Environ. Res. Lett., 12(7), 74026.CrossRefGoogle Scholar
Franke, J., et al. (2011). 200 years of European temperature variability: Insights from and tests of the proxy surrogate reconstruction analog method. Clim. Dyn., 37(1–2), 133–150. https://doi.org/10.1007/s00382-010-0802-6.CrossRefGoogle Scholar
Franke, J., et al. (2017). A monthly global paleo-reanalysis of the atmosphere from 1600 to 2005 for studying past climatic variations. Sci. Data, 4, 170076. https://doi.org/10.1038/sdata.2017.76.CrossRefGoogle ScholarPubMed
Franke, J., et al. (2020). The importance of input data quality and quantity in climate field reconstructions-results from the assimilation of various tree-ring collections. Clim. Past, 16, 1061–1074. https://doi.org/10.5194/cp-16-1061-2020.CrossRefGoogle Scholar
Fukumori, I. (2006). What is data assimilation really solving, and how is the calculation actually done? In Chassignet, E. P. and Verron, J. (eds.), Ocean Weather Forecasting, 317–342.Google Scholar
Gebhardt, C., et al. (2008). Reconstruction of Quaternary temperature fields by dynamically consistent smoothing. Clim. Dyn., 30, 421–437. https://doi.org/10.1007/s00382-007-0299-9.CrossRefGoogle Scholar
Gelb, A. (1974). Applied Optimal Estimation. The Analytic Science Corporation/MIT, Cambridge, p 374.Google Scholar
Ghil, M. and Malanotte-Rizzoli, P. (1991). Data assimilation in meteorology and oceanography. Adv. Geophys., 33, 141–266.CrossRefGoogle Scholar
Gomez-Navarro, J. J., et al. (2017). Pseudo-proxy tests of the analogue method to reconstruct spatially resolved global temperature during the Common Era. Clim. Past, 13, 629–648. https://doi.org/10.5194/cp-13-629-2017.CrossRefGoogle Scholar
Goodwin, I. D., et al. (2014a). Climate windows for Polynesian voyaging to New Zealand and Easter Island. P. Natl. A. Sci. USA, 111(41), 14716–14721. https://doi.org/10.1073/pnas.1408918111.CrossRefGoogle Scholar
Goodwin, I. D., et al. (2014b). A reconstruction of extratropical Indo-Pacific sea-level pressure patterns during the Medieval Climate Anomaly. Clim. Dyn., 43(5–6), 1197–1219. https://doi.org/10.1007/s00382-013-1899-1.CrossRefGoogle Scholar
Goodwin, I. D., et al. (in prep 2025). Tasman Sea wave climate, coastal impacts and extreme storm activity in the context of large-scale atmospheric circulation, since 1500 CE.Google Scholar
Goosse, H. (2016). An additional step toward comprehensive paleoclimate reanalyses, J. Adv. Model. Earth Syst., 8, 1501–1503. https://doi.org/10.1002/2016MS000739.CrossRefGoogle Scholar
Goosse, H., et al. (2006). Using paleo-climate proxy-data to select optimal realisations in an ensemble of simulations of the climate of the past millennium. Clim. Dyn., 27(2–3), 165–184.CrossRefGoogle Scholar
Goosse, H., et al. (2010). Reconstructing surface temperature changes over the past 600 years using climate model simulations with data assimilation. J. Geophys. Res. Atmos., 115, D09108. https://doi.org/10.1029/2009jd012737.CrossRefGoogle Scholar
Graham, N. E., et al. (2007). Tropical Pacific – Mid-latitude teleconnections in medieval times. Clim. Change, 83(1–2), 241–285.CrossRefGoogle Scholar
Hakim, G. J., et al. (2013). Overview of data assimilation methods. PAGES News, 21(2), 72–73.CrossRefGoogle Scholar
Hakim, G. J., et al. (2016). The Last Millennium Climate Reanalysis Project: Framework and first results. J. Geophys. Res. Atmos., 121, 6745–6764. https://doi.org/10.1002/2016JD024751.CrossRefGoogle Scholar
Hibbert, F. D., et al. (2016). Coral indicators of past sea-level change: A global repository of U-series dated benchmarks. Quat. Sci. Rev., 145, 1–56. https://doi.org/10.1016/j.quascirev.2016.04.019.CrossRefGoogle Scholar
Hibbert, F. D., et al. (2018). A database of biological and geomorphological sea-level markers from the Last Glacial Maximum to present. Sci. Data, 5, 180088. https://doi.org/10.1038/sdata.2018.88.CrossRefGoogle ScholarPubMed
Horton, P., et al. (2012). Spatial relationship between the atmospheric circulation and the precipitation measured in the western Swiss Alps by means of the analogue method. Nat. Hazards Earth Syst. Sci., 12, 777–784. https://doi.org/10.5194/nhess-12-777-2012.CrossRefGoogle Scholar
Horton, P., et al. (2017). The analogue method for precipitation prediction: Finding better analogue situations at a sub-daily time step. Hydrol. Earth Syst. Sci., 21, 3307–3323. https://doi.org/10.5194/hess-21-3307-2017.CrossRefGoogle Scholar
Huntley, H. S. and Hakim, G. J. (2010). Assimilation of time-averaged observations in a quasi-geostrophic atmospheric jet model. Clim. Dyn., 35, 995–1009. https://doi.org/10.1007/s00382-009-0714-5.CrossRefGoogle Scholar
Jensen, M. F., et al. (2018). A spatiotemporal reconstruction of sea-surface temperatures in the North Atlantic during Dansgaard–Oeschger events 5–8. Clim. Past, 14(6), 901–922. https://doi.org/10.5194/cp-14-901-2018.CrossRefGoogle Scholar
Jones, M. and Dee, S. G. (2018). Global-scale proxy system modelling of oxygen isotopes in lacustrine carbonates: New insights from isotope-enabled-model proxy-data comparison. Quat. Sci. Rev., 202,12, 19–29. https://doi.org/10.1016/j.quascirev.2018.09.009.CrossRefGoogle Scholar
Kalman, R. E. (1960). A new approach to linear filtering and prediction problems. Trans. ASME, J. Basic Eng., 35–45.CrossRefGoogle Scholar
Kalnay, E. (2003). Atmospheric Modeling, Data Assimilation, and Predictability. Cambridge University Press. https://doi.org/10.1256/00359000360683511.Google Scholar
Kalnay, E., et al. (1996). The NCEP/NCAR 40-year reanalysis project. B. Am. Meteorol. Soc., 77, 437–471. https://doi.org/10.1175/1520-0477(1996)077<0437:tnyrp>2.0.co;2.2.0.CO;2>CrossRefGoogle Scholar
Kang, W., et al. (2013). A survey of observers for nonlinear dynamical systems. Chapter 1. In Park, S. K. and Xu, L. (eds.), Data Assimilation for Atmospheric, Oceanic and Hydrologic Applications (Vol. II). Springer-Verlag, Berlin Heidelberg, pp. 1–25. https://doi.org/10.1007/978-3-642-35088-7.Google Scholar
Kaufman, D., et al. (2020). A global database of Holocene paleotemperature records. Sci. Data, 7, 115. https://doi.org/10.1038/s41597-020-0445-3.CrossRefGoogle ScholarPubMed
Khatiwala, S., et al. (2001). Age tracers in an ocean GCM. Deep Sea Res. Part I, 48, 1423–1441.CrossRefGoogle Scholar
Khider, D., et al. (2019). PaCTS 1.0: A crowdsourced reporting standard for paleoclimate data. Paleoceanograph. Paleoclimatol., 34, 1570–1596. https://doi.org/10.1029/2019PA003632.CrossRefGoogle Scholar
Klein, F. and Goosse, H. (2017). Reconstructing East African rainfall and Indian Ocean sea surface temperatures over the last centuries using data assimilation. Clim. Dyn., 1(12), 1–21. https://doi.org/10.1007/s00382-017-3853-0.Google Scholar
Klein, F., et al. (2019). Assessing the robustness of Antarctic temperature reconstructions over the past two millennia using pseudoproxy and data assimilation experiments. Clim. Past, 15, 661–684. https://doi.org/10.5194/cp-2018-90.CrossRefGoogle Scholar
Kruizinga, S. and Murphy, A. H. (1983). Use of an analogue procedure to formulate objective probabilistic temperature forecasts in the Netherlands. Mon. Weather Rev., 111, 2244–2254.2.0.CO;2>CrossRefGoogle Scholar
Lahoz, W. and Schneider, P. (2014). Data assimilation: Making sense of Earth observation. Front. Environ. Sci., 2, 16. https://doi.org/10.3389/fenvs.2014.00016.CrossRefGoogle Scholar
Landrum, L., et al. (2013). Last millennium climate and its variability in CCSM4. J. Clim., 26, 1085–1111. https://doi.org/10.1175/JCLI-D-11-00326.1.CrossRefGoogle Scholar
Le Dimet, F. X. and Talagrand, O. (1986). Variational algorithms for analysis and assimilation of meteorological observations. Tellus, 38A, 97–110.Google Scholar
Le Dimet, F. X., et al. (2017). Variational data assimilation: Optimization and optimal control. Chapter 1. In Park, S. K. and Xu, L. (eds.), Data Assimilation for Atmospheric, Oceanic and Hydrologic Applications (Vol. III). pp. 1–55. https://doi.org/10.1007/978-3-319-43415-5_1.Google Scholar
Lewis, J. M., et al. (2006). Dynamic Data Assimilation: A Least Squares Approach. Cambridge University Press, Cambridge, 654pp.CrossRefGoogle Scholar
Lewis, J. M. and Lakshmivarahan, S. (2008). Sasaki’s pivotal contribution: Calculus of variations applied to weather map analysis. Mon. Weather Rev., 136, 3553–3567. https://doi.org/10.1175/2008MWR2400.1.CrossRefGoogle Scholar
Lewis, J. M. and Lakshmivarahan, S. (2013). A question of adequacy of observations in variational data assimilation. Chapter 5. In Park, S. K. and Xu, L. (eds.), Data Assimilation for Atmospheric, Oceanic and Hydrologic Applications (Vol. II). Springer-Verlag, Berlin Heidelberg, pp. 111–124. https://doi.org/10.1007/978-3-642-35088-7.Google Scholar
Liu, H., et al. (2017). A systematic comparison of particle filter and EnKF in assimilating time–averaged observations. J. Geophys. Res. Atmos., 122, 13155–13173. https://doi.org/10.1002/2017JD026798.CrossRefGoogle Scholar
Lomb, N. (1975). A spectrographic study of beta Centauri. Mon. Not. R. Astron. Soc., 172(3), 639–647.CrossRefGoogle Scholar
Lorenz, E. (1956). Empirical orthogonal functions and statistical weather prediction. Tech. Rep., Massachusetts Institute of Technology, Department of Meteorology.Google Scholar
Lorenz, E. (1969). Atmospheric predictability as revealed by naturally occurring analogues. J. Atmos. Sci., 26, 636–646. https://doi.org/10.1175/1520-0469(1969)26<636:aparbn>2.0.co;2.2.0.CO;2>CrossRefGoogle Scholar
Mairesse, A., et al. (2013). Investigating the consistency between proxy-based reconstructions and climate models using data assimilation: A mid-Holocene case study. Clim. Past, 9, 2741–2757. www.clim-past.net/9/2741/2013/.CrossRefGoogle Scholar
Mann, M., et al. (2007). Robustness of proxy-based climate field reconstruction methods. J. Geophys. Res., 112(D12), D12109.38.Google Scholar
Matsikaris, A., et al. (2015). On-line and off-line data assimilation in palaeoclimatology: A case study. Clim. Past, 11, 81–93. https://doi.org/10.5194/cp-11-81-2015.CrossRefGoogle Scholar
McKay, N. P. and Emile-Geay, J. (2016). Technical note: The Linked Paleo Data framework-a common tongue for paleoclimatology. Clim. Past, 12, 1093–1100.CrossRefGoogle Scholar
Menviel, L., et al. (2011). Deconstructing the Last Glacial termination: The role of millennial and orbital-scale forcings. Quat. Sci. Rev., 30, 1155–1172. https://doi.org/10.1016/j.quascirev.2011.02.005.CrossRefGoogle Scholar
Moore, A. M., et al. (2013). A 4D-Var analysis system for the California Current: A prototype for an operational regional ocean data assimilation system. In Park, S. K. and Xu, L. (eds.), Data Assimilation for Atmospheric, Oceanic and Hydrologic Applications (Vol. II). Springer-Verlag, Berlin Heidelberg, https://doi.org/10.1007/978-3-642-35088-7.Google Scholar
Neukom, R., et al. (2019). No evidence for globally coherent warm and cold periods over the preindustrial Common Era. Nature, 571, 550–554. https://doi.org/10.1038/s41586-019-1401-2.CrossRefGoogle ScholarPubMed
Okazaki, A. and Yoshimura, K. (2017). Development and evaluation of a system of proxy data assimilation for paleoclimate reconstruction. Clim. Past, 13, 379–393. https://doi.org/10.5194/cp-13-379-2017.CrossRefGoogle Scholar
Okazaki, A. and Yoshimura, K. (2019). Global evaluation of proxy system models for stable water isotopes with realistic atmospheric forcing. J. Geophys. Res. Atmos., 124, 8972–8993. https://doi.org/10.1029/2018JD029463.CrossRefGoogle Scholar
Otto-Bliesner, B. L., et al. (2016). Climate variability and change since 850 CE. An ensemble approach with the Community Earth System Model (CESM). B. Am. Meteor. Soc., 735–754. https://doi.org/10.1175/BAMS-D-14-00233.1.CrossRefGoogle Scholar
Park, S. K. and Xu, L. (eds.). (2013). Data Assimilation for Atmospheric, Oceanic and Hydrologic Applications (Vol. II). Springer-Verlag, Berlin Heidelberg. https://doi.org/10.1007/978-3-642-35088-7.CrossRefGoogle Scholar
Perkins, W. A. and Hakim, G. J. (2017). Reconstructing paleoclimate fields using online data assimilation with a linear inverse model. Clim. Past, 13, 421–436. www.clim-past.net/13/421/2017/.CrossRefGoogle Scholar
Pfister, L., et al. (2020). Statistical reconstruction of daily precipitation and temperature fields in Switzerland back to 1864. Clim. Past, 16(2), 663–678. https://doi.org/10.5194/cp-16-663-2020.CrossRefGoogle Scholar
Rössler, O. and Brönnimann, S. (2018). The effect of the Tambora eruption on Swiss flood generation in 1816/1817. Sci. Total Environ., 627, 1218–1227. https://doi.org/10.1016/j.scitotenv.2018.01.254.CrossRefGoogle ScholarPubMed
Routray, A., et al. (2016). Introduction to data assimilation techniques and ensemble Kalman Filter. Chapter 11. In Mohanty, U. C., et al., Advanced Numerical Modeling and Data Assimilation Techniques for Tropical Cyclone Prediction. Capital Publishing Company, pp. 307–330. https://doi.org/10.5822/978-94-024-0896-6_11.Google Scholar
Rovere, A., et al. (2023). The World Atlas of Last Interglacial Shorelines (version 1.0). Earth Syst. Sci. Data, 15, 1–23. https://doi.org/10.5194/essd-15-1-2023.CrossRefGoogle Scholar
Scaife, A. A. and Smith, D. (2018). A signal-to-noise paradox in climate science. npj Clim. Atmos. Sci., 1, 28. https://doi.org/10.1038/s41612-018-0038-4.CrossRefGoogle Scholar
Scargle, J. D. (1982). Studies in astronomical time series analysis II. Statistical aspects of spectral analysis of unevenly spaced data. Astrophys. J., 263, 835–853.CrossRefGoogle Scholar
Schenk, F. and Zorita, E. (2012). Reconstruction of high resolution atmospheric fields for Northern Europe using analog-upscaling. Clim. Past, 8(5), 1681–1703. https://doi.org/10.5194/cp-8-1681-2012.CrossRefGoogle Scholar
Schneider, D. and Deser, C. (2018). Tropically driven and externally forced patterns of Antarctic sea ice change: Reconciling observed and modeled trends. Clim. Dyn., 50, 4599–4618. https://doi.org/10.1007/s00382-017-3893-5.CrossRefGoogle Scholar
Steiger, N. J., et al. (2014). Assimilation of time-averaged pseudo- proxies for climate reconstruction. J. Clim., 27, 426–441. https://doi.org/10.1175/JCLI-D-12-00693.1.CrossRefGoogle Scholar
Steiger, N. and Hakim, G. (2016). Multi-timescale data assimilation for atmosphere–ocean state estimates. Clim. Past, 12, 1375–1388. www.clim-past.net/12/1375/2016/.CrossRefGoogle Scholar
Steiger, N. J. and Smerdon, J. E. (2017). A pseudoproxy assessment of data assimilation for reconstructing the atmosphere–ocean dynamics of hydroclimate extremes. Clim. Past, 13, 1435–1449. https://doi.org/10.5194/cp-13-1435-2017.CrossRefGoogle Scholar
Steiger, N. J., et al. (2018). A reconstruction of global hydroclimate and dynamical variables over the Common Era. Sci. Data, 5(1), 180086.CrossRefGoogle ScholarPubMed
Stevens, B., et al. (2013). Atmospheric component of the Earth System Model: ECHAM6. J. Adv. Model. Earth Syst., 5, 146–172. https://doi.org/10.1002/jame.20015.CrossRefGoogle Scholar
Sun, H., et al. (2022). An analog offline EnKF for paleoclimate data assimilation. J. Adv. Model. Earth Syst., 14, e2021MS002674. https://doi.org/10.1029/2021MS002674.CrossRefGoogle Scholar
Tardiff, R., et al. (2019). Last Millennium Reanalysis with an expanded proxy database and seasonal proxy modeling. Clim. Past, 15, 1251–1273. https://doi.org/10.5194/cp-15-1251-2019.Google Scholar
Tebaldi, C. and Knutti, R. (2007). The use of the multi-model ensemble in probabilistic climate projections. Phil. Trans. Roy. Soc. A: Math. Phys. Eng. Sci., 365(1857), 2053–2075. https://doi.org/10.1098/rsta.2007.2076.Google ScholarPubMed
Thomson, D. J. (1982). Spectrum estimation and harmonic analysis. Proc. IEEE, 70(9), 1055–1096.CrossRefGoogle Scholar
Trouet, V., et al. (2009a). Interannual variations in fire weather, fire extent, and synoptic-scale circulation patterns in northern California and Oregon. Theor. Appl. Climatol., 95, 349–360.CrossRefGoogle Scholar
Trouet, V., et al. (2009b). Persistent positive North Atlantic Oscillation mode dominated the Medieval Climate Anomaly. Science, 324, 78–80. https://doi.org/10.1126/science.1166349.CrossRefGoogle Scholar
Valler, V., et al. (2019). Impact of different estimations of the background-error covariance matrix on climate reconstructions based on data assimilation. Clim. Past, 15(4), 1427–1441. https://doi.org/10.5194/cp-15-1427-2019.CrossRefGoogle Scholar
van den Dool, H. M. (1994). Searching for analogues, how long must we wait? Tellus A, 46(3), 314–324. https://doi.org/10.1034/j.1600-0870.1994.t01-2-00006.x.CrossRefGoogle Scholar
van der Schrier, G. and Barkmeijer, J. (2005). Bjerknes’ hypothesis on the coldness during AD 1790–1820 revisited. Clim. Dyn., 25, 537–553.CrossRefGoogle Scholar
Von Storch, H. and van Zwiers, F. W. (1999). Statistical Analysis in Climate Research. Cambridge University Press, Cambridge, 484pp.Google Scholar
von Storch, H., et al. (2000). Combining paleoclimatic evidence and GCMs by means of Data Assimilation Through Upscaling and Nudging. In Proc. 11th Symp. Glob. Clim. Change Studies, AMS Long Beach, CA, pp. 28–31. [Available at www.hvonstorch.de/klima/pdf/DATUN-2000.pdf.]Google Scholar
Wahl, E., et al. (2019). Jet stream dynamics, hydroclimate, and fire in California from 1600 CE to present. Proc. Natl. Acad. Sci. USA, 116(12), 5393–5398. www.pnas.org/cgi/doi/10.1073/pnas.1815292116.CrossRefGoogle ScholarPubMed
Wahl, E., et al. (2022). The Kalman Filter as post-processor for analog data–model assimilation in paleoclimate reconstruction. J. Climate, 35, 5501–5518. https://doi.org/10.1175/JCLI-D-21-0454.1.CrossRefGoogle Scholar
Whitaker, J. S., et al. (2008). Ensemble data assimilation with the NCEP Global Forecast System. Mon. Weather Rev., 136, 463–482. https://doi.org/10.1175/2007MWR2018.1.CrossRefGoogle Scholar
Widmann, M., et al. (2010). Using data assimilation to study extratropical Northern Hemi- sphere climate over the last millennium. Clim. Past, 6, 627–644. https://doi.org/10.5194/cp-6-627-2010.CrossRefGoogle Scholar
Wong, C. I. and Breecker, D. O. (2018). Advancements in the use of speleothems as climate archives. Quat. Sci. Rev., 127, 1–18. https://doi.org/10.1016/j.quascirev.2015.07.019.Google Scholar
Wu, Z., et al. (2007). On the trend, detrending, and variability of nonlinear and nonstationary time series. Proc. Natl. Acad. Sci. USA, 104(38), 14889–14894.CrossRefGoogle ScholarPubMed
Zorita, E., et al. (1995). Stochastic characterisation of regional circulation patterns for climate model diagnosis and estimation of local precipitation. J. Climate, 8(5), 1023–1042.2.0.CO;2>CrossRefGoogle Scholar
Zorita, E. and von Storch, H. (1999). The analog method as a simple statistical downscaling technique: Comparison with more complicated methods. J. Climate, 12, 2474–2489.2.0.CO;2>CrossRefGoogle Scholar
Abram, N. J., et al. (2014). Evolution of the Southern Annular Mode during the past millennium. Nat. Clim. Change, 4, 564–569. https://doi.org/10.1038/NCLIMATE2235.CrossRefGoogle Scholar
Anchukaitis, K. J., et al. (2010). Influence of volcanic eruptions on the climate of the Asian monsoon region. Geophys. Res. Lett., 37, L22703. https://doi.org/10.1029/2010GL044843.CrossRefGoogle Scholar
Anderson, D. M., et al. (2019). Additions to the Last Millennium reanalysis multi-proxy database. Data Sci. J., 18(2), 1–11. https://doi.org/10.5334/dsj-2019-002.CrossRefGoogle Scholar
Boisier, J. P., et al. (2018). Anthropogenic drying in central-southern Chile evidenced by long-term observations and climate model simulations. Elem. Sci. Anth., 6, 74. https://doi.org/10.1525/elementa.328.CrossRefGoogle Scholar
Bracegirdle, T. J., et al. (2019). Back to the future: Using long-term observational and paleo-proxy reconstructions to improve model projections of Antarctic climate. Geosci., 9, 255. https://doi.org/10.3390/geosciences9060255.CrossRefGoogle Scholar
Braganza, K., et al. (2009). A multi-proxy index of the El Niño – Southern Oscillation, A.D. 1525–1982. J. Geophys. Res., 114(D5), D05106.Google Scholar
Catto, J. L., et al. (2019). The future of midlatitude cyclones. Curr. Clim. Change Rep., 5, 407–420. https://doi.org/10.1007/s40641-019-00149-4.CrossRefGoogle Scholar
Chadwick, M., et al. (2020). Analysing the timing of peak warming and minimum winter sea-ice extent in the Southern Ocean during MIS 5e. Quat. Sc. Rev., 229, https://doi.org/10.1016/j.quascirev.2019.106134.CrossRefGoogle Scholar
Chang, E. K. M. (2017). Projected significant increase in the number of extreme extratropical cyclones in the Southern Hemisphere. J. Clim., 30, 4915–4935. https://doi.org/10.1175/JCLI-D-16-0553.1.CrossRefGoogle Scholar
Chemke, R. (2022). The future poleward shift of Southern Hemisphere summer mid-latitude storm tracks stems from ocean coupling. Nat. Clim. Change, 13(1), 1730–1730. https://doi.org/10.1038/s41467-022-29392-4.Google ScholarPubMed
Chiswell, S. M. (2021). Atmospheric wave number-4 driven South Pacific marine heat waves and marine cool spells. Nature Comms., 12, 4779. https://doi.org/10.1038/s41467-021-25160-y.CrossRefGoogle Scholar
Christiansen, B. and Ljungqvist, F. C. (2017). Challenges and perspectives for large-scale temperature reconstructions of the past two millennia. Rev. Geophys., 55, 40–96. https://doi.org/10.1002/2016RG000521.CrossRefGoogle Scholar
Clem, K. R. and Fogt, R. L. (2013). Varying roles of ENSO and SAM on the Antarctic Peninsula climate in austral spring. J. Geophys. Res. Atmos., 118, 11481–11492. https://doi.org/10.1002/jgrd.50860.CrossRefGoogle Scholar
Clement, A., et al. (1996). An ocean dynamical thermostat. J. Clim., 9, 2190–2196.2.0.CO;2>CrossRefGoogle Scholar
Coats, S., et al. (2015a). North American pancontinental droughts in model simulations of the last millennium. J. Clim., 28, 2025–2043. https://doi.org/10.1175/jcli-d-14-00634.1.CrossRefGoogle Scholar
Coats, S., et al. (2015b). Are simulated megadroughts in the North American southwest forced?, J. Clim., 28, 124–142. https://doi.org/10.1175/jcli-d-14-00071.1.CrossRefGoogle Scholar
Cobb, K. M., et al. (2001). A central tropical Pacific coral demonstrates Pacific, Indian, and Atlantic decadal climate connections. Geophys. Res. Lett., 28, 2209–2212. https://doi.org/10.1029/2001GL012919.CrossRefGoogle Scholar
Cobb, K. M., et al. (2003). El Niño/Southern Oscillation and tropical Pacific climate during the last millennium. Nature, 424, 271–276. https://doi.org/10.1038/nature01779.CrossRefGoogle ScholarPubMed
Cobb, K. M., et al. (2013). Highly variable El Niño-Southern Oscillation throughout the Holocene. Science, 339, 67–70. http://dx.doi.org/10.1126/science.1228246.CrossRefGoogle ScholarPubMed
Cole, J. E., et al. (1993). Recent variability in the Southern Oscillation isotopic results from a Tarawa atoll coral. Science, 260, 1790–1793.CrossRefGoogle ScholarPubMed
Cook, B. I., et al. (2015). Unprecedented 21st century drought risk in the American Southwest and Central Plains, Sci. Adv., 1, e1400082. https://doi.org/10.1126/sciadv.1400082.CrossRefGoogle ScholarPubMed
Cook, E. R., et al. (2002). A well-verified, multiproxy reconstruction of the winter North Atlantic oscillation index since A.D. 1400. J.Clim., 15, 1754–1764.2.0.CO;2>CrossRefGoogle Scholar
Cortese, G., et al. (2013). Southwest Pacific Ocean response to a warmer world: Insights from Marine Isotope Stage 5e. Paleoceanography, 28(3), 585–598.CrossRefGoogle Scholar
Cotté, C. and Guinet, C. (2007). Historical whaling records reveal major regional retreat of Antarctic sea ice. Deep Sea Res., 54, 243–252. https://doi.org/10.1016/j.dsr.2006.11.001.CrossRefGoogle Scholar
Crosta, X., et al. (2021). Multi-decadal trends in Antarctic sea-ice extent driven by ENSO–SAM over the last 2,000 years. Nature, 14, 156–160. https://doi.org/10.1038/s41561-021-00697-1.Google Scholar
D’Arrigo, R., et al. (2005). On the variability of ENSO over the past six centuries. Geophys. Res. Lett., 32, L03711. https://doi.org/10.1029/2004GL022055.Google Scholar
Dätwyler, C., et al. (2018). Teleconnection stationarity, variability and trends of the Southern Annular Mode (SAM) during the last millennium. Clim. Dyn., 51, 2321–2339. https://doi.org/10.1007/s00382-017-4015-0.CrossRefGoogle Scholar
Dätwyler, C., et al. (2020). Teleconnections and relationship between the El Niño–Southern Oscillation (ENSO) and the Southern Annular Mode (SAM) in reconstructions and models over the past millennium. Clim. Past, 16, 743–756. https://doi.org/10.5194/cp-16-743-2020.CrossRefGoogle Scholar
Dee, S. G., et al. (2020a). Enhanced North American ENSO teleconnections during the Little Ice Age revealed by paleoclimate data assimilation. Geophys. Res. Lett., 47, e2020GL087504. https://doi.org/10.1029/2020GL087504.CrossRefGoogle Scholar
Dee, S. G., et al. (2020b). No consistent ENSO response to volcanic forcing over the last millennium. Science, 367(6485), 1477–1481. https://doi.org/10.1126/science.aax2000.CrossRefGoogle Scholar
Dee, S. G. and Steiger, N. J. (2022). ENSO’s response to volcanism in a data assimilation-based paleoclimate reconstruction over the Common Era. Paleoceanogr. Paleoclimatol., 37, e2021PA004290. https://doi.org/10.1029/2021PA004290.CrossRefGoogle Scholar
de la Mare, W. K. (1997). Abrupt mid-twentieth century decline in Antarctic sea-ice extent from whaling records. Nature, 389, 57–60.Google Scholar
de la Mare, W. K. (2002). Whaling records and sea ice: Consistency with historical records. Polar Rec., 38, 355–358.CrossRefGoogle Scholar
de la Mare, W. K. (2009). Changes in Antarctic sea-ice extent from direct historical observations and whaling records. Clim. Change, 92, 461–493. https://doi.org/10.1007/s10584-008-9473-2.CrossRefGoogle Scholar
Diaz, H. F. and Markgraf, V. (eds.). (1992). El Niño: Historical and Paleoclimatic Aspects of the Southern Oscillation. Cambridge University Press, Cambridge.Google Scholar
Diaz, H. F. and Markgraf, V. (eds.). (2000). El Niño and the Southern Oscillation; Multiscale Variability and Global and Regional Impacts. Cambridge University Press, Cambridge.Google Scholar
Di Lorenzo, E. and Mantua, N. (2016). Multi-year persistence of the 2014/15 North Pacific marine heatwave. Nat. Climate Change, 6, 1042–1048.CrossRefGoogle Scholar
Doblas-Reyes, F. J., et al. (2021). Linking global to regional climate change. Chapter 10. In Masson-Delmotte, V., et al. (eds.), Climate Change 2021: The Physical Science Basis. Cambridge University Press, Cambridge, pp. 1363–1512.Google Scholar
Domeisen, D. I., et al. (2019). The teleconnection of El Niño Southern Oscillation to the stratosphere. Rev. Geophys., 57, 5–47. https://doi.org/10.1029/2018RG000596.CrossRefGoogle Scholar
Edinburgh, T. and Day, J. J. (2016). Estimating the extent of Antarctic summer sea ice during the Heroic Age of Antarctic Exploration. Cryosphere, 10, 2721–2730. www.the-cryosphere.net/10/2721/2016/. https://doi.org/10.5194/tc-10-2721-2016.CrossRefGoogle Scholar
Emile-Geay, J., et al. (2013a). Estimating central equatorial Pacific SST variability over the past millennium: Part I. Methodology and validation. J. Clim., 26(7), 2302–2328.Google Scholar
Emile-Geay, J., et al. (2013b). Estimating central equatorial Pacific SST variability over the past millennium: Part II. Reconstructions and implications. J. Clim., 26(7), 2329–2352.Google Scholar
Emile-Geay, J., et al. (2016). Links between tropical Pacific seasonal, interannual and orbital variability during the Holocene. Nat. Geosci., 9, 168–173.CrossRefGoogle Scholar
Emile-Geay, J., et al. (2021). Past ENSO variability: Reconstructions, models, and implications. In McPhaden, M. J., et al. (eds.), El Niño Southern Oscillation in a Changing Climate. Geophysical Monograph 253, First Edition. American Geophysical Union, Washington, DC, pp. 87–118.Google Scholar
Falster, G., et al. (2023). Forced changes in the Pacific Walker Circulation over the past millennium. Nature, 622(7981), 93–100. https://doi.org/10.1038/s41586-023-06447-0.CrossRefGoogle ScholarPubMed
Feng, M., et al. (2021). Multi-year marine cold-spells off the west coast of Australia and effects on fisheries. J. Mar. Systems, 214, https://doi.org/10.1016/j.jmarsys.2020.103473.CrossRefGoogle Scholar
Fletcher, M.-S. and Moreno, P. I. (2011). Zonally symmetric changes in the strength and position of the Southern Westerlies drove atmospheric CO2 variations over the past 14 k.y. Geology, 39, 419–422. https://doi.org/10.1130/G31807.1.CrossRefGoogle Scholar
Fletcher, M.-S. and Moreno, P. I. (2012). Have the Southern Westerlies changed in a zonally symmetric manner over the last 14,000 years? A hemisphere-wide take on a controversial problem. Quat. Int., 253, 32–46. http://dx.doi.org/10.1016/j.quaint.2011.04.042.CrossRefGoogle Scholar
Fogt, R. L. and Bromwich, D. H. (2006). Decadal variability of the ENSO teleconnection to the high-latitude South Pacific governed by coupling with the southern annular mode. J. Clim., 19(6), 979–997.CrossRefGoogle Scholar
Fogt, R. L., et al. (2009). Historical SAM variability. Part II: Twentieth-century variability and trends from reconstructions, observations, and the IPCC AR4 models. J. Clim., 22, 5346–5365. https://doi.org/10.1175/2009JCLI2786.1.CrossRefGoogle Scholar
Fogt, R. L., et al. (2011). Understanding the SAM influence on the South Pacific ENSO teleconnection. Clim. Dyn., 36(7–8), 1555–1576. https://doi.org/10.1007/s00382-010-0905-0.CrossRefGoogle Scholar
Fogt, R. L. and Marshall, G. J. (2020). The Southern Annular Mode: Variability, trends, and climate impacts across the Southern Hemisphere. WIREs Clim Change., 11(4), e652. https://doi.org/10.1002/wcc.652.CrossRefGoogle Scholar
Franke, J., et al. (2017). Data descriptor: A monthly global paleo-reanalysis of the atmosphere from 1600 to 2005 for studying past climatic variations. Sci. Data, 4, 170076. https://doi.org/10.1038/sdata.2017.76.CrossRefGoogle Scholar
Frölicher, T. L., et al. (2018). Marine heatwaves under global warming. Nature, 560, 360–364.CrossRefGoogle ScholarPubMed
Gao, C., et al. (2008). Volcanic forcing of climate over the past 1500 years: An improved ice core- based index for climate models. J. Geophys. Res., 113(D23), https://doi.org/10.1029/2008JD010239.Google Scholar
Gergis, J. and Fowler, A. (2005). Classification of synchronous oceanic and atmospheric El Niño–Southern Oscillation (ENSO) events for palaeoclimate reconstruction. Int. J. Climatol., 25, 1541–1565.CrossRefGoogle Scholar
Gergis, J., et al. (2006). Reconstructing El Niño – Southern Oscillation (ENSO) from high-resolution palaeoarchives. J. Quat. Sci., 21, 707–722.CrossRefGoogle Scholar
Gergis, J. and Fowler, A. (2009). A history of El Niño–Southern Oscillation (ENSO) events since A.D. 1525: Implications for future climate change. Clim. Change, 92(3), 343–387.CrossRefGoogle Scholar
Glantz, M. (2006). Essay: Problem climates or problem societies? In Bridgeman, H. A. and Oliver, J. E. (eds.), The Global Climate System: Patterns, Processes, and Teleconnections. Cambridge University Press, Cambridge.Google Scholar
Goodwin, I. D., et al. (2004). Mid latitude winter climate variability in the south Indian and south-west Pacific regions since 1300 AD. Clim. Dyn., 22, 783–794. https://doi.org/10.1007/S00382-004-0403-3.CrossRefGoogle Scholar
Goodwin, I. D., et al. (2014). A reconstruction of extratropical Indo-Pacific sea-level pressure patterns during the Medieval Climate Anomaly. Clim. Dyn., 43(5–6), 1197–1219. https://doi.org/10.1007/s00382-013-1899-1.CrossRefGoogle Scholar
Goodwin, I. D., et al. (2023). Robbins Island: The index site for regional Last Interglacial sea level, wave climate and the subtropical ridge around Bass Strait, Australia. Quat. Sc. Rev., 305, 107996. https://doi.org/10.1016/j.quascirev.2023.107996.CrossRefGoogle Scholar
Goodwin, I. D. (in prep 2025). The South Pacific Split Jet and climate anomalies between the Subtropical Pacific and Antarctica during MIS5e.Google Scholar
Goosse, H., et al. (2012a). The role of forcing and internal dynamics in explaining the Medieval Climate Anomaly. Clim. Dyn., 39, 2847–2866. https://doi.org/10.1007/s00382-012-1297-0.CrossRefGoogle Scholar
Goosse, H., et al. (2012b). The medieval climate anomaly in Europe: Comparison of the summer and annual mean signals in two reconstructions and in simulations with data assimilation. Glob. Planet. Change, 84–85, 35–47. https://doi.org/10.1016/j.gloplacha.2011.07.002.Google Scholar
Graham, N. E. (1994). Decadal-scale climate variability in the tropical and North Pacific during the 1970s and 1980s: Observations and model results. Clim. Dyn., 10, 135–162.CrossRefGoogle Scholar
Grise, K. M., et al. (2018). Regional and seasonal characteristics of the recent expansion of the Tropics. J. Clim., 31, 6839–6856. https://doi.org/10.1175/JCLI-D-18-0060.1.CrossRefGoogle Scholar
Grise, K. M., et al. (2019). Recent tropical expansion: Natural variability or forced response? J. Clim., 32, 1551–1571. https://doi.org/10.1175/JCLI-D-18-0444.1.CrossRefGoogle Scholar
Guilyardi, E., et al. (2021). ENSO Modelling. In McPhaden, M. J., et al. (eds.), El Niño Southern Oscillation in a Changing Climate. Geophysical Monograph 253, First Edition. American Geophysical Union, pp. 199–226.Google Scholar
Hakim, G. J., et al. (2016). The Last Millennium Climate Reanalysis Project: Framework and first results. J. Geophys. Res. Atmos., 121, 6745–6764. https://doi.org/10.1002/2016JD024751.CrossRefGoogle Scholar
Hessl, A., et al. (2017). Reconstructions of the Southern Annular Mode (SAM) during the last millennium. Prog. Phys.Geogr., 41(6), 834–849.CrossRefGoogle Scholar
Hinojosa, J. L., et al. (2017). A New Zealand perspective on centennial-scale Southern Hemisphere westerly wind shifts during the last two millennia. Quat. Sci. Rev., 172, 32–43. http://dx.doi.org/10.1016/j.quascirev.2017.07.016.CrossRefGoogle Scholar
Huiskamp, W. and McGregor, S. (2021). Quantifying Southern Annular Mode paleo- reconstruction skill in a model framework. Clim. Past, 17(5), 1819–1839.CrossRefGoogle Scholar
Hurrell, J. W., et al. (2003). An overview of the North Atlantic Oscillation. In Hurrell, J. W., et al. (eds.), The North Atlantic Oscillation, Climatic Significance and Environmental Impact. AGU Geophysical Monograph, vol. 134, American Geophysical Union, Washington, DC, pp. 1–35.CrossRefGoogle Scholar
Jones, P. D., et al. (2009). High-resolution palaeoclimatology of the last millennium: A review of current status and future prospects. Holocene, 19(1), 3–49.CrossRefGoogle Scholar
Josey, S. A., et al. (2018). The recent Atlantic cold anomaly: Causes, consequences, and related phenomena. Annu. Rev. Mar. Sci., 10, 475–501. https://doi.org/10.1146/annurev-marine-121916-063102.CrossRefGoogle ScholarPubMed
Khodri, M., et al. (2017). Tropical explosive volcanic eruptions can trigger El Niño by cooling tropical Africa. Nat. Commun., 8(1), 778. https://doi.org/10.1038/s41467-017-00755-6.Google ScholarPubMed
King, J., et al. (2023). Trends and variability in the Southern Annular Mode over the Common Era. Nat. Commun., 14(1), 2324–2324. https://doi.org/10.1038/s41467-023-37643-1.CrossRefGoogle ScholarPubMed
Klein, F., et al. (2019). Assessing the robustness of Antarctic temperature reconstructions over the past two millennia using pseudoproxy and data assimilation experiments. Clim. Past, 15, 661–684.CrossRefGoogle Scholar
Konecky, B. L., et al. (2020). The Iso2k database: A global compilation of paleo-δ18O and δ2H records to aid understanding of common era climate. Earth Sys. Sci. Data, 12(3), 2261–2288. https://doi.org/10.5194/essd-12-2261-2020.CrossRefGoogle Scholar
Kreutz, K. J., et al. (2000). Sea level pressure variability in the Amundsen Sea region inferred from a West Antarctic glaciochemical record. J. Geophys. Res., 105(D3), 4047–4059.Google Scholar
Lamy, F., et al. (2010). Holocene changes in the position and intensity of the southern westerly wind belt. Nat. Geosci., 3, 695–699.CrossRefGoogle Scholar
Li, J., et al. (2017). Quantifying climatic variability in monsoonal northern China over the last 2200 years and its role in driving Chinese dynastic changes. Quat. Sci. Rev., 159, 35–46. https://doi.org/10.1016/j.quascirev.2017.01.009.CrossRefGoogle Scholar
Linsley, B. K., et al. (2015). Decadal changes in South Pacific sea surface temperatures and the relationship to the Pacific decadal oscillation and upper ocean heat content. Geophys. Res. Lett., 42, 2358–2366. https://doi.org/10.1002/2015GL063045.CrossRefGoogle Scholar
Luterbacher, J., et al. (2002). Reconstruction of sea level pressure fields over the Eastern North Atlantic and Europe back to 1500. Clim. Dyn., 18, 545–561.CrossRefGoogle Scholar
Luterbacher, J., et al. (2016). European summer temperatures since Roman times. Environ. Res. Lett., 11, 024001. https://doi.org/10.1088/1748-9326/11/2/024001.CrossRefGoogle Scholar
Mackintosh, N. A. and Herdman, H. F. P. (1940). Distribution of the pack-ice in the Southern Ocean. Discovery Rep., 19, 285–296.Google Scholar
Mann, M. E., et al. (2009). Global signatures and dynamical origins of the Little Ice Age and Medieval Climate Anomaly. Science, 326, 1256–1260. https://doi.org/10.1126/science.1177303.CrossRefGoogle ScholarPubMed
Marshall, G. J. (2003). Trends in the Southern Annular Mode from observations and reanalyses. J. Clim., 16, 4134–4143.2.0.CO;2>CrossRefGoogle Scholar
Masson-Delmotte, V., et al. (2013). Information from paleoclimate archives. In Stocker, T. F., et al., (eds.), Climate Change 2013 – The Physical Science Basis: Working Group I Contribution to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, pp. 383–464.Google Scholar
McCulloch, M. T. and Esat, T. (2000). The coral record of Last Interglacial sea levels and sea surface temperatures. Chem. Geol., 169, 107–129.CrossRefGoogle Scholar
McGregor, S., et al. (2013). Inferred changes in El Niño–Southern Oscillation variance over the past six centuries. Clim. Past, 9(5), 2269–2284. https://doi.org/10.5194/cp-9-2269-2013.CrossRefGoogle Scholar
McGregor, H., et al. (2016). Data, age uncertainties and ocean δ18O under the spotlight for Ocean2k Phase 2. Past Glob. Changes Mag., 24(1), 44. https://doi.org/10.22498/pages.24.1.44.Google Scholar
McKay, R. C., et al. (2023). Can southern Australian rainfall decline be explained? A review of possible drivers. WIREs Clim. Change, e820, 1–24. https://doi.org/10.1002/wcc.820.Google Scholar
McPhaden, M. J., et al. (eds.), (2021). El Niño Southern Oscillation in a Changing Climate. Geophysical Monograph 253, First Edition. American Geophysical Union, Washington, DC, pp. 199–226.Google Scholar
Menviel, L., et al. (2011). Deconstructing the Last Glacial termination: The role of millennial and orbital-scale forcings. Quat. Sci. Rev., 30, 1155–1172. https://doi.org/10.1016/j.quascirev.2011.02.005.CrossRefGoogle Scholar
Michel, S., et al. (2020). Reconstructing climatic modes of variability from proxy records using ClimIndRec version 1.0. Geosci. Model Dev., 13, 841–858. https://doi.org/10.5194/gmd-13-841-2020.CrossRefGoogle Scholar
Miller, A. J., et al. (1994). The 1976–77 climate shift of the Pacific Ocean. Oceanography, 7(1), 21–26. http://dx.doi.org/10.5670/oceanog.1994.11.CrossRefGoogle Scholar
Mindlin, J., et al. (2020). Storyline description of Southern Hemisphere midlatitude circulation and precipitation response to greenhouse gas forcing. Clim. Dyn., 54, 4399−4421. https://doi.org/10.1007/s00382-020-05234-1.CrossRefGoogle ScholarPubMed
Moreno, P. I., et al. (2018). Onset and evolution of Southern Annular Mode-like changes at centennial timescale. Sci. Rep., 8, 3458.CrossRefGoogle ScholarPubMed
Moy, C. M., et al. (2002). Variability of El Niño southern oscillation activity at millennial timescales during the Holocene epoch. Nature, 420, 162–165. http://dx.doi.org/10.1038/nature01194.CrossRefGoogle ScholarPubMed
Moy, C. M., et al. (2008). Isotopic evidence for hydrologic change related to the westerlies in SW Patagonia, Chile, during the last millennium. Quat. Sci. Rev., 27, 1335–1349.CrossRefGoogle Scholar
Moy, C. M., et al. (2009). Climate change in southern South America during the last two millennia. In Vimeux, F., et al. (eds.), Past Climate Variability in South America and Surrounding Regions. Springer, New York, pp. 353–393. http://dx.doi.org/10.1007/978-90-481-2672-9_15.Google Scholar
Neukom, R., et al. (2014). Inter-hemispheric temperature variability over the past millennium, Nat. Clim. Change, 4, 362–367. https://doi.org/10.1038/nclimate2174.CrossRefGoogle Scholar
Neukom, R., et al. (2015). Facing unprecedented drying of the Central Andes? Precipitation variability over the period AD 1000–2100. Environ. Res. Lett., 10, 084017. https://doi.org/10.1088/1748-9326/10/8/084017.CrossRefGoogle Scholar
Neukom, R., et al. (2019a). Consistent multidecadal variability in global temperature reconstructions and simulations over the Common Era. Nat. Geosci., 536, 411. https://doi.org/10.1038/s41561-019-0400-0.Google Scholar
Neukom, R., et al. (2019b). No evidence for globally coherent warm and cold periods over the preindustrial Common Era. Nature, 571, 550–554. https://doi.org/10.1038/s41586-019-1401-2.CrossRefGoogle Scholar
Oliver, E. C. J., et al. (2018). Longer and more frequent marine heatwaves over the past century. Nat. Commun., 9, 1324.CrossRefGoogle ScholarPubMed
Oliver, E. C. J., et al. (2021). Marine heatwaves. Annu. Rev. Mar. Sci., 13, 313–342.CrossRefGoogle ScholarPubMed
Ortega, P., et al. (2015). A model-tested North Atlantic Oscillation reconstruction for the past millennium, Nature, 523, 71. https://doi.org/10.1038/nature14518.CrossRefGoogle ScholarPubMed
Otto-Bliesner, B. L., et al. (2021). Large-scale features of Last Interglacial climate: Results from evaluating the lig127k simulations for the Coupled Model Intercomparison Project (CMIP6)–Paleoclimate Modeling Intercomparison Project (PMIP4). Clim. Past, 17, 63–94. https://doi.org/10.5194/cp-17-63-2021.CrossRefGoogle Scholar
PAGES 2k Consortium. (2013). Continental-scale temperature variability during the past two millennia. Nat. Geosci., 6, 5. https://doi.org/10.1038/ngeo1797.Google Scholar
PAGES 2k-PMIP3 group. (2015). Continental-scale temperature variability in PMIP3 simulations and PAGES 2k regional temperature reconstructions over the past millennium. Clim. Past, 11, 1673–1699. https://doi.org/10.5194/cp-11-1673-2015.Google Scholar
PAGES Hydro 2k Consortium. (2017). Comparing proxy and model estimates of hydroclimate variability and change over the Common Era. Clim. Past, 13, 1851–1900. https://doi.org/10.5194/cp-13-1851-2017.Google Scholar
PAGES 2k Consortium. (2017a). A global multiproxy database for temperature reconstructions of the Common Era. Sci. Data, 4(170088), 1–33. https://doi.org/10.1038/sdata.2017.88.Google Scholar
PAGES 2k Consortium. (2017b). PAGES 2k Global 2,000 Year Multiproxy Database. Available at: www.ncdc.noaa.gov/paleo/study/21171.Google Scholar
Partin, J. W., et al. (2015). Iso2k: A community-driven effort to develop a global database of paleo- water isotopes covering the past two millennia. Eos Trans. AGU, Fall Meet. Suppl., 2015.Google Scholar
Pausata, F. S., et al. (2020). ITCZ shift and extratropical teleconnections drive ENSO response to volcanic eruptions. Sci. Adv., 6(23), eaaz5006. https://doi.org/10.1126/sciadv.aaz5006.CrossRefGoogle ScholarPubMed
Predybaylo, E., et al. (2020). El Niño/Southern Oscillation response to low-latitude volcanic eruptions depends on ocean pre-conditions and eruption timing. Commun. Earth Environ., 1–12. https://doi.org/10.1038/s43247-020-0013-y.CrossRefGoogle Scholar
Priestley, M. D. K. and Catto, J. L. (2022). Future changes in the extratropical storm tracks and cyclone intensity, wind speed, and structure. Weather Clim. Dyn. Discuss., 3(1), 337–360. https://doi.org/10.5194/wcd-2021-75.Google Scholar
Rodrigues, R. R., et al. (2019). Common cause for severe droughts in South America and marine heatwaves in the South Atlantic. Nat. Geosci., 12, 620.CrossRefGoogle Scholar
Rypdal, M. and Rypdal, K. (2016). Late Quaternary temperature variability described as abrupt transitions on a 1∕f noise background. Earth Syst. Dynam., 7(1), 281–293.CrossRefGoogle Scholar
Sachs, J. P., et al. (2009). Southward movement of the Pacific intertropical convergence zone AD 1400–1850. Nat. Geosci., 2, 519–525.CrossRefGoogle Scholar
Schmidt, G. A., et al. (2014). Using palaeo-climate comparisons to constrain future projections in CMIP5. Clim. Past, 10, 221–250. https://doi.org/10.5194/cp-10-221-2014.CrossRefGoogle Scholar
Screen, J. A. and Simmonds, I. (2014). Amplified mid-latitude planetary waves favour particular regional weather extremes. Nat. Clim. Change, 4, 704–709.CrossRefGoogle Scholar
Screen, J. A., et al. (2018). Polar climate change as manifest in atmospheric circulation. Curr. Clim. Change Rep., 4, 383–395. https://doi.org/10.1007/s40641-018-0111-4.CrossRefGoogle ScholarPubMed
Seager, R., et al. (2019). Climate variability and change of Mediterranean-type climates. J. Clim., 32, 2887–2915.CrossRefGoogle Scholar
Sen Gupta, A., et al. (2020). Drivers and impacts of the most extreme marine heatwave events. Sci Rep, 10, 19359. https://doi.org/10.1038/s41598-020-75445-3.CrossRefGoogle Scholar
Senapati, B., et al. (2021). Global wave number-4 pattern in the southern subtropical sea surface temperature. Sci. Rep., 11, 142.CrossRefGoogle ScholarPubMed
Shao, Z.-G. and Ditlevsen, P. D. (2016). Contrasting scaling properties of interglacial and glacial climates. Nature Comm., 7, 10951. https://doi.org/10.1038/ncomms10951.CrossRefGoogle ScholarPubMed
Shaw, T. A., et al. (2016). Storm track processes and the opposing influences of climate change. Nat. Geosci., 9, 656–664. https://doi.org/10.1038/ngeo2783.CrossRefGoogle Scholar
Shepherd, T. G. (2014). Atmospheric circulation as a source of uncertainty in climate change projections. Nat. Geosci., 7, 703–708. https://doi.org/10.1038/ngeo2253.CrossRefGoogle Scholar
Shepherd, T. G., et al. (2018). Storylines: An alternative approach to representing uncertainty in physical aspects of climate change. Clim. Change, 151, 555–571.CrossRefGoogle ScholarPubMed
Shepherd, T. G. (2019). Storyline approach to the construction of regional climate change information. Proc R Soc A., 475, 20190013.CrossRefGoogle Scholar
Shepherd, T. G. (2021). Bringing physical reasoning into statistical practice in climate-change science. Clim. Change, 169(1–2), 2. https://doi.org/10.1007/s10584-021-03226-6.CrossRefGoogle Scholar
Shepherd, T. G. and Lloyd, E. A. (2021). Meaningful climate science. Clim. Change, 169, 17. https://doi.org/10.1007/s10584-021-03246-2.CrossRefGoogle Scholar
Smerdon, J. E. and Pollack, H. N. (2016). Reconstructing Earth’s surface temperature over the past 2000 years: The science behind the headlines. WIREs Clim. Change, 7, 746–771. https://doi.org/10.1002/wcc.418.CrossRefGoogle Scholar
Stahle, D. W. (1998). Experimental dendroclimatic reconstruction of the Southern Oscillation. Bull. Am. Meteorol. Soc., 79(10), 2137–2152.2.0.CO;2>CrossRefGoogle Scholar
Staten, P. W., et al. (2020). Tropical widening: From global variations to regional impacts. Bull. Am. Meteorol. Soc., 101, E897–E904. https://doi.org/10.1175/BAMS-D-19-0047.1.CrossRefGoogle Scholar
Steiger, N. J., et al. (2018). A reconstruction of global hydroclimate and dynamical variables over the Common Era. Sci. Data, 5(1), 180086.CrossRefGoogle ScholarPubMed
Stenni, B., et al. (2017). Antarctic climate variability on regional and continental scales over the last 2000 years. Clim. Past, 13, 1609–1634. https://doi.org/10.5194/cp-13-1609-2017.CrossRefGoogle Scholar
Tardiff, R., et al. (2019). Last Millennium Reanalysis with an expanded proxy database and seasonal proxy modeling. Clim. Past, 15, 1251–1273. https://doi.org/10.5194/cp-15-1251-2019.Google Scholar
Teleti, P. R., et al. (2019). A historical Southern Ocean climate dataset from whaling ships’ logbooks. Geosci. Data J., 6, 30–40. https://doi.org/10.1002/gdj3.65.CrossRefGoogle Scholar
Terray, L. (2023). A storyline approach to the June 2021 northwestern North American heatwave. Geophys. Res. Lett., 50, e2022GL101640. https://doi.org/10.1029/2022GL101640.CrossRefGoogle Scholar
Thomas, E. R., et al. (2019). Antarctic sea ice proxies from marine and ice core archives suitable for reconstructing sea ice over the past 2000 years. Geosciences, 9(12), 506. https://doi.org/10.3390/geosciences9120506.CrossRefGoogle Scholar
Tierney, J. E., et al. (2010). Coordinated hydrological regimes in the Indo-Pacific region during the past two millennia. Paleoceanog., 25(1), PA1102. https://doi.org/10.1029/2009PA001871.CrossRefGoogle Scholar
Tierney, J. E., et al. (2015). Tropical sea surface temperatures for the past four centuries reconstructed from coral archives. Paleoceanog., 30, 226–252. https://doi.org/10.1002/2014PA002717.CrossRefGoogle Scholar
Timmermann, A., et al. (2018). El Niño–Southern Oscillation complexity. Nature, 559, 535–545. https://doi.org/10.1038/s41586-018-0252-6.CrossRefGoogle ScholarPubMed
Trewartha, G. T. (1961). The Earth’s Problem Climates. The University of Wisconsin Press, Madison, WI.Google Scholar
Trouet, V., et al. (2009). Persistent positive North Atlantic Oscillation mode dominated the Medieval Climate Anomaly. Science, 324, 78–80. https://doi.org/10.1126/science.1166349.CrossRefGoogle ScholarPubMed
Tselioudis, G., et al. (2016). Midlatitude cloud shifts, their primary link to the Hadley cell, and their diverse radiative effects. Geophys. Res. Lett., 43(9), 4594–4601. https://doi.org/10.1002/2016GL068242.CrossRefGoogle Scholar
Tselioudis, G., et al. (2024). Oceanic cloud trends during the satellite era and their radiative signatures. Clim. Dyn., 62, 9319–9332. https://doi.org/10.1007/s00382-024-07396-8.CrossRefGoogle Scholar
Tsonis, A. A. (2018). Insights in climate dynamics from climate networks. In Advances in Nonlinear Geosciences. Springer International Publishing, pp. 631–649. https://doi.org/10.1007/978-3-319-58895-7_29.CrossRefGoogle Scholar
Tsonis, A. A., et al. (2007). A new dynamical mechanism for major climate shifts. Geophys. Res. Lett., 34, L13705. https://doi.org/10.1029/2007GL030288.CrossRefGoogle Scholar
Tudhope, A. W., et al. (1995). Recent changes in climate in the far western equatorial Pacific and their relation- ship to the Southern Oscillation: Oxygen isotope records from massive corals, Papua New Guinea. Earth Planet. Sci. Lett., 136, 575–590. https://doi.org/10.1016/0012-821X(95)00156-7.CrossRefGoogle Scholar
Tudhope, A. W., et al. (2001). Variability in the El Niño– Southern Oscillation through a glacial-interglacial cycle. Science, 291, 1511–1517.CrossRefGoogle ScholarPubMed
Turney, C., et al. (2016). A 250 year periodicity in Southern Hemisphere westerly winds over the last 2600 years. Clim. Past., 12, 189–200.CrossRefGoogle Scholar
Turney, C. J., et al. (2017). Reconstructing atmospheric circulation over southern New Zealand: Establishment of modern westerly airflow 5500 years ago and implications for Southern Hemisphere Holocene climate change. Quat. Sci. Rev., 159, 77–87.CrossRefGoogle Scholar
Urban, F. E., et al. (2000). Influence of mean climate change on climate variability from a 155-year tropical Pacific coral record. Nature, 407(6807), 989–993.CrossRefGoogle ScholarPubMed
Valler, V., et al. (2020). Assimilating monthly precipitation data in a paleoclimate data assimilation framework. Clim. Past, 16, 1309–1323. https://doi.org/10.5194/cp-16-1309-2020.CrossRefGoogle Scholar
Valler, V., et al. (2021). An updated global atmospheric paleo-reanalysis covering the last 400 years. Geosci. Data J., 9, 89–107. https://doi.org/10.1002/gdj3.121.Google ScholarPubMed
van Garderen, L. and Mindlin, J. (2022). A storyline attribution of the 2011/2012 drought in Southeastern South America. Weather, 77(6), 212–218.CrossRefGoogle Scholar
van der Schrier, G. and Barkmeijer, J. (2005). Bjerknes’ hypothesis on the coldness during AD 1790–1820 revisited. Clim. Dyn., 25, 537–553.CrossRefGoogle Scholar
Villalba, R., et al. (2012). Unusual Southern Hemisphere tree growth patterns induced by changes in the Southern Annular Mode. Nature Geosci., 5, 793–798.CrossRefGoogle Scholar
von Humboldt, A. (1850). Views of Nature. Accessible from www.gutenberg.org/files/67684/67684-h/67684-h.htm.Google Scholar
Von Storch, H., et al. (2018). The concept of large-scale conditioning of climate model simulations of atmospheric coastal dynamics: Current state and perspectives. Atmosphere, 9, 337. https://doi.org/10.3390/atmos9090337.CrossRefGoogle Scholar
Wehrli, K., et al. (2019). Identifying key driving processes of major recent heat waves. J. Geophys. Res.: Atmos., 124, 11746–11765. https://doi.org/10.1029/2019JD030635.CrossRefGoogle Scholar
Wehrli, K., et al. (2020). Storylines of the 2018 Northern Hemisphere heatwave at pre-industrial and higher global warming levels. Earth Syst. Dynam., 11, 855–873. https://doi.org/10.5194/esd-11-855-2020.CrossRefGoogle Scholar
Wilkinson, C. and Wilkinson, S. (2018). Report on the Imaging of Sources of Historic Ice, Meteorological and Oceanographic Data in the Southern Ocean. Archive of the Sea Mammal Research Unit, University of St. Andrews, Scotland, Recovery of Logbooks and International Marine Data, RECLAIM Project, 2018.Google Scholar
Yan, H., et al. (2015). Dynamics of the intertropical convergence zone over the western Pacific during the Little Ice Age. Nat. Geosci., 8, 315–320.CrossRefGoogle Scholar
Zinke, J., et al. (2014). Corals record long-term Leeuwin Current variability during Ningaloo Niño/Niña since 1795. Nat. Commun., 5(1), 3607. https://doi.org/10.1038/ncomms4607.CrossRefGoogle ScholarPubMed
Zinke, J., et al. (2021). The West Pacific Gradient tracks ENSO and zonal Pacific sea surface temperature gradient during the last Millennium. Sci. Rep., 11, 20395. https://doi.org/10.1038/s41598-021-99738-3.CrossRefGoogle ScholarPubMed

Accessibility standard: WCAG 2.0 A

Why this information is here

This section outlines the accessibility features of this content - including support for screen readers, full keyboard navigation and high-contrast display options. This may not be relevant for you.

Accessibility Information

The PDF of this book conforms to version 2.0 of the Web Content Accessibility Guidelines (WCAG), ensuring core accessibility principles are addressed and meets the basic (A) level of WCAG compliance, addressing essential accessibility barriers.

Content Navigation

Table of contents navigation
Allows you to navigate directly to chapters, sections, or non‐text items through a linked table of contents, reducing the need for extensive scrolling.
Index navigation
Provides an interactive index, letting you go straight to where a term or subject appears in the text without manual searching.

Reading Order & Textual Equivalents

Single logical reading order
You will encounter all content (including footnotes, captions, etc.) in a clear, sequential flow, making it easier to follow with assistive tools like screen readers.
Short alternative textual descriptions
You get concise descriptions (for images, charts, or media clips), ensuring you do not miss crucial information when visual or audio elements are not accessible.
Full alternative textual descriptions
You get more than just short alt text: you have comprehensive text equivalents, transcripts, captions, or audio descriptions for substantial non‐text content, which is especially helpful for complex visuals or multimedia.

Structural and Technical Features

ARIA roles provided
You gain clarity from ARIA (Accessible Rich Internet Applications) roles and attributes, as they help assistive technologies interpret how each part of the content functions.

Save book to Kindle

To save this book to your Kindle, first ensure no-reply@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • References
  • Ian D. Goodwin, Macquarie University and ClimaLab
  • Book: Synoptic Paleoclimatology
  • Online publication: 09 September 2025
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • References
  • Ian D. Goodwin, Macquarie University and ClimaLab
  • Book: Synoptic Paleoclimatology
  • Online publication: 09 September 2025
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • References
  • Ian D. Goodwin, Macquarie University and ClimaLab
  • Book: Synoptic Paleoclimatology
  • Online publication: 09 September 2025
Available formats
×