Hostname: page-component-8448b6f56d-gtxcr Total loading time: 0 Render date: 2024-04-18T23:22:08.683Z Has data issue: false hasContentIssue false

The Atlantic Meridional Overturning Circulation as productivity regulator of the North Atlantic Subtropical Gyre

Published online by Cambridge University Press:  14 November 2018

Sílvia Nave*
Affiliation:
Laboratório Nacional de Energia e Geologia, I.P. (LNEG), Estrada da Portela, Apt. 7586. 2610-999 Amadora, Portugal
Susana Lebreiro
Affiliation:
Instituto Geológico y Minero de España, Calle Ríos Rosas 23, 28003 Madrid, Spain
Elisabeth Michel
Affiliation:
Laboratoire des Sciences du Climat et de l’Environnement, IPSL, CEA/CNRS/UVSQ, Université Paris-Saclay, Avenue de la Terrasse, 91198 Gif-sur-Yvette Cedex, France
Catherine Kissel
Affiliation:
Laboratoire des Sciences du Climat et de l’Environnement, IPSL, CEA/CNRS/UVSQ, Université Paris-Saclay, Avenue de la Terrasse, 91198 Gif-sur-Yvette Cedex, France
Maria Ondina Figueiredo
Affiliation:
Laboratório Nacional de Energia e Geologia, I.P. (LNEG), Estrada da Portela, Apt. 7586. 2610-999 Amadora, Portugal
Abel Guihou
Affiliation:
Laboratoire des Sciences du Climat et de l’Environnement, IPSL, CEA/CNRS/UVSQ, Université Paris-Saclay, Avenue de la Terrasse, 91198 Gif-sur-Yvette Cedex, France
António Ferreira
Affiliation:
British Geological Survey, Environmental Science Centre, Keyworth, Nottingham NG12 5GG, United Kingdom
Laurent Labeyrie
Affiliation:
Laboratoire des Sciences du Climat et de l’Environnement, IPSL, CEA/CNRS/UVSQ, Université Paris-Saclay, Avenue de la Terrasse, 91198 Gif-sur-Yvette Cedex, France
Ana Alberto
Affiliation:
Laboratório Nacional de Energia e Geologia, I.P. (LNEG), Estrada da Portela, Apt. 7586. 2610-999 Amadora, Portugal
*
*Corresponding author at: Laboratório Nacional de Energia e Geologia, I.P. (LNEG), Estrada da Portela, Apt. 7586, 2610-999 Amadora, Portugal. E-mail address: silvia.nave@lneg.pt (S. Nave)

Abstract

Spatially extensive and intense phytoplankton blooms observed off Iberia, in satellite pictures, are driven by significant nutrient supply by upper-ocean vertical mesoscale activity rather than by horizontal advection by coastal upwelling. Productivity of oligotrophic regions is still poorly depicted by discrete instrumental and model data sets. The paleoproductivity reconstructions of these areas represent the mean productivity over long periods, bringing new insights into the total biomass fluxes. Here, we present paleoproductivity records from the oceanic Tore Seamount region, covering the period from 140 to 60 ka. They show higher nutrient supplies during Termination II, Marine Oxygen Isotope Stage (MIS) 4, MIS 6, and warming transitions of the MIS 5 sub-stages. The highest nutrient content (higher productivity) in phase with tracers of bottom-water ventilation (benthic δ13C,231Pa/230Th) establishes a strong linkage with variability of Southern Ocean-sourced waters. Low productivity and ventilation over warm sub-stages of MIS 5 respond instead to North Atlantic Deep Water. Assuming that the Tore Seamount is representative of oligotrophic regions, the glacial-interglacial relationship observed between paleoproductivity and Atlantic Meridional Overturning Circulation strength opens new insights into the importance of estimating the total biomass in these regions. The subtropical gyres might play a considerable role in the carbon cycle over (sub-)glacial-interglacial time scales than previously thought.

Type
Research Article
Copyright
Copyright © University of Washington. Published by Cambridge University Press, 2018 

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.)

References

REFERENCES

Abrantes, F., 1991. Increased upwelling off Portugal during the last glaciation: diatom evidence. Marine Micropaleontology 17, 285310.Google Scholar
Abreu, L.d., Shackleton, N.J., Schönfeld, J., Hall, M., Chapman, M., 2003. Millennial-scale oceanic climate variability of the Western Iberian margin during the last two glacial periods. Marine Geology 196, 120.Google Scholar
Adkins, J.F., Ingersoll, A.P., Pasquero, C., 2005. Rapid climate change and conditional instability of the glacial deep ocean from the thermobaric effect and geothermal heating. Quaternary Science Reviews 24, 581594.Google Scholar
Arístegui, J., Álvarez-Salgado, X.A., Barton, E.D., Figueiras, F.G., Hernández-León, S., Roy, C., Santos, A.M.P., 2004. Oceanography and Fisheries of the Canary current/Iberian region of the Eastern North Atlantic. In: Robinson, A.R., Brink, K.H. (Eds.), The Sea. Harvard University Press, Cambridge, pp. 877931.Google Scholar
Bazin, L., Landais, A., Lemieux-Dudon, B., Toyé Mahamadou Kele, H., Veres, D., Parrenin, F., Martinerie, P., et al., 2013. An optimized multi-proxy, multi-site Antarctic ice and gas orbital chronology (AICC2012): 120–800 ka. Climate of the Past 9, 17151731.Google Scholar
Behrenfeld, M.J., O/‘Malley, R.T., Siegel, D.A., McClain, C.R., Sarmiento, J.L., Feldman, G.C., Milligan, A.J., Falkowski, P.G., Letelier, R.M., Boss, E.S., 2006. Climate-driven trends in contemporary ocean productivity. Nature 444, 752755.Google Scholar
Bianchi, G.G., Hall, I.R., McCave, I.N., Joseph, L., 1999. Measurement of the sortable silt current speed proxy using the Sedigraph 5100 and Coulter Multisizer IIe: precision and accuracy. Sedimentology 46, 10011014.Google Scholar
Brachfeld, S., Kissel, C., Laj, C., Mazaud, A., 2004. Viscous behavior of U channels during acquisition and demagnetization of remanences: implications for paleomagnetic and rock‐magnetic investigations. Physics of the Earth and Planetary Interiors 145. http://dx.doi.org/10.1016/j.pepi.2003.1012.1011.Google Scholar
Bradtmiller, L.I., McGee, D., Awalt, M., Evers, J., Yerxa, H., Kinsley, C.W., deMenocal, P.B., 2016. Changes in biological productivity along the northwest African margin over the past 20,000 years. Paleoceanography 31, 185202.Google Scholar
Broecker, W.S., 1982. Glacial to interglacial changes in ocean chemistry. Progress In Oceanography 11, 151197.Google Scholar
Broecker, W.S. 1982b. Ocean chemistry during glacial time. Geochimica et Cosmochimica Acta 46, 16891705.Google Scholar
Cayre, O., Lancelot, Y., Vincent, E., Hall, M.A., 1999. Paleoceanographic reconstructions from planktonic foraminifera off the Iberian Margin: temperature, salinity, and Heinrich events. Paleoceanography 14, 384396.Google Scholar
Channell, J.E.T., Hodell, D.A., Margari, V., Skinner, L.C., Tzedakis, P.C., Kesler, M.S., 2013. Biogenic magnetite, detrital hematite, and relative paleointensity in Quaternary sediments from the Southwest Iberian Margin. Earth and Planetary Science Letters 376, 99109.Google Scholar
Coplen, T.B., 1988. Normalization of oxygen and hydrogen isotope data. Chemical Geology 72, 293297.Google Scholar
Curry, W.B., Oppo, D.W., 2005. Glacial water mass geometry and the distribution of δ13C of ΣCO2 in the western Atlantic Ocean. Paleoceanography 20, PA1017. http://dx.doi.org/10.1029/2004PA001021.Google Scholar
Dufois, F., Hardman-Mountford, N.J., Greenwood, J., Richardson, A.J., Feng, M., Matear, R.J., 2016. Anticyclonic eddies are more productive than cyclonic eddies in subtropical gyres because of winter mixing. Science Advances 2, e1600282.Google Scholar
Eberwein, A., Mackensen, A., 2006. Regional primary productivity differences off Morocco (NW-Africa) recorded by modern benthic foraminifera and their stable carbon isotopic composition. Deep Sea Research Part I: Oceanographic Research Papers 53, 13791405.Google Scholar
Emerson, S., Quay, P., Karl, D., Winn, C., Tupas, L., Landry, M., 1997. Experimental determination of the organic carbon flux from open-ocean surface waters. Nature 389, 951954.Google Scholar
Farr, T.G., Rosen, P.A., Caro, E., Crippen, R., Duren, R., Hensley, S., Kobrick, M., et al., 2007. The Shuttle Radar Topography Mission. Reviews of Geophysics 45, RG2004. http://dx.doi.org/10.1029/2005RG000183.Google Scholar
Filippelli, G.M., Latimer, J.C., Murray, R.W., Flores, J.-A., 2007. Productivity records from the Southern Ocean and the equatorial Pacific Ocean: testing the glacial Shelf-Nutrient Hypothesis. Deep Sea Research Part II: Topical Studies in Oceanography 54, 24432452.Google Scholar
Fiúza, A.F.G., 1983. Upwelling patterns off Portugal. In: Suess, E., Thiede, J. (Eds.), Coastal Upwelling Its Sediment Record. Plenum Press, New York, pp. 8598.Google Scholar
Fiúza, A.F.G., 1984. Hidrologia e dinamica das aguas costeiras de Portugal. PhD dissertation, Faculdade de Ciências da Universidade de Lisboa, Universidade de Lisboa, Lisboa, Portugal.Google Scholar
François, R., Frank, M., van der Loeff, M.M.R., Bacon, M.P., 2004. 230Th normalization: an essential tool for interpreting sedimentary fluxes during the late Quaternary. Paleoceanography 19, PA1018. http://dx.doi.org/1010.1029/2003PA000939.Google Scholar
Gil, I.M., Keigwin, L.D., Abrantes, F.G., 2009. Deglacial diatom productivity and surface ocean properties over the Bermuda Rise, northeast Sargasso Sea. Paleoceanography 24, PA4101. http://dx.doi.org/10.1029/2008PA001729.Google Scholar
Govin, A., Chiessi, C.M., Zabel, M., Sawakuchi, A.O., Heslop, D., Hörner, T., Zhang, Y., Mulitza, S., 2014. Terrigenous input off northern South America driven by changes in Amazonian climate and the North Brazil Current retroflection during the last 250 ka. Climate of the Past 10, 843862.Google Scholar
Guihou, A., Pichat, S., Govin, A., Nave, S., Michel, E., Duplessy, J.-C., Telouk, P., Labeyrie, L., 2011. Enhanced Atlantic Meridional Overturning Circulation supports the Last Glacial Inception. Quaternary Science Reviews 30, 15761582.Google Scholar
Guihou, A., Pichat, S., Nave, S., Govin, A., Labeyrie, L., Michel, E., Waelbroeck, C., 2010. Late slowdown of the Atlantic Meridional Overturning Circulation during the Last Glacial Inception: new constraints from sedimentary (231Pa/230Th). Earth and Planetary Science Letters 289, 520529.Google Scholar
Hall, I.R., McCave, I.N., 2000. Palaeocurrent reconstruction, sediment and thorium focussing on the Iberian margin over the last 140 ka. Earth and Planetary Science Letters 178, 151164.Google Scholar
Hendry, K.R., Gong, X., Knorr, G., Pike, J., Hall, I.R., 2016. Deglacial diatom production in the tropical North Atlantic driven by enhanced silicic acid supply. Earth and Planetary Science Letters 438, 122129.Google Scholar
Incarbona, A., Martrat, B., Stefano, E.D., Grimalt, J.O., Pelosi, N., Patti, B., Tranchida, G., 2010. Primary productivity variability on the Atlantic Iberian Margin over the last 70,000 years: evidence from coccolithophores and fossil organic compounds. Paleoceanography 25, PA2218. http://dx.doi.org/10.1029/2008PA001709.PA2218.Google Scholar
Jennings, S., Kaiser, M.J., Reynolds, J.D., 2001. Marine Fisheries Ecology. Blackwell, Oxford.Google Scholar
King, G.C.P., Banerjee, S.K., Marvin, J., Özdemir, Ö., 1982. A comparison of different magnetic methods for determining the relative grain size of magnetite in natural materials: some results from lake sediments. Earth and Planetary Science Letters 59, 404419.Google Scholar
Kissel, C., 2005. Magnetic signature of rapid climatic variations in glacial North Atlantic, a review. Comptes Rendus Geosciences 337, 908918.Google Scholar
Kukla, G.J., Bender, M.L., de Beaulieu, J.-L., Bond, G., Broecker, W.S., Cleveringa, P., Gavin, J.E., et al., 2002. Last interglacial climates. Quaternary Research 58, 213.Google Scholar
Lebreiro, S.M., Francés, G., Abrantes, F.F.G., Diz, P., Bartels-Jónsdóttir, H.B., Stroynowski, Z.N., Gil, I.M., et al., 2006. Climate change and coastal hydrographic response along the Atlantic Iberian margin (Tagus Prodelta and Muros Ría) during the last two millennia. The Holocene 16, 10031015.Google Scholar
Lebreiro, S.M., Moreno, J.C., Abrantes, F.F., Pflaumann, U., 1997. Productivity and paleoceanographic implications on the Tore Seamount (Iberian Margin) during the last 225 Kyr: foraminiferal evidence. Paleoceanography 12, 718727.Google Scholar
Lebreiro, S.M., Voelker, A.H.L., Vizcaino, A., Abrantes, F.G., Alt-Epping, U., Jung, S., Thouveny, N., Gràcia, E., 2009. Sediment instability on the Portuguese continental margin under abrupt glacial climate changes (last 60 kyr). Quaternary Science Reviews 28, 32113223.Google Scholar
Lévy, M., 2003. Mesoscale variability of phytoplankton and of new production: Impact of the large-scale nutrient distribution. Journal of Geophysical Research 108. http://dx.doi.org10.1029/2002JC001577.Google Scholar
Lisiecki, L.E., Raymo, M.E., 2005. A Pliocene-Pleistocene stack of 57 globally distributed benthic δ18O records. Paleoceanography 20, PA1003. http://dx.doi.org/1010.1029/2004PA001071.Google Scholar
Lombard, F., Labeyrie, L., Michel, E., Bopp, L., Cortijo, E., Retailleau, S., Howa, H., Jorissen, F., 2011. Modelling planktic foraminifer growth and distribution using an ecophysiological multi-species approach. Biogeosciences 8, 853873.Google Scholar
Marino, M., Maiorano, P., Tarantino, F., Voelker, A., Capotondi, L., Girone, A., Lirer, F., Flores, J.-A., Naafs, B.D.A., 2014. Coccolithophores as proxy of seawater changes at orbital-to-millennial scale during middle Pleistocene Marine Isotope Stages 14–9 in North Atlantic core MD01-2446. Paleoceanography 29, doi:10.1002/2013PA002574.Google Scholar
Martins, M.V.A, Rey, D., Pereira, E. Plaza-Morlote, M., Salgueiro, E., Moreno, J., Duleba, W., et al, 2017. Influence of dominant wind patterns in a distal region of the NW Iberian margin during the last glaciation. Journal of the Geological Society 175, 321.Google Scholar
Martrat, B., Grimalt, J.O., Shackleton, N.J., de Abreu, L., Hutterli, M.A., Stocker, T.F., 2007. Four climate cycles of recurring deep and surface water destabilizations on the Iberian Margin. Science 317, 502507.Google Scholar
McCave, I.N., Hall, I.R., 2006. Size sorting in marine muds: Processes, pitfalls, and prospects for paleoflow-speed proxies. Geochemistry, Geophysics, Geosystems 7, Q10N05. http://dx.doi.org/10.1029/2006GC001284.Google Scholar
McCave, I.N., Manighetti, B., Robinson, S.G., 1995. Sortable silt and fine sediment size/composition slicing: parameters for palaeocurrent speed and palaeoceanography. Paleoceanography 10, 593610.Google Scholar
McCave, I.N., Thornalley, D.J.R., Hall, I.R., 2017. Relation of sortable silt grain-size to deep-sea current speeds: calibration of the ‘Mud Current Meter’. Deep Sea Research Part I: Oceanographic Research Papers 127, 112.Google Scholar
McGillicuddy, D.J., Anderson, L.A., Bates, N.R., Bibby, T., Buesseler, K.O., Carlson, C.A., Davis, C.S., et al., 2007. Eddy/wind interactions stimulate extraordinary mid-ocean plankton blooms. Science 316, 10211026.Google Scholar
McGillicuddy, D.J., Robinson, A.R., Siegel, D.A., Jannasch, H.W., Johnson, R., Dickey, T.D., McNeil, J., Michaels, A.F., Knap, A.H., 1998. Influence of mesoscale eddies on new production in the Sargasso Sea. Nature 394, 263266.Google Scholar
Meckler, A.N., Sigman, D.M., Gibson, K.A., Francois, R., Martinez-Garcia, A., Jaccard, S.L., Rohl, U., Peterson, L.C., Tiedemann, R., Haug, G.H., 2013. Deglacial pulses of deep-ocean silicate into the subtropical North Atlantic Ocean. Nature 495, 495498.Google Scholar
Michel, E., Labeyrie, L.D., Duplessy, J.-C., Gorfti, N., Labracherie, M., Turon, J.-L., 1995. Could deep subantarctic convection feed the world deep basins during the last glacial maximum? Paleoceanography 10, 927941.Google Scholar
Mortlock, R.A., Froelich, P., 1989. A simple method for the rapid determination of biogenic opal in pelagic marine sediments. Deep-Sea Research 36, 14151426.Google Scholar
Nave, S., Freitas, P., Abrantes, F., 2001. Coastal upwelling in the Canary Island region: spatial variability reflected by the surface sediment diatom record. Marine Micropaleontology 42, 123.Google Scholar
Oliveira, P.B., Peliz, A., Dubert, J., Rosa, T.L., Santos, A.M.P., 2004. Winter geostrophic currents and eddies in the western Iberia coastal transition zone. Deep Sea Research Part I: Oceanographic Research Papers 51, 367381.Google Scholar
Oliver, K.I.C., Hoogakker, B.A.A., Crowhurst, S., Henderson, G.M., Rickaby, R.E.M., Edwards, N.R., Elderfield, H., 2010. A synthesis of marine sediment core δ13C data over the last 150 000 years. Climate of the Past 6, 645673.Google Scholar
Oschlies, A., 2002. Can eddies make ocean deserts bloom? Global Biogeochemical Cycles 16, 1106.Google Scholar
Oschlies, A., Garcon, V., 1998. Eddy-induced enhancement of primary production in a model of the North Atlantic Ocean. Nature 394, 266269.Google Scholar
Pahnke, K., Goldstein, S.L., Hemming, S.R., 2008. Abrupt changes in Antarctic Intermediate Water circulation over the past 25,000 years. Nature Geoscience 1, 870874.Google Scholar
Paillard, D., Labeyrie, L., Yiou, P., 1996. Macintosh program performs time-series analysis. Eos: Transactions of the American Geophysical Union 77, 379.Google Scholar
Pailler, D., Bard, E., 2002. High frequency palaeoceanographic changes during the past 140 000 yr recorded by the organic matter in sediments of the Iberian Margin. Palaeogeography, Palaeoclimatology, Palaeoecology 181, 431452.Google Scholar
Pelegri, J.L., Aristegui, J., Cana, L., Gonzalez-Davila, M., Hernandez-Guerra, A., Hernandez-Leon, S., Marrero-Diaz, A., Montero, M.F., Sangra, P., Santana-Casiano, M., 2005. Coupling between the open ocean and the coastal upwelling region off northwest Africa: water recirculation and offshore pumping of organic matter. Journal of Marine Systems 54, 337.Google Scholar
Peliz, A., Dubert, J., Santos, A.M.P., Oliveira, P.B., Cann, B.L., 2005. Winter upper ocean circulation in the Western Iberian Basin—fronts, eddies and poleward flows: an overview. Deep Sea Research Part I: Oceanographic Research Papers 52, 621646.Google Scholar
Pflaumann, U., Duprat, J., Pujol, C., Labeyrie, L.D., 1996. SIMMAX: A modern analog technique to deduce Atlantic sea surface temperatures from planktonic foraminifera in deep-sea sediments. Paleoceanography 11, 1535.Google Scholar
Pflaumann, U., Sarnthein, M., Chapman, M., d’Abreu, L., Funnell, B., Huels, M., Kiefer, T., et al., 2003. Glacial North Atlantic: sea-surface conditions reconstructed by GLAMAP 2000. Paleoceanography 18, 1065. http://dx.doi.org/1010.1029/2002PA000774.Google Scholar
Rickaby, R.E.M., Elderfield, H., 2005. Evidence from the high-latitude North Atlantic for variations in Antarctic Intermediate water flow during the last deglaciation. Geochemistry, Geophysics, Geosystems 6, Q05001. http://dx.doi.org/10.1029/2004GC000858.Google Scholar
Romero, O.E., Kim, J.-H., Donner, B., 2008. Submillennial-to-millennial variability of diatom production off Mauritania, NW Africa, during the last glacial cycle. Paleoceanography 23, PA3218. http://dx.doi.org/10.1029/2008PA001601.Google Scholar
Shackleton, N.J., Hall, M.A., Vincent, E., 2000. Phase relationships between millennial-scale events 64,000–24,000 years ago. Paleoceanography 15, 565569. http://dx.doi.org/10.1029/2000PA000513 Google Scholar
Salgueiro, E., Voelker, A.H.L., de Abreu, L., Abrantes, F., Meggers, H., Wefer, G., 2010. Temperature and productivity changes off the western Iberian margin during the last 150 ky. Quaternary Science Reviews 29, 680695.Google Scholar
Sanchez Goni, M.F., Loutre, M.F., Crucifix, M., Peyron, O., Santos, L., Duprat, J., Malaize, B., Turon, J.-L., Peypouquet, J.-P., 2005. Increasing vegetation and climate gradient in Western Europe over the Last Glacial Inception (122–110 ka): data-model comparison. Earth and Planetary Science Letters 231, 111130.Google Scholar
Sarmiento, J.L., Gruber, N., 2002. Sinks for anthropogenic carbon. Physics Today 55, 3036.Google Scholar
Schmieder, F., von Dobeneck, T., Bleil, U., 2000. The Mid-Pleistocene climate transition as documented in the deep South Atlantic Ocean: initiation, interim state and terminal event. Earth and Planetary Science Letters 179, 539549.Google Scholar
Schmittner, A., 2005. Decline of the marine ecosystem caused by a reduction in the Atlantic overturning circulation. Nature 434, 628633.Google Scholar
Schwab, C., Kinkel, H., Weinelt, M., Repschläger, J., 2012. Coccolithophore paleoproductivity and ecology response to deglacial and Holocene changes in the Azores Current System. Paleoceanography 27, PA3210. http://dx.doi.org/10.1029/2012PA002281.Google Scholar
Shackleton, N.J., Hall, M.A., Line, J., Cang, S., 1983. Carbon isotope data in core V19-30 confirm reduced carbon dioxide concentration in the ice age atmosphere. Nature 306, 319322.Google Scholar
Shackleton, N.J., Hall, M.A., Vincent, E., 2000. Phase relationships between millennial-scale events 64,000–24,000 years ago. Paleoceanography 15, 565569.Google Scholar
Shackleton, N.J., Opdyke, N.D., 1973. Oxygen isotope and palaeomagnetic stratigraphy of Equatorial Pacific core V28-238: oxygen isotope temperatures and ice volumes on a 105 year and 106 year scale. Quaternary Research 3, 3955.Google Scholar
Sousa, F.M., Bricaud, A., 1992. Satellite-derived phytoplankton pigment structures in the Portuguese upwelling area. Journal of Geophysical Research 97, 1134311356.Google Scholar
Stein, R., 1990. Organic carbon content/sedimentation rate relationship and its paleoenvironmental significance for marine sediments. Geo-Marine Letters 10, 3744.Google Scholar
Thomson, J., Nixon, S., Summerhayes, C.P., Rohling, E.J., Schönfeld, J., Zahn, R., Zahn, P., Abrantes, F.G., Gaspar, L., Vaqueiro, S., 2000. Enhanced productivity on the Iberian margin during glacial/interglacial transitions revealed by barium and diatoms. Journal of the Geological Society 157, 667677.Google Scholar
van Aken, H.M., 2000. The hydrography of the mid-latitude northeast Atlantic Ocean: I: the deep water masses. Deep Sea Research Part I: Oceanographic Research Papers 47, 757788.Google Scholar
Veres, D., Bazin, L., Landais, A., Toyé Mahamadou Kele, H., Lemieux-Dudon, B., Parrenin, F., Martinerie, P., et al., 2013. The Antarctic ice core chronology (AICC2012): an optimized multi-parameter and multi-site dating approach for the last 120 thousand years. Climate of the Past 9, 17331748.Google Scholar
Waelbroeck, C., Labeyrie, L., Michel, E., Duplessy, J.C., McManus, J.F., Lambeck, K., Balbon, E., Labracherie, M., 2002. Sea-level and deep water temperature changes derived from benthic foraminifera isotopic records. Quaternary Science Reviews 21, 295305.Google Scholar
Waelbroeck, C., Skinner, L.C., Labeyrie, L., Duplessy, J.C., Michel, E., Vazquez Riveiros, N., Gherardi, J.M., Dewilde, F., 2011. The timing of deglacial circulation changes in the Atlantic. Paleoceanography 26, PA3213. http://dx.doi.org/10.1029/2010PA002007.Google Scholar
Weeks, R., Laj, C., Endignoux, L., Fuller, M., Roberts, A., Manganne, R., Blanchard, E., Goree, W., 1993. Improvements in long‐core measurement techniques: Applications in palaeomagnetism and palaeoceanography. Geophysical Journal International 114, 651662.Google Scholar
Williams, R.G., 2011. Ocean eddies and plankton blooms. Nature Geosciences 4, 739740.Google Scholar
Williams, R.G., Follows, M.J., 1998. Eddies make ocean deserts bloom. Nature 394, 228229.Google Scholar
Yamazaki, T., Ikehara, M., 2012. Origin of magnetic mineral concentration variation in the Southern Ocean. Paleoceanography 27, PA2206. http://dx.doi.org/10.1029/2011PA002271.Google Scholar
Supplementary material: File

Nave et al. supplementary material

Nave et al. supplementary material 1

Download Nave et al. supplementary material(File)
File 4.9 MB