Hostname: page-component-8448b6f56d-wq2xx Total loading time: 0 Render date: 2024-04-25T04:19:20.673Z Has data issue: false hasContentIssue false

An environmental snapshot of the Bølling interstadial in Southern Iberia

Published online by Cambridge University Press:  20 January 2017

Antonio García-Alix*
Affiliation:
Departamento de Didáctica de la Ciencias Experimentales, Universidad de Granada, Granada, Spain Instituto Andaluz de Ciencias de la Tierra CSIC-UGR, Granada, Spain
Gonzalo Jiménez-Moreno
Affiliation:
Departamento de Estratigrafía y Paleontología, Universidad de Granada, Granada, Spain
Francisco J. Jiménez-Espejo
Affiliation:
Institute of Biogeosciences, Japan Agency for Marine-Earth Science and Technology, Yokosuka, Japan Department of Earth and Planetary Sciences, Graduate School of Environmental Studies, Nagoya University, Nagoya, Japan
Fernando García-García
Affiliation:
Departamento de Geología, Universidad de Jaén, Jaén, Spain
Antonio Delgado Huertas
Affiliation:
Instituto Andaluz de Ciencias de la Tierra CSIC-UGR, Granada, Spain
*
*Corresponding author at: Instituto Andaluz de Ciencias de la Tierra CSIC-UGR, Granada, Spain. E-mail address:agalix@ugr.es (A. García-Alix).

Abstract

The Bølling–Allerød interstadial is the closest warm time period to the Holocene. The study of the climate variability during this most recent warm scenario provides a natural record of potential environmental changes related with global temperature variations. Little is known about this interstadial in the Southern Iberian Peninsula. Therefore, the exceptional climatic record of the Otiñar paleo-lake (ca. 14.5–14.0 cal ka BP), provides environmental information about the first part of this interstadial (Bølling) in this key region. Although the studied high-resolution isotopic record point to almost invariant hydrological conditions in the paleo-lake, with little change in the carbon budget and important limestone dissolution, the pollen record shows an increase in forest species that can be interpreted as a warming trend and an increase in humidity during the Bølling in the area. This record is one of the few continental archives that show this climatic trend in Southern Iberia, agreeing with many other regional records from the western Mediterranean. This does not agree with higher latitude records that show an opposite trend. This opposite pattern in precipitation between the western Mediterranean and more northern latitudes could be explained by a persistent and increasing negative NAO mode during the Bølling in this area.

Type
Research Article
Copyright
University of Washington

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

Barcena, M.A., Cacho, I., Abrantes, F., Sierro, F.J., Grimalt, J.O., and Flores, J.A. Paleoproductivity variations related to climatic conditions in the Alboran Sea (western Mediterranean) during the last glacial–interglacial transition: the diatom record. Palaeogeography, Palaeoclimatology, Palaeoecology 167, (2001). 337357.CrossRefGoogle Scholar
Benavente, J. Investigaciones hidrogeológicas en la Sierra de Jaén. (PhD Thesis) (1978). University of Granada, Google Scholar
Benson, L.V., White, L.D., and Rye, R. Carbonate deposition. Pyramid Lake Subbasin, Nevada, 4. Comparison of the stable isotope values of carbonate deposits (tufas) and the Lafiontan lake-level record. Palaeogeography, Palaeoclimatology, Palaeoecology 122, (1996). 4576.Google Scholar
Bischoff, J.L., Stafford, T.W. Jr., and Rubin, M. A time–depth scale for Owens Lake sediments of core OL-92: radiocarbon dates and constant mass-accumulation rate. Geological Society of America, Special Paper 317, (1997). 9199.Google Scholar
Björck, S., Kromer, B., Johnsen, S., Bennike, O., Hammarlund, D., Lemdahl, G., Possnert, G., Rasmussen, T.L., Wohlfarth, B., Hammer, C.U., and Spurk, M. Synchronized terrestrial–atmospheric deglacial records around the North Atlantic. Science 274, (1996). 11551160.Google Scholar
Björck, S., Walker, M.J.C., Cwynar, L., Johnsen, S.J., Knudsen, K.-L., Lowe, J.J., Wohlfarth, B., INTIMATE Members, An event stratigraphy for the Last Termination in the North Atlantic region based on the Greenland Ice Core record: a proposal by the INTIMATE group. Journal of Quaternary Science 13, (1998). 283292.Google Scholar
Brauer, A., Endres, C., Gunter, C., Litt, T., Stebich, M., and Negendank, J.F.W. High resolution sediment and vegetation responses to Younger Dryas climate change in varved lake sediments from Meerfelder Maar, Germany. Quaternary Science Reviews 18, (1999). 321329.CrossRefGoogle Scholar
Broecker, W.S., Denton, G.H., Edwards, R.L., Cheng, H., Alley, R.B., and Putnam, A.E. Putting the Younger Dryas cold event into context. Quaternary Science Reviews 29, (2010). 10781081.Google Scholar
Carrión, J.S. Patterns and processes of Late Quaternary environmental change in a montane region of southwestern Europe. Quaternary Science Reviews 21, (2002). 20472066.Google Scholar
Carrión, J.S., and Dupré, M. Late Quaternary vegetational history at Navarrés, eastern Spain. A two-core approach. New Phytologist 134, (1996). 177191.Google Scholar
Carrión, J.S., and van Geel, B. Fine-resolution Upper Weichselian and Holocene palynological record from Navarrés (Valencia, Spain) and a discussion about factors of Mediterranean forest succession. Review of Palaeobotany and Palynology 106, (1999). 209236.Google Scholar
Combourieu Nebout, N., Peyron, O., Dormoy, I., Desprat, S., Beaudouin, C., Kotthoff, U., and Marret, F. Rapid climatic variability in the west Mediterranean during the last 25,000 years from high resolution pollen data. Climate of the Past 5, (2009). 503521.Google Scholar
Dansgaard, W., Johnsen, S.J., Clausen, H.B., Dahl-Jensen, D., Gundestrup, N.S., Hammer, C.U., Hvidberg, C.S., Steffensen, J.P., Sveinbjörnsdóttir, A.E., Jouzel, J., and Bond, G.C. Evidence for general instability of past climate from a 250 kyr ice-core record. Nature 264, (1993). 218220.Google Scholar
deMenocal, P., Ortiz, J., Guilderson, T., Adkins, J., Sarnthein, M., Baker, L., and Yarusinsky, M. Abrupt onset and termination of the African Humid Period: rapid climate responses to gradual insolation forcing. Quaternary Science Reviews 19, (2000). 347361.CrossRefGoogle Scholar
Denton, G.H., Anderson, R.F., Toggweiler, J.R., Edwards, R.L., Schaefer, J.M., and Putnam, A.E. The last glacial termination. Science 328, (2010). 16521656.Google Scholar
Faegri, K., and Iversen, J. Textbook of Pollen Analysis, IV. (1989). Wiley, New York.Google Scholar
Fernández, S., Fuentes, N., Carrión, J.S., González-Sampériz, P., Montoya, E., Gil, G., Vega Toscano, G., and Riquelme, J.A. The Holocene and Upper Pleistocene pollen sequence of Carihuela Cave, southern Spain. Geobios 40, (2007). 7590.Google Scholar
Fleitmann, D., Mudelsee, M., Burns, S.J., Bradley, R.S., Kramers, J., and Matter, A. Evidence for a widespread climatic anomaly at around 9.2 ka before present. Paleoceanography 23, (2008). http://dx.doi.org/10.1029/2007PA001519Google Scholar
Fletcher, W.J., and Sanchez Goñi, M.F. Orbital- and sub-orbital-scale climate impacts on vegetation of the western Mediterranean basin over the last 48,000 yr. Quaternary Research 70, (2008). 451464.CrossRefGoogle Scholar
Fletcher, W., Boski, T., and Moura, D. Palynological evidence for environmental and climatic change in the lower Guadiana valley (Portugal) during the last 13,000 years. The Holocene 17, (2007). 479492.Google Scholar
Fletcher, W.J., Sanchez Goñi, M.F., Peyron, O., and Dormoy, I. Abrupt climate changes of the last deglaciation detected in a Western Mediterranean forest record. Climate of the Past 6, (2010). 245264.Google Scholar
García�a-Alix, A., Delgado Huertas, A., and Martín Suárez, E. Unravelling the Late Pleistocene habitat of the southernmost woolly mammoths in Europe. Quaternary Science Reviews 32, (2012). 7585.CrossRefGoogle Scholar
García�a-Alix, A., Jiménez-Moreno, G., Anderson, R.S., Jiménez-Espejo, F., and Delgado-Huertas, A. Holocene paleoenvironmental evolution of a high-elevation wetland in Sierra Nevada, southern Spain, deduced from an isotopic record. Journal of Paleolimnology 48, (2012). 471484.Google Scholar
García�a-Alix, A., Jimenez Espejo, F.J., Lozano, J.A., Jimenez-Moreno, G., Martínez-Ruiz, F., García�a-Sanjuán, L., Aranda Jiménez, G., García�a Alfonso, E., Ruiz-Puertas, G., and Anderson, R.S. Anthropogenic impact and lead pollution throughout the Holocene in Southern Iberia. Science of the Total Environment 449, (2013). 451460.Google Scholar
García�a-García�a, F., Sánchez-Gómez, M., Navarro, V., and Pla, S. Formation, infill, and dissection of a latest-Pleistocene landslide-dammed reservoir (Betic Cordillera, Southern Spain): upstream and downstream geomorphological and sedimentological evidence. Quaternary International 233, (2011). 6171.Google Scholar
Genty, D., Blamart, D., Ghaleb, B., Plagnes, V., Causse, C., Bakalowicz, M., Zouari, K., Chkir, N., Hellstrom, J., Wainer, K., and Bourges, F. Timing and dynamics of the last deglaciation from European and North African δ13C stalagmite profiles e comparison with Chinese and South Hemisphere stalagmites. Quaternary Science Reviews 25, (2006). 21182142.CrossRefGoogle Scholar
Griffiths, S.J., Street-Perrott, F.A., Holmes, J.A., Leng, M.J., and Tzedakis, G. Chemical and isotopic composition of modern water bodies in the Lake Kopais Basin, central Greece: analogues for the interpretation of the lacustrine sedimentary sequence. Sedimentary Geology 148, (2002). 79103.Google Scholar
Grimm, E.C. CONISS: a FORTRAN 77 program for stratigraphically constrained cluster analysis by the method of incremental sum of squares. Computers and Geosciences 13, (1987). 1335.Google Scholar
Hatun, H., Sando, A.B., Drange, H., Hansen, B., and Valdimarsson, H. Influence of the Atlantic Subpolar Gyre on the thermohaline circulation. Science 309, (2005). 18411844.Google Scholar
Hodell, D.A., and Schelske, C.L. Production, sedimentation, and isotopic composition of organic matter in Lake Ontario. Limnology and Oceanography 43, (1998). 200214.Google Scholar
Hurrel, J.H. Decadal trends in the North Atlantic Oscillation: regional temperatures and precipitation. Science 269, (1995). 676679.Google Scholar
Hurrel, J.H., Kushnir, Y., Ottersen, G., and Visbeck, M. An overview of the North Atlantic Oscillation. Geophysical Monograph 134, (2003). 135.Google Scholar
Jiménez-Espejo, F.J., Martínez-Ruiz, F., Rogerson, M., González-Donoso, J.M., Romero, O.E., Linare, D., Sakamoto, T., Gallego-Torres, D., Rueda Ruiz, J.L., Ortega-Huertas, M., and Perez Claros, J.A. Detrital input, productivity fluctuations, and water mass circulation in the westernmost Mediterranean Sea since the Last Glacial Maximum. Geochemistry, Geophysics, Geosystems 9, (2008). Q11U02 http://dx.doi.org/10.1029/2008GC002096Google Scholar
Jiménez-Moreno, G., and Anderson, R.S. Holocene vegetation and climate change recorded in alpine bog sediments from the Borreguiles de la Virgen, Sierra Nevada, southern Spain. Quaternary Research 77, (2012). 4453.Google Scholar
Jiménez-Moreno, G., García�a-Alix, A., Hernández-Corbalán, M.D., Anderson, R.S., and Delgado-Huertas, A. Vegetation, fire, climate and human disturbance history in the southwestern Mediterranean area during the late Holocene. Quaternary Research 79, (2013). 110122. http://dx.doi.org/10.1016/j.yqres.2012.11.008Google Scholar
Johnsen, S.J. GRIP Oxygen Isotopes. (1999). http://dx.doi.org/10.1594/PANGAEA,55091Google Scholar
Johnsen, S.J., Dahl-Jensen, D., Gundestrup, N., Steffensen, J.P., Clausen, H.B., Miller, H., Masson-Delmotte, V., Sveinbjörnsdottir, A.E., and White, J. Oxygen isotope and palaeotemperature records from six Greenland ice-core stations: Camp Century, Dye-3, GRIP, GISP2, Renland and NorthGRIP. Journal of Quaternary Science 16, (2001). 299307.Google Scholar
Johnson, G.C., and Gruber, N. Decadal water mass variations along 20°W in the Northeastern Atlantic Ocean. Progress in Oceanography 73, (2007). 277295.Google Scholar
Kirsch, H.J. Illite crystallinity: recommendations on sample preparation, X-ray diffraction settings, and interlaboratory samples. Journal of Metamorphic Geology 9, (1991). 665670.Google Scholar
Li, H.-C., and Ku, T.-L. δ13C–δ18O covariance as a paleohydrological indicator for closed-basin lakes. Palaeogeography, Palaeoclimatology, Palaeoecology 133, (1997). 6980.Google Scholar
Liu, Z., Otto-Bliesner, B.L., He, F., Brady, E.C., Tomas, R., Clark, P.U., Carlson, A.E., Lynch-Stieglitz, J., Curry, W., Brook, E., Erickson, D., Jacob, R., Kutzbach, J., and Cheng, J. Transient simulation of last deglaciation with a new mechanism for Bolling–Allerod warming. Science 325, (2009). 310314.Google Scholar
Lowe, J.J., Hoek, W., INTIMATE Group, Inter-regional correlation of palaeoclimatic records for the last glacial–interglacial transition: a protocol for improved precision recommended by the INTIMATE project group. Quaternary Science Reviews 20, (2001). 11751187.Google Scholar
Lowe, J.J., Rasmussen, S.O., Bjorck, S., Hoek, W.Z., Steffensen, J.P., Walker, M.J.C., Yu, Z.C., INTIMATE Group, Synchronisation of palaeoenvironmental events in the North Atlantic region during the last termination: a revised protocol recommended by the INTIMATE group. Quaternary Science Reviews 27, (2008). 617.Google Scholar
Martin, J.D. Using XPowder: A Software Package for Powder X-Ray Diffraction Analysis. (2004). (DL GR 1001/04: Spain. http://www.xpowder.com)Google Scholar
Martrat, B., Grimalt, J.O., López-Martínez, C., Cacho, I., Sierro, F.J., Flores, J.A., Zahn, R., Canals, M., Curtis, J.H., and Hodell, D.A. Abrupt temperature changes in the Western Mediterranean over the past 250,000 years. Science 306, (2004). 17621765.Google Scholar
Martrat, B., Grimalt, J.O., Shackleton, N.J., Abreu, L., Hutterli, M.A., and Stocker, T.F. Four climate cycles of recurring deep and surface water destabilizations on the Iberian Margin. Science 317, (2007). 502507.Google Scholar
McCrea, J.M. On the isotopic chemistry of carbonates and a paleotemperature scale. Journal of Chemical Physics 18, (1950). 849857.Google Scholar
McKenzie, J.A. Carbon isotopes and productivity in the lacustrine and marine environments. Stumm, W. Chemical Processes in Lakes. (1985). Wiley, New York. 99118.Google Scholar
McManus, J.F., Francois, R., Gherardi, J.-M., Keigwin, L., and Brown-Leger, S. Collapse and rapid resumption of Atlantic meridional circulation linked to deglacial climate changes. Nature 428, (2004). 834837.CrossRefGoogle ScholarPubMed
Meyers, P.A. Preservation of elemental and isotopic source identification of sedimentary organic matter. Chemical Geology 113, (1994). 289302.Google Scholar
Meyers, P.A., and Teranes, J.L. Sediment organic matter. Last, W.M., and Smol, J.P. Tracking Environmental Changes Using Lake Sediments: Physical and Chemical Techniques. (2001). Kluwer Academic Publishers, Dordrecht. 239270.Google Scholar
Muñoz Sobrino, C., Heiri, O., Hazekamp, M., Velden, D., van der Kirilova, E.P., García�a-Moreiras, I., and Lotter, A.F. New data on the Lateglacial period of SW Europe: a high resolution multiproxy record from Laguna de la Roya (NW Iberia). Quaternary Science Reviews 80, (2013). 5877.Google Scholar
Morellón, M., Valero-García�s, B.L., González-Sampériz, P., Vegas-Vilarrúbia, T., Rubio, E., Rieradevall, M., Delgado-Huertas, A., Mata, P., Romero, O., Engstrom, D.R., López-Vicente, M., Navas, A., and Soto, J. Climate changes and human activities recorded in the sediments of Lake Estanya (NE Spain) during the Medieval Warm Period and Little Ice Age. Journal of Paleolimnology 46, (2011). 423452.Google Scholar
Moreno, A., Stoll, H., Jiménez-Sánchez, M., Cacho, I., Valero-García�s, B., Ito, E., and Edwards, R.L. A speleothem record of glacial (25–11.6 kyr BP) rapid climatic changes from northern Iberian Peninsula. Global and Planetary Change 71, (2010). 218231.Google Scholar
Moreno, A., López-Merino, L., Leira, M., Marco-Barba, J., González-Sampériz, P., Valero-García�s, B.L., López-Sáez, A., Santos, L., Mata, P., and Ito, E. Revealing the last 13,500 years of environmental history from the multiproxy record of a mountain lake (Lago Enol, northern Iberian Peninsula). Journal of Paleolimnology 46, (2011). 327349.Google Scholar
O'Doherty, L., Sandoval, J., Bartolini, A., Bruchez, S., Bill, M., and Guex, J. Carbon-isotope stratigraphy and ammonite faunal turnover for the Middle Jurassic in the Southern Iberian palaeomargin. Palaeogeography, Palaeoclimatology, Palaeoecology 239, (2006). 311333.CrossRefGoogle Scholar
O'Leary, M.H. Carbon isotopes in photosynthesis. Bioscience 38, (1988). 328336.Google Scholar
Ortiz, J.E., Torres, T., Delgado, A., Julia, R., Lucini, M., Llamas, F.J., Reyes, E., Soler, V., and Valle, M. The palaeoenvironmental and palaeohydrological evolution of Padul Peat Bog (Granada, Spain) over one million years, from elemental, isotopic and molecular organic geochemical proxies. Organic Geochemistry 35, (2004). 12431260.Google Scholar
Prado, A. El sistema termal de Alicún de las Torres (Granada) como análogo natural de escape de CO2 en forma de DIC: implicaciones paleoclimáticas y como sumidero de CO2. (PhD Thesis) (2012). Universidad Complutense de Madrid, Google Scholar
Rasmussen, S.O., Andersen, K.K., Svensson, A.M., Steffensen, J.P., Vinther, B.M., Clausen, H.B., Siggaard-Andersen, M.-L., Johnsen, S.J., Larsen, L.B., Dahl-Jensen, D., Bigler, M., Röthlisberger, R., Fischer, H., Goto-Azuma, K., Hansson, M.E., and Ruth, U. A new Greenland ice core chronology for the last glacial termination. Journal of Geophysical Research 111, (2006). D06102 http://dx.doi.org/10.1029/2005JD006079Google Scholar
Rasmussen, S.O., Seierstad, I.K., Andersen, K.K., Bigler, M., Dahl-Jensen, D., and Johnsen, S.J. Synchronization of the NGRIP, GRIP and GISP2 ice cores across MIS2 and palaeoclimatic implications. Quaternary Science Reviews 27, (2008). 1828.Google Scholar
Reyes, E., Pérez del Villar, L., Delgado, A., Cortecci, G., Núñez, R., Pelayo, M., and Cózar, J.S. Carbonatation processes at the El Berrocal natural analogue granitic system (Spain): inferences from mineralogical and stable isotope studies. Chemical Geology 150, (1998). 293315.Google Scholar
Rodrigo Gámiz, M., Martínez Ruiz, F., Jiménez Espejo, F.J., Gallego Torres, D., Nieto Moreno, V., Martín Ramos, D., Ariztegui, D., and Romero, O. Impact of climate variability in the western Mediterranean during the last 20,000 years: oceanic and atmospheric responses. Quaternary Science Reviews 15–16, (2011). 20182034.Google Scholar
Rodrigo-Gámiz, M., Martínez-Ruiz, F., Rampen, S.W., Schouten, S., and Sinninghe Damsté, J.S. Sea surface temperature variations in the western Mediterrenean Sea over the last 20 kry: a dual-organic proxy (Uk37 and LDI) approach. Paleoceanography (2013). http://dx.doi.org/10.1002/2013PA002466Google Scholar
Romanek, C.S., Grossman, E.L., and Morse, J.W. Carbon isotopic fractionation in synthetic aragonite and calcite: effects of temperature and precipitation rate. Geochimica et Cosmochimica Acta 56, (1992). 419430.CrossRefGoogle Scholar
Street-Perrott, F.A., Huang, Y., Perrott, R.A., Eglinton, G., Barker, P., Khelifa, L.B., Harkness, D.D., and Olago, D.O. Impact of lower atmospheric carbon dioxide on tropical mountain ecosystems. Science 278, (1997). 14221426.Google Scholar
Stuiver, M., Reimer, P.J., Bard, E., Beck, J.W., Burr, G.S., Hughen, K.A., Kromer, B., McCormac, F.G., Plicht, J., and Spurk, M. INTCAL98 radiocarbon age calibration 24,000–0 cal BP. Radiocarbon 40, (1998). 10411083.Google Scholar
Talbot, M.R. A review of the palaeohydrological interpretation of carbon and oxygen isotopic ratios in primary lacustrine carbonates. Chemical Geology. Isotope Geoscience Section 80, (1990). 261279.CrossRefGoogle Scholar
Talbot, M.R., and Laerdal, T. The Lake Pleistocene–Holocene palaeolimnology of Lake Victoria, East Africa, based upon elemental and isotopic analyses of sedimentary organic matter. Journal of Paleolimnology 23, (2000). 141164.Google Scholar
van Raden, U.J., Colombaroli, D., Gilli, A., Schwanderc, J., Bernasconia, S.M., van Leeuwend, J., Leuenbergerc, M., and Eicheet, U. High-resolution late-glacial chronology for the Gerzensee lake record (Switzerland): δ18O correlation between a Gerzensee-stack and NGRIP. Palaeogeography, Palaeoclimatology, Palaeoecology 391, (2013). 1324.Google Scholar
Von Grafenstein, U., Erlenkauser, H., Brauer, A., Jouzel, J., and Johnsen, S.J. A mid-European decadal isotope-climate record from 15,500 to 5,000 years BP. Science 284, (1999). 16541657.Google Scholar
Wolfe, B.B., Edwards, T.W.D., Beuning, K.R.M., and Elgood, R.J. Carbon and oxygen isotope analysis of lake sediment cellulose: methods and applications. Last, W.M., and Smol, J.P. Tracking Environmental Changes Using Lake Sediments: Physical and Chemical Techniques. (2001). Kluwer Academic Publishers, Dordrecht. 373400.Google Scholar
Yu, Z.C., and Eicher, U. Abrupt climate oscillations during the last deglaciation in central North America. Science 282, (1998). 22352238.Google Scholar