Hostname: page-component-7c8c6479df-995ml Total loading time: 0 Render date: 2024-03-29T02:06:01.482Z Has data issue: false hasContentIssue false

Oxygen isotope evidence of Little Ice Age aridity on the Caribbean slope of the Cordillera Central, Dominican Republic

Published online by Cambridge University Press:  20 January 2017

Chad S. Lane*
Affiliation:
Department of Geography, University of Tennessee, Knoxville, Tennessee, 37996, TN, USA
Sally P. Horn
Affiliation:
Department of Geography, University of Tennessee, Knoxville, Tennessee, 37996, TN, USA
Kenneth H. Orvis
Affiliation:
Department of Geography, University of Tennessee, Knoxville, Tennessee, 37996, TN, USA
John M. Thomason
Affiliation:
Department of Geography, University of Tennessee, Knoxville, Tennessee, 37996, TN, USA
*
Corresponding author. Fax: +1 910 962 7077.

E-mail address:lanec@uncw.edu (C.S. Lane).

Abstract

Climate change during the so-called Little Ice Age (LIA) of the 15th to 19th centuries was once thought to be limited to the high northern latitudes, but increasing evidence reflects significant climate change in the tropics. One of the hypothesized features of LIA climate in the low latitudes is a more southerly mean annual position of the Intertropical Convergence Zone (ITCZ), which produced more arid conditions through much of the northern tropics. High-resolution stable oxygen isotope data and other sedimentary evidence from Laguna de Felipe, located on the Caribbean slope of the Cordillera Central of the Dominican Republic, support the hypothesis that the mean annual position of the ITCZ was displaced significantly southward during much of the LIA. Placed within the context of regional paleoclimate and paleoceanographic records, and reconstructions of global LIA climate, this shift in mean annual ITCZ position appears to have been induced by lower solar insolation and internal dynamical responses of the global climate system. Our results from Hispaniola further emphasize the global nature of LIA climate change and the sensitivity of circum-Caribbean climate conditions to what are hypothesized to be relatively small variations in global energy budgets.

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.)

Footnotes

1

Current Address: Department of Geography and Geology, University of North Carolina-Wilmington, Wilmington, NC, 28403, USA.

References

Anchukaitis, K.J., and Horn, S.P. A 2000-year reconstruction of forest disturbance from southern Pacific Costa Rica. Palaeogeography Palaeoclimatology Palaeoecology 221, (2005). 3554.Google Scholar
Black, D.E., Thunell, R.C., Kaplan, A., Peterson, L.C., and Tappa, E.J. A 2000-year record of Caribbean and tropical North Atlantic hydrographic variability. Paleoceanography 19, (2004). PA2022 CrossRefGoogle Scholar
Black, D.E., Abahazi, M.A., Thunell, R.C., Kaplan, A., Tappa, E.J., and Peterson, L.C. An 8-century tropical Atlantic SST record from the Cariaco Basin: baseline variability, twentieth-century warming, and Atlantic hurricane frequency. Paleoceanography 22, (2007). PA4204 CrossRefGoogle Scholar
Bolay, E. The Dominican Republic: a Country Between Rain Forest and Desert; Contributions to the Ecology of a Caribbean Island. (1997). Margraf Verlag, Weikersheim.Google Scholar
Broecker, W.S. Was a change in thermohaline circulation responsible for the Little Ice Age?. Proceedings of the National Academy of Sciences of the United States of America 97, (2000). 13391342.CrossRefGoogle ScholarPubMed
Brown, E.T., and Johnson, T.C. Coherence between tropical East African and South American records of the Little Ice Age. Geochemistry Geophysics Geosystems 6, (2005). Q12005 CrossRefGoogle Scholar
Chiang, J.C.H., Kushnir, Y., and Giannini, A. Deconstructing Atlantic Intertropical Convergence Zone variability: influence of the local cross-equatorial sea surface temperature gradient and remote forcing from the eastern equatorial Pacific. Journal of Geophysical Research—Atmospheres 107, (2002). ACL3.1–ACL3.19 Google Scholar
Chiang, J.C.H., Cheng, W., and Bitz, C.M. Teleconnection mechanisms to the tropical Atlantic from an abrupt freshening of the North Atlantic Ocean. Geophysical Research Letters 35, (2008). L07704 CrossRefGoogle Scholar
Chivas, A.R., De Deckker, P., and Shelley, J.M.G. Magnesium and strontium in nonmarine ostracod shells as indicators of paleosalinity and paleotemperature. Hydrobiologia 143, (1986). 135142.CrossRefGoogle Scholar
Clement, A.C., Seager, R., Cane, M.A., and Zebiak, S.E. An ocean dynamical thermostat. Journal of Climate 9, (1996). 21902196.Google Scholar
Cohen, A.S. Paleolimnology: the History and Evolution of Lake Systems. (2003). Oxford University Press, New York.CrossRefGoogle Scholar
Colinvaux, P.A., De Oliveira, P.E., and Moreno, J.E. Amazon Pollen Manual and Atlas. (1999). Hardwood Academic Publishers, Amsterdam.Google Scholar
Covich, A., and Stuiver, M. Changes in oxygen-18 as a measure of long-term fluctuations in tropical lake levels and molluscan populations. Limnology and Oceanography 19, (1974). 682691.CrossRefGoogle Scholar
Craig, H. The measurement of oxygen isotope paleotemperatures. Tongiorgi, E. Stable Isotopes in Oceanographic Studies and Paleotemperatures. (1965). Congiglio Nazionale della Richereche, Laboratorio de Geologia Nucleare, Pisa. 9130.Google Scholar
Cronin, T.M., Dwyer, G.S., Kamiya, T., Schwede, S., and Willard, D.A. Medieval Warm Period, Little Ice Age and 20th century temperature variability from Chesapeake Bay. Global and Planetary Change 36, (2003). 1729.CrossRefGoogle Scholar
Cronin, T.M., Thunell, R., Dwyer, G.S., Saenger, C., Mann, M.E., Vann, C., and Seal, R.R. Multiproxy evidence of Holocene climate variability from estuarine sediments, eastern North America. Paleoceanography 20, (2005). PA4006 Google Scholar
Curtis, J.H., and Hodell, D.A. An isotopic and trace element study of ostracods from Lake Miragoane, Haiti: a 10,500 year record of paleosalinity and paleotemperature changes in the Caribbean. Geophysical Monograph 78, (1993). 135152.Google Scholar
Curtis, J.H., Hodell, D.A., and Brenner, M. Climate variability on the Yucatan Peninsula (Mexico) during the past 3500 years, and implications for Maya cultural evolution. Quaternary Research 46, (1996). 3747.Google Scholar
Diefendorf, A.F., Patterson, W.P., Mullins, H.T., Tibert, N., and Martini, A. Evidence for high-frequency late Glacial to mid-Holocene (16,800 to 5500 cal yr B.P.) climate variability from oxygen isotope values of Lough Inchiquin, Ireland. Quaternary Research 65, (2006). 7886.Google Scholar
Donnelly, J.P., and Woodruff, J.D. Intense hurricane activity over the past 5, 000 years controlled by El Nino and the West African monsoon. Nature 447, (2007). 465468.Google Scholar
Engstrom, D.R., and Nelson, S.R. Paleosalinity from trace-metals in fossil ostracodes compared with observational records at Devils Lake, North-Dakota, USA. Palaeogeography Palaeoclimatology Palaeoecology 83, (1991). 295312.CrossRefGoogle Scholar
Epstein, S., Bruchsbaum, R., Lowenstam, H.A., and Urey, H.C. Revised carbonate-water isotopic temperature scale. Geological Society of America Bulletin 64, (1953). 13151326.Google Scholar
Fontes, J.C., and Gonfiantini, R. Comportement isotopique au cours de l'evaporation de deux bassins sahariens. Earth and Planetary Science Letters 3, (1967). 258266.CrossRefGoogle Scholar
Gasse, F., Tehet, R., Durand, A., Gilber, E., and Fontes, J.C. The arid–humid transition in the Sahara and the Sahel during the last deglaciation. Nature 346, (1990). 141146.CrossRefGoogle Scholar
Giannini, A., Kushnir, Y., and Cane, M.A. Interannual variability of Caribbean rainfall, ENSO, and the Atlantic Ocean. Journal of Climate 13, (2000). 297311.2.0.CO;2>CrossRefGoogle Scholar
Giannini, A., Cane, M.A., and Kushnir, Y. Interdecadal changes in the ENSO teleconnection to the Caribbean region and the North Atlantic oscillation. Journal of Climate 14, (2001). 28672879.2.0.CO;2>CrossRefGoogle Scholar
Giannini, A., Chiang, J.C.H., Cane, M.A., Kushnir, Y., and Seager, R. The ENSO teleconnection to the tropical Atlantic Ocean: contributions of the remote and local SSTs to rainfall variability in the tropical Americas. Journal of Climate 14, (2001). 45304544.2.0.CO;2>CrossRefGoogle Scholar
Goulden, M.L., Litvak, M., and Miller, S.D. Factors that control Typha marsh evapotranspiration. Aquatic Botany 86, (2007). 97106.Google Scholar
Gupta, A.K., Anderson, D.M., and Overpeck, J.T. Abrupt changes in the Asian southwest monsoon during the Holocene and their links to the North Atlantic Ocean. Nature 421, (2003). 354357.Google Scholar
Haase-Schramm, A., Bohm, F., Eisenhauer, A., Dullo, W.C., Joachimski, M.M., Hansen, B., and Reitner, J. Sr/Ca ratios and oxygen isotopes from sclerosponges: temperature history of the Caribbean mixed layer and thermocline during the Little Ice Age. Paleoceanography 18, (2003). 1073 Google Scholar
Haug, G.H., Hughen, K.A., Sigman, D.M., Peterson, L.C., and Rohl, U. Southward migration of the intertropical convergence zone through the Holocene. Science 293, (2001). 13041308.Google Scholar
Haug, G.H., Gunther, D., Peterson, L.C., Sigman, D.M., Hughen, K.A., and Aeschlimann, B. Climate and the collapse of Maya civilization. Science 299, (2003). 17311735.Google Scholar
Heaton, T.H.E., Holmes, J.A., and Bridgwater, N.D. Carbon and oxygen isotope variations among lacustrine ostracods: implications for palaeoclimatic studies. Holocene 5, (1995). 428434.CrossRefGoogle Scholar
Hegerl, G., Crowley, T., Baum, S., Kim, K., and Hyde, W. Detection of volcanic, solar, and greenhouse signals in paleo-reconstructions of Northern Hemisphere temperature. Geophysical Research Letters 30, (2003). 1242 http://dx.doi.org/10.1029/2002GL0166335Google Scholar
Hegerl, G., Crowley, T., Hyde, W., and Frame, D. Constraints on climate sensitivity from temperature reconstructions of the last millennium. Nature 440, (2006). 10291032.Google Scholar
Helama, S., Timonen, M., Holopainen, J., Ogurtsov, M.G., Mielikainen, K., Eronen, M., Lindholm, M., and Merilainen, J. Summer temperature variations in Lapland during the Medieval Warm Period and the Little Ice Age relative to natural instability of thermohaline circulation on multi-decadal and multi-centennial scales. Journal of Quaternary Science 24, (2009). 450456.Google Scholar
Hodell, D.A., Curtis, J.H., and Brenner, M. Possible role of climate in the collapse of Classic Maya civilization. Nature 375, (1995). 391394.Google Scholar
Hodell, D.A., Brenner, M., Curtis, J.H., and Guilderson, T. Solar forcing of drought frequency in the Maya lowlands. Science 292, (2001). 13671370.CrossRefGoogle ScholarPubMed
Hodell, D.A., Brenner, M., and Curtis, J.H. Terminal Classic drought in the northern Maya lowlands inferred from multiple sediment cores in Lake Chichancanab (Mexico). Quaternary Science Reviews 24, (2005). 14131427.Google Scholar
Hodell, D.A., Brenner, M., Curtis, J.H., Medina-Gonzalez, R., Can, E.I.C., Albornaz-Pat, A., and Guilderson, T.P. Climate change on the Yucatan Peninsula during the Little Ice Age. Quaternary Research 63, (2005). 109121.CrossRefGoogle Scholar
Holzhauser, H., Magny, M., and Zumbuhl, H.J. Glacier and lake-level variations in west-central Europe over the last 3500 years. Holocene 15, (2005). 789801.Google Scholar
Horn, S.P. Postglacial vegetation and fire history in the Chirripó páramo of Costa Rica. Quaternary Research 40, (1993). 107116.Google Scholar
Horn, S.P., and Sanford, R.L. Holocene fires in Costa-Rica. Biotropica 24, (1992). 354361.Google Scholar
Ito, E., De Deckker, P., and Eggins, S.M. Ostracodes and their shell chemistry: implications for paleohydrologic and paleoclimatologic applications. Paleontological Society Papers 9, (2003). 119152.Google Scholar
Keatings, K.W., Heaton, T.H.E., and Holmes, J.A. Carbon and oxygen isotope fractionation in non-marine ostracods: results from a 'natural culture' environment. Geochimica Et Cosmochimica Acta 66, (2002). 17011711.Google Scholar
Keigwin, L. The Little Ice Age and Medieval warm period in the Sargasso Sea. Science 274, (1996). 15041507.Google Scholar
Kim, S.T., and O'Neil, J.R. Equilibrium and nonequilibrium oxygen isotope effects in synthetic carbonates. Geochimica Et Cosmochimica Acta 61, (1997). 34613475.Google Scholar
Kreutz, K.J., Mayewski, P.A., Meeker, L.D., Twickler, M.S., Whitlow, S.I., and Pittalwala, I.I. Bipolar changes in atmospheric circulation during the Little Ice Age. Science 277, (1997). 12941296.Google Scholar
Lafleur, P.M. Evapotranspiration from sedge-dominated wetland surfaces. Aquatic Botany 37, (1990). 341353.Google Scholar
Lane, C.S., Horn, S.P., Mora, C.I., and Orvis, K.H. Late-Holocene paleoenvironmental change in the mid-elevations of the Dominican Republic: a multi-site, multi-proxy analysis. Quaternary Science Reviews 28, (2009). 22392260.Google Scholar
Linsley, B., Dunbar, R., Wellington, G.M., and Mucciarone, D. A coral-based reconstruction of Intertropical Convergence Zone variability over Central America since 1707. Journal of Geophysical Research 99, (1994). 99779994.CrossRefGoogle Scholar
Lister, G.S. Stable isotopes from lacustrine ostracoda as tracers for continental palaeoenvironments. De Deckker, P., Colin, J.P., and Peypouquet, J.P. Ostracoda in the Earth Sciences. (1988). Elsevier, Amsterdam. 201218.Google Scholar
Lozano-Garcia, M., Caballero, M., Ortega, B., Rodriguez, A., and Sosa, S. Tracing the effects of the Little Ice Age in the tropical lowlands of eastern Mesoamerica. Proceedings of the National Academy of Sciences of the United States of America 104, (2007). 1620016203.CrossRefGoogle ScholarPubMed
Lund, D.C., and Curry, W.B. Late Holocene variability in Florida current surface density: patterns and possible causes. Paleoceanography 19, (2004). CrossRefGoogle Scholar
Lund, D.C., Lynch-Stieglitz, J., and Curry, W.B. Gulf Stream density structure and transport during the past millennium. Nature 444, (2006). 601604.Google Scholar
Malmgren, B.A., Winter, A., and Chen, D. El-Niño-southern oscillation and North Atlantic oscillation control of climate in Puerto Rico. Journal of Climate 11, (1998). 27132717.Google Scholar
Mann, M.E., Zhang, Z., Rutherford, S., Bradley, R.S., Hughes, M.K., Shindell, D., Ammann, C., Faluvegi, C., and Ni, F. Global signatures and dynamical origins of the Little Ice Age and Medieval Climate Anomaly. Science 326, (2009). 12561260.Google Scholar
Marchitto, T.M., and deMenocal, P.B. Late Holocene variability of upper North Atlantic deep water temperature and salinity. Geochemistry Geophysics Geosystems 4, (2003). Google Scholar
Markgraf, V., Baumgartner, T.R., Bradbury, J.P., Diaz, H.F., Dunbar, R.B., Luckman, B.H., Seltzer, G.O., Swetnam, T.W., and Villalba, R. Paleoclimate reconstruction along the Pole-Equator–Pole transect of the Americas (PEP 1). Quaternary Science Reviews 19, (2000). 125140.Google Scholar
Mayewski, P.A., Rohling, E.E., Stager, J.C., Karlen, W., Maasch, K.A., Meeker, L.D., Meyerson, E.A., Gasse, F., van Kreveld, S., Holmgren, K., Lee-Thorp, J., Rosqvist, G., Rack, F., Staubwasser, M., Schneider, R.R., and Steig, E.J. Holocene climate variability. Quaternary Research 62, (2004). 243255.Google Scholar
Moberg, A., Sonechkin, D.M., Holmgren, K., Datsenko, N.M., and Karlen, W. Highly variable Northern Hemisphere temperatures reconstructed from low- and high-resolution proxy data. Nature 443, (2005). 613617.Google Scholar
Nyberg, J., Kuijpers, A., Malmgren, B.A., and Kunzendorf, H. Late Holocene changes in precipitation and hydrography recorded in marine sediments from the northeastern Caribbean Sea. Quaternary Research 56, (2001). 87102.CrossRefGoogle Scholar
Nyberg, J., Malmgren, B.A., Kuijpers, A., and Winter, A. A centennial-scale variability of tropical North Atlantic surface hydrography during the late Holocene. Palaeogeography Palaeoclimatology Palaeoecology 183, (2002). 2541.Google Scholar
Oana, S., and Deevey, E.S. Carbon 13 in lake waters and its possible bearing on paleolimnology. American Journal of Science 258A, (1960). 253272.Google Scholar
O'Brien, S.R., Mayewski, P.A., Meeker, L.D., Meese, D.A., Twickler, M.S., and Whitlow, S.I. Complexity of Holocene climate as reconstructed from a Greenland ice core. Science 270, (1995). 19621964.Google Scholar
Peterson, L.C., and Haug, G.H. Variability in the mean latitude of the Atlantic Intertropical Convergence Zone as recorded by riverine input of sediments to the Cariaco Basin (Venezuela). Palaeogeography Palaeoclimatology Palaeoecology 234, (2006). 97113.Google Scholar
Polissar, P.J., Abbott, M., Wolfe, A.P., Bezada, M., Rull, V., and Bradley, R.S. Solar modulation of Little Ice Age climate in the tropical Andes. Proceedings of the National Academy of Sciences of the United States of America 103, (2006). 89378942.CrossRefGoogle ScholarPubMed
Price, J.S. Evapotranspiration from a lakeshore Typha marsh on Lake Ontario. Aquatic Botany 48, (1994). 261272.Google Scholar
Purper, I. Cytheridella boldii Purper, sp. nov. (Ostracoda) from Venezuela and a revision of the Genus Cytheridella Daday, 1905. Anais da Academia Brasileira de Cîencias 46, (1974). 636662.Google Scholar
Rabatel, A., Francou, B., Jomelli, V., Naveau, P., and Grancher, D. A chronology of the Little Ice Age in the tropical Andes of Bolivia (16°S) and its implications for climatic reconstruction. Quaternary Research 70, (2008). 198212.CrossRefGoogle Scholar
Reimer, P.J., Baillie, M.G.L., Bard, E., Bayliss, A., Beck, J.W., Blackwell, P.G., Bronk Ramsey, C., Buck, C.E., Burr, G.S., Edwards, R.L., Friedrich, M., Grootes, P.M., Guilderson, T.P., Hajdas, I., Heaton, T.J., Hogg, A.G., Hughen, K.A., Faiser, K.F., Kromer, F.G., McCormac, F.G., Manning, S.W., Reimer, R.W., Richards, D.A., Southon, J.R., Talamo, S., Turney, C.S.M., van der Plicht, J., and Weyhenmeyer, C.E. IntCal09 and Marine09 radiocarbon age calibration curves, 0–50, 000 years cal BP. Radiocarbon 51, (2009). 11111150.Google Scholar
Russell, J.M., and Johnson, T.C. Little Ice Age drought in equatorial Africa: Intertropical Convergence Zone migrations and El Nino-Southern Oscillation variability. Geology 35, (2007). 2124.Google Scholar
Sachs, J.P., Sachse, D., Smittenberg, R.H., Zhang, Z., Battisti, D.S., and Golubic, S. Southward movement of the Pacific intertropical convergence zone AD 1400–1850. Nature Geoscience 2, (2009). 519525.Google Scholar
Saenger, C., Cohen, A.L., Oppo, D.W., Halley, R.B., and Carilli, J.E. Surface-temperature trends and variability in the low-latitude North Atlantic since 1552. Nature Geoscience 2, (2009). 492495.Google Scholar
Shindell, D.T., Schmidt, G.A., Mann, M.E., Rind, D., and Waple, A. Solar forcing of regional climate change during the Maunder minimum. Science 294, (2001). 21492152.Google Scholar
Stone, R. Tree rings tell of Angkor's dying days. Science 323, (2009). 999 Google Scholar
Stuiver, M. Oxygen and carbon isotope ratios of fresh-water carbonates as climatic indicators. Journal of Geophysical Research 75, (1970). 52475257.CrossRefGoogle Scholar
Stuiver, M., and Braziunas, T.F. Atmospheric 14C and century-scale solar oscillations. Nature 338, (1989). 405408.Google Scholar
Stuiver, M., and Reimer, P.J. Extended 14C database and revised CALIB 3.0 14C age calibration program. Radiocarbon 35, (1993). 215230.CrossRefGoogle Scholar
Tartaglione, C.A., Smith, S.R., and O'Brien, J.J. ENSO impact on hurricane landfall probabilities for the Caribbean. Journal of Climate 16, (2003). 29252931.2.0.CO;2>CrossRefGoogle Scholar
Telford, R.J., Heegaard, E., and Birks, H.J.B. The intercept is a poor estimate of a calibrated radiocarbon age. Holocene 14, (2004). 296298.Google Scholar
Thomason, J. M. (2007). "Modern and Fossil Ostracods in the Sediments of Las Lagunas, Dominican Republic.". Unpublished B.A. thesis, University of Tennessee, .Google Scholar
Thompson, L.G., Yao, T., Mosley-Thompson, E., Davis, M.E., Henderson, K.A., and Lin, P.N. A high-resolution millennial record of the South Asian monsoon from Himalayan ice cores. Science 289, (2000). 19161919.Google Scholar
Thompson, L.G., Mosley-Thompson, E., Brecher, H., Davis, M., Leon, B., Les, D., Lin, P.N., Mashiotta, T., and Mountain, K. Abrupt tropical climate change: past and present. Proceedings of the National Academy of Sciences of the United States of America 103, (2006). 1053610543.CrossRefGoogle ScholarPubMed
Trouet, V., Esper, J., Graham, N.E., Baker, A., Scourse, J.D., and Frank, D.C. Persistent positive North Atlantic Oscillation mode dominated the Medieval Climate Anomaly. Science 324, (2009). 7880.Google Scholar
Vare, L.L., Masse, G., Gregory, T.R., Smart, C.W., and Belt, S.T. Sea ice variations in the central Canadian Arctic Archipelago during the Holocene. Quaternary Science Reviews 28, (2009). 13541366.Google Scholar
von Grafenstein, U., Erlernkeuser, H., and Trimborn, P. Oxygen and carbon isotopes in modern fresh-water ostracod valves: assessing vital offsets and autoecological effects of interest for palaeoclimate studies. Palaeogeography Palaeoclimatology Palaeoecology 148, (1999). 133152.Google Scholar
Watanabe, T., Winter, A., and Oba, T. Seasonal changes in sea surface temperature and salinity during the Little Ice Age in the Caribbean Sea deduced from Mg/Ca and O-18/O-16 ratios in corals. Marine Geology 173, (2001). 2135.Google Scholar
Winter, A., Ishioroshi, H., Watanabe, T., Oba, T., and Christy, J. Caribbean sea surface temperatures: two-to-three degrees cooler than present during the Little Ice Age. Geophysical Research Letters 27, (2000). 33653368.Google Scholar
Wood, R.D. Charophytes of North America: a Guide to the Species of Charophytes of North America, Central America, and the West Indies. (1967). Stella's Printing, West Kingston, Rhode Island.Google Scholar
Wood, R.A., and Imahori, K. A Revision of the Characeae. (1964). Weinheim, New York.Google Scholar