Hostname: page-component-848d4c4894-4hhp2 Total loading time: 0 Render date: 2024-05-21T09:19:49.848Z Has data issue: false hasContentIssue false

15,000-yr Pollen Record of Vegetation change in the High Altitude Tropical Andes at Laguna Verde Alta, Venezuela

Published online by Cambridge University Press:  20 January 2017

Valentí Rull
Affiliation:
Departament de Biologia Animal, Biologia Vegetal i Ecologia, Universitat Autònoma de Barcelona, C1-215, Bellaterra, 08193 Barcelona, Spain
Mark B. Abbott
Affiliation:
Department of Geology and Planetary Science, University of Pittsburgh, Pittsburgh, PA 15260-3332, USA
Pratigya J. Polissar
Affiliation:
Department of Geosciences, University of Massachusetts, Amherst, MA 01003-5820, USA
Alexander P. Wolfe
Affiliation:
Department of Earth and Atmospheric Sciences, University of Alberta, Edmonton, Alberta, Canada AB T6G 2E3
Maximiliano Bezada
Affiliation:
Department of Earth Sciences, Universidad Pedagógica Experimental Libertador, Caracas, Venezuela
Raymond S. Bradley
Affiliation:
Department of Geosciences, University of Massachusetts, Amherst, MA 01003-5820, USA

Abstract

Pollen analysis of sediments from a high-altitude (4215 m), Neotropical (9°N) Andean lake was conducted in order to reconstruct local and regional vegetation dynamics since deglaciation. Although deglaciation commenced ∼15,500 cal yr B.P., the area around the Laguna Verde Alta (LVA) remained a periglacial desert, practically unvegetated, until about 11,000 cal yr B.P. At this time, a lycopod assemblage bearing no modern analog colonized the superpáramo. Although this community persisted until ∼6000 cal yr B.P., it began to decline somewhat earlier, in synchrony with cooling following the Holocene thermal maximum of the Northern Hemisphere. At this time, the pioneer assemblage was replaced by a low-diversity superpáramo community that became established ∼9000 cal yr B.P. This replacement coincides with regional declines in temperature and/or available moisture. Modern, more diverse superpáramo assemblages were not established until ∼4600 cal yr B.P., and were accompanied by a dramatic decline in Alnus, probably the result of factors associated with climate, humans, or both. Pollen influx from upper Andean forests is remarkably higher than expected during the Late Glacial and early to middle Holocene, especially between 14,000 and 12,600 cal yr B.P., when unparalleled high values are recorded. We propose that intensification of upslope orographic winds transported lower elevation forest pollen to the superpáramo, causing the apparent increase in tree pollen at high altitude. The association between increased forest pollen and summer insolation at this time suggests a causal link; however, further work is needed to clarify this relationship.

Type
Special issue articles
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

Bennett, K.D. (1996). Determination of the number of zones in a biostratigraphical sequence. New Phytologist 132, 155170.Google Scholar
Bennett, K.D. (2002). Documentation for psimpoll 4.10 and pscomb 1.03. C Programs for Plotting Pollen Diagrams and Analysing Pollen Data. University of Cambridge, Cambridge.125 pp.Google Scholar
Bennett, K.D., and Willis, K.J. (2001). Pollen. Smol, J.P., Birks, H.J.B., and Last, W.M. Tracking Environmental Change using Lake Sediments. Terrestrial, Algal, and Siliceous Indicators vol. 3, Kluwer, Dordrecht.532.Google Scholar
Berger, A., and Loutre, M.F. (1991). Insolation values for the climate of the last 10 million years. Quaternary Sciences Reviews 10, 297317.Google Scholar
Birks, H.J.B., and Birks, H.H. (1980). Quaternary Palaeoecology. Arnold, London.289 pp.Google Scholar
Bradbury, J.P., Leyden, B., Salgado-Labouriau, M.L., Lewis, W.M., Schubert, C., Binford, M.W., Frey, D.G., Whitehead, D.R., and Weibezahn, F.H. (1981). Late Quaternary environmental history of Lake Valencia. Science 214, 12991305.CrossRefGoogle ScholarPubMed
Bradley, R.S., Yuretich, R.F., Salgado-Labouriau, M.L., and Weingarten, B. (1985). Late Quaternary palaeoenvironmental reconstruction using lake sediments from the Venezuelan Andes, preliminary results. Zeischrift für Gletscherkunde und Glazialgeologie 21, 97106.Google Scholar
Cleef, A. (1980). Secuencia altitudinal de la vegetación de los páramos de la Cordillera Oriental de Colombia. Revista del Instituto Geográfico Agustín Codazzi 5067.Google Scholar
Cleveland, W.S. (1994). The Elements of Graphing Data. Hobart, Summit.297 pp.Google Scholar
Curtis, J.H., Brenner, M., and Hodell, D.A. (1999). Climate change in the Lake Valencia Basin, Venezuela, 12600 cal yr BP to present. The Holocene 9, 609619.CrossRefGoogle Scholar
Grabandt, R.A.J. (1980). Pollen rain in relation to arboreal vegetation in the Colombian Cordillera Oriental. Review of Palaeobotany and Palynology 29, 65147.Google Scholar
Grabandt, R.A.J., and Nieuwland, J. (1985). Pollen rain in relation to páramo vegetation in the Colombian Cordillera Oriental. The Quaternary of Colombia 11, 1110.Google Scholar
Hansen, B.C.S. (1995). A review of late glacial pollen records from Ecuador and Peru with reference to the Younger Dryas event. Quaternary Science Reviews 14, 853865.CrossRefGoogle Scholar
Hansen, B.C.S., and Rodbell, D.T. (1995). A late-glacial/Holocene pollen record from the eastern Andes of Peru. Quaternary Research 44, 216227.Google Scholar
Hansen, B.C.S., Rodbell, D.T., Seltzer, G.O., León, B., Young, K.R., and Abbott, M. (2003). Late-glacial and Holocene vegetational history from two sites in the western Cordillera of southwestern Ecuador. Palaeogeography, Palaeoclimatology, Palaeoecology 194, 79108.Google Scholar
Haug, G., Hughen, K., Sigman, D., Petrson, L., and Röhl, U. (2001). Southward migration of the intertropical convergence zone during the Holocene. Science 293, 13041307.Google Scholar
Hooghiemstra, H. (1984). Vegetational and climatic history of the High Plain of Bogotá, Colombia, A continuous record of the last 3.5 million years. Dissertationes Botanicae 79. J. Cramer, Vaduz.Google Scholar
Hughen, K.A., Eglinton, T.I., Xu, L., and Makou, M. (2004). Abrupt tropical vegetation response to rapid climate changes. Science 304, 19551959.CrossRefGoogle ScholarPubMed
Iriondo, M.H. (1997). Models of deposition of loess and loessoids in the upper Quaternary of South America. Journal of South American Earth Sciences 10, 7179.Google Scholar
Leyden, B.W. (1985). Late Quaternary aridity and Holocene moisture fluctuations in the Lake Valencia Basin, Venezuela. Ecology 66, 12791295.CrossRefGoogle Scholar
Luteyn, J.L. (1999). Páramos. A checklist of plant diversity, geographical distribution and botanical literature. Memoirs of the New York Botanical Garden 84, New York Botanical Garden, New York.278 pp.Google Scholar
Marchant, R., and Hooghiemstra, H. (2004). Rapid environmental change in African and South American tropics around 4000 years before present: a review. Earth-Science Reviews 66, 217260.Google Scholar
Marchant, R., Almeida, L., Behling, H., Berrio, J.C., Bush, M., Cleef, A., Duivenvoorden, J., Kappelle, M., De Oliveira, P., Teixeira, A., Lozano-García, S., Hooghiemstra, H., Ledru, M.-P., Ludlow-Wiechers, B., Markgraf, V., Mancini, V., Páez, M., Prieto, A., Rangel, O., and Salgado-Labouriau, M.L. (2002a). )Distribution and ecology of parent taxa of pollen lodged within the Latin American Pollen Database. Review of Palaeobotany and Palaeoecology 121, 175.Google Scholar
Marchant, R., Behling, H., Berrío, J.C., Cleef, A., Duivenvoorden, J., Hooghiemstra, H., Kuhry, P., Melief, B., Schreve-Brinkman, E., van Geel, B., van der Hammen, T., van Reenen, G., and Wille, M. (2002b). )Pollen-based biome reconstructions fpr Colombia at 3000, 6000, 9000, 12 000, 15 000 and 18 000 14C yr ago: Late Quaternary tropical vegetation dynamics. Journal of Quaternary Science 17, 113129.Google Scholar
McGregor, G.R., and Nieuwolt, S. (1998). Tropical Climatology. An Introduction to the Climates of the Low Latitudes. Wiley, Chichester.339 pp.Google Scholar
Monasterio, M. (1979). El páramo desértico en el altiandino de Venezuela. Salgado-Labouriau, M.L. El Medio Ambiente Páramo. CEA-IVIC, Caracas.117146.Google Scholar
Monasterio, M. (1980). Las formaciones vegetales de los páramos de Venezuela. Monasterio, M. Estudios Ecológicos En Los Páramos Andinos. Universidad de Los Andes, Mérida.93158.Google Scholar
Monasterio, M., and Reyes, S. (1980). Diversidad ambiental y variación de la vegetación en los páramos de los Andes venezolanos. Monasterio, M. Estudios Ecológicos En Los Páramos Andinos. Universidad de Los Andes, Mérida.4791.Google Scholar
Murillo, M.T., and Bless, M.J.M. (1974). Spores of recent Colombian pteridophyta: I. Trilete spores. Review of Palaeobotany and Palynology 18, 223269.Google Scholar
Murillo, M.T., and Bless, M.J.M. (1978). Spores of recent Colombian pteridophyta: II. Monolete spores. Review of Palaeobotany and Palynology 25, 319365.Google Scholar
Paduano, G.M., Bush, M.B., Baker, P.A., Fritz, Sh.C., and Seltzer, G.O. (2003). A vegetation and fire history of Lake Titicaca since the Last Glacial Maximum. Palaeogeography, Palaeoeclimatology, Palaeoecology 194, 259279.Google Scholar
Rull, V. (1987). A note on pollen counting in palaeoecology. Pollen et Spores 29, 471480.Google Scholar
Rull, V. (1996). Late Pleistocene and Holocene climates of Venezuela. Quaternary International 31, 8594.Google Scholar
Rull, V. (1998). Palaeoecology of pleniglacial sediments from the Venezuelan Andes. Palynological record of El Caballo stadial, sedimentation rates and glacier retreat. Review of Palaeobotany and Palynology 99, 95114.Google Scholar
Rull, V. (1999). Palaeoclimatology and sea-level history in Venezuela. New data, land-sea correlations, and proposal for future studies in the frame of the IGBP-PAGES Project. Interciencia 24, 92101.Google Scholar
Rull, V., Bezada, M., and Mahaney, W.C. (1999). The Midle-Wisconsin ‘El Pedregal’ interstadial in the Venezuelan Andes, palynological record. Current Research in the Pleistocene 16, 111113.Google Scholar
Salgado-Labouriau, M.L. (1979). Modern pollen deposition in the Venezuelan Andes. Grana 18, 5368.Google Scholar
Salgado-Labouriau, M.L. (1980). A pollen diagram of the Pleistocene-Holocene boundary of Lake Valencia, Venezuela. Review of Palaeobotany and Palynology 30, 297312.Google Scholar
Salgado-Labouriau, M.L. (1984). Late Quaternary palynological studies in the Venezuelan Andes. Erdwissenschaftliche Forschung 18, 279293.Google Scholar
Salgado-Labouriau, M.L. (1988). Sequence of colonization by plants in the Venezuelan Andes after the last Pleistocene glaciation. Journal of Palynology 23/24, 189204.Google Scholar
Salgado-Labouriau, M.L. (1989). Late Quaternary climatic oscillations in the Venezuelan Andes. Biology International 18, 1214.Google Scholar
Salgado-Labouriau, M.L., and Rull, V. (1986). A method of introducing exotic pollen for paleoecological analysis of sediments. Review of Palaeobotany and Palynology 47, 97103.Google Scholar
Salgado-Labouriau, M.L., Rull, V., Schubert, C., and Valastro, S. (1988). The establishment of vegetation after Late Pleistocene deglaciation in the Páramo de Miranda, Venezuelan Andes. Review of Palaeobotany and Palynology 55, 517.Google Scholar
Salgado-Labouriau, M.L., Bradley, R.S., Yuretich, R.F., and Weingarten, B. (1992). Palaeoecological analysis of the sediments of lake Mucubají, Venezuelan Andes. Journal of Biogeography 19, 317327.Google Scholar
Schubert, C. (1979). La zona del páramo: Morfología glacial y periglacial de los Andes de Venezuela. Salgado-Labouriau, M.L. El Medio Ambiente Páramo. CEA-IVIC, Caracas.1127.Google Scholar
Stuiver, M., Reimer, P.J., Bard, E., Beck, J.W., Burr, G.S., Hughen, K.A., Kromer, B., McCormac, F.G., van der Plicht, J., and Spurk, M. (1998a). )INTCAL98 Radiocarbon age calibration 24,000-0 cal BP. Radiocarbon 40, 10411083.Google Scholar
Stuiver, M., Reimer, P.J., and Braziunas, T.F. (1998b). )High-precision radiocarbon age calibration for terrestrial and marine samples. Radiocarbon 40, 11271151.Google Scholar
Tryon, A.F., and Lugardon, B. (1991). Spores of the Pteridophyta. Springer, New York.648 pp.Google Scholar
Van der Hammen, T., and González, E. (1960). Upper Pleistocene and Holocene climate and vegetation of the ‘Sabana de Bogotá’ (Colombia, South America). Leidse Geologische Mededelingen 25, 261315.Google Scholar
Van der Hammen, T., and Hooghiemstra, H. (1995). The El Abra Stadial, a Younger Dryas equivalent in Colombia. Quaternary Science Reviews 14, 841851.Google Scholar
Wagner, E. (1979). Arqueología de los Andes venezolanos. Los páramos y la Tierra Fría. Salgado-Labouriau, M.L. El Medio Ambiente Páramo. Centro de Estudios Avanzados-IVIC, Caracas.207218.Google Scholar
Weingarten, B., and Salgado-Labouriau, M.L., Yuretich, R.F., and Bradley, R. (1991). Late Quaternary Environmental History of the Venezuelan Andes. Yuretich, R.F. Late Quaternary climatic fluctuations of the Venezuelan Andes. Department of Geology and Geography Contribution, University of Massachusetts, Amherst.6394.65Google Scholar
Weng, Ch., Bush, M.B., and Chepstow-Lusty, A.J. (2004). Holocene changes of andean alder (Alnus acuminata) in highland Ecuador and Peru. Journal of Quaternary Science 19, 685691.Google Scholar
Wright, H.E., Mann, D.H., and Glaser, pp.H. (1984). Piston corers for peat and lake sediments. Ecology 65, 657659.Google Scholar