Full Text View

Volume 20 Issue 7 (July 2010)

GSA Today

Bookmark and Share

Article, pp. 4-10 | Abstract | PDF (4.7MB)

Resolving uplift of the northern Andes using detrital zircon age signatures

Brian K. Horton1, Mauricio Parra2, Joel E. Saylor2, Junsheng Nie2, Andrés Mora3, Vladimir Torres3, Daniel F. Stockli4, Manfred R. Strecker5

1 Dept. of Geological Sciences and Institute for Geophysics, Jackson School of Geosciences, The University of Texas at Austin, Austin, Texas 78712-0254, USA
2 Dept. of Geological Sciences, Jackson School of Geosciences, The University of Texas at Austin, Austin, Texas 78712-0254, USA
3 Instituto Colombiano del Petróleo, Ecopetrol, Bucaramanga, Colombia
4 Dept. of Geology, University of Kansas, Lawrence, Kansas 66045, USA
5 Institut für Geowissenschaften, Universität Potsdam, 14476 Potsdam, Germany

Table of Contents
Search GoogleScholar for

Search GSA Today


 

Abstract

Uplift of the Eastern Cordillera in the northern Andes has been linked to orographic climate change and genesis of South America’s largest river systems. The timing of initial uplift remains poorly constrained, with most estimates ranging from ca. 60 to ca. 5 Ma. New detrital zircon U-Pb ages from proximal fill of the Llanos foreland basin in Colombia reveal a pronounced mid-Cenozoic shift in provenance from an Amazonian craton source to an Andean fold-thrust belt source. This shift corresponds with changes in detrital zircon (U-Th)/He ages, a conglomeratic unroofing sequence, and a sharp increase in foredeep accumulation rates. These nearly simultaneous changes in zircon age spectra, clast compositions, and sediment accumulation are attributable to latest Oligocene uplift of the eastern flank of the Eastern Cordillera. The timing relationships suggest an early activation of the frontal thrust system, implying a long-term (up to 25 m.y.) cessation of orogenic wedge advance, potentially driven by structural inheritance and/or climate change.

Manuscript received 12 Aug. 2009; accepted 6 Jan. 2010

DOI: 10.1130/GSATG76A.1

INTRODUCTION

Surface uplift of the Eastern Cordillera in the northern Andes has had a profound effect on orographic climate change (Mora et al., 2008), growth of large continental drainage systems (Fig. 1) (Amazon, Orinoco, and Magdalena rivers; Hoorn et al., 1995; Díaz de Gamero, 1996), and biologic evolution of neotropical rainforests (Hooghiemstra and Van der Hammen, 1998; Jaramillo et al., 2006). Most estimates for the onset of uplift along the eastern flank of the Colombian Andes (Fig. 2) range from Paleocene to Pliocene time (Van der Hammen et al., 1973; Dengo and Covey, 1993; Cooper et al., 1995; Bayona et al., 2008; Parra et al., 2009a).

Fig. 1Figure 1

Map of South America showing main river systems (Magdalena, Orinoco, Amazon, and Parana) and Precambrian crustal provinces of the Amazonian craton (after Cordani et al., 2000; Chew et al., 2007).

Fig. 2Figure 2

Shaded relief map showing tectonomorphic provinces of the northern Andes (after Mora et al., 2006) and 12 sample locations (white dots). CC—Central Cordillera; EC—Eastern Cordillera; MV—Magdalena Valley; WC—Western Cordillera.

Initial uplift has proven difficult to constrain by conventional methods. First, recent zircon fission track data provide a minimum age but do not uniquely pinpoint the precise onset of earliest uplift-induced exhumation (Parra et al., 2009b). Second, insights from synorogenic growth strata are commonly limited by inadequate exposure, poor seismic resolution, and minimal variation in stratal dip (e.g., Toro et al., 2004). Third, clastic compositional records of erosional unroofing are hindered by the uniformly high-maturity (quartz-dominated) sand compositions imposed by intense tropical weathering (e.g., Johnsson et al., 1988).

In this study, we utilize U-Pb and (U-Th)/He ages of detrital zircon grains from the Colombian Andes to demonstrate that initial uplift-induced exhumation along the eastern flank of the fold-thrust belt had commenced by ca. 26–23 Ma. Timing relationships revealed by geochronological data coincide with shifts in conglomerate clast compositions and sediment accumulation rates and provide new insights into the pace of orogenic wedge advance.

Geologic Setting

The Eastern Cordillera of Colombia forms a 1–4-km-high orographic barrier separating the intermontane Magdalena Valley from the Llanos foreland basin (Fig. 2). The 100–200-km-wide range is bounded by a frontal thrust system consisting of inverted normal faults and newly formed fold-thrust structures (Cooper et al., 1995; Mora et al., 2006). Following Jurassic–Early Cretaceous rifting, Andean orogenesis began with latest Cretaceous–Paleocene shortening in the Central Cordillera and early foreland basin evolution in the present-day Magdalena Valley and Eastern Cordillera (Cooper et al., 1995; Gómez et al., 2003). The locus of deformation advanced eastward, reaching the western edge of the Eastern Cordillera by middle Eocene–Oligocene time (Restrepo-Pace et al., 2004; Gómez et al., 2005a; Montes et al., 2005; Parra et al., 2009b; Nie et al., 2010).

The Eastern Cordillera consists of Cretaceous quartzose sandstone and mudrock with subordinate Cenozoic, Jurassic, and Paleozoic clastic units capping localized occurrences of crystalline basement. Phanerozoic clastic units are quartz rich (Villamil, 1999), rendering sandstone petrography less effective in addressing provenance history (Johnsson et al., 1988). Prior to Andean uplift, terrigenous clastic sediment was derived from basement exposures to the east, in the Guyana shield of the northern Amazonian craton (Fig. 1) (Cooper et al., 1995). A reversal in sediment transport was triggered by uplift of principally magmatic-arc rocks in the Central Cordillera (Villamil, 1999; Gómez et al., 2005b) followed by uplift and erosional recycling of sedimentary rocks in the Eastern Cordillera (Cooper et al., 1995; Bayona et al., 2008).

The Cretaceous-Cenozoic stratigraphic successions of the Eastern Cordillera and Llanos basin (Fig. 3) consist of clastic units assigned to synrift, postrift, and foreland basin deposition (Cooper et al., 1995). Principally marine facies characterize the 3–8-km-thick Cretaceous succession, including diagnostic glauconitic sandstones in the Une, Chipaque, and Guadalupe units. The overlying 2–3-km-thick Paleogene section contains a range of nonmarine sandstone units that can be correlated from the interior of the Eastern Cordillera into the Llanos basin (e.g., Cacho, Barco, Regadera, and Mirador Formations). The youngest basin fill includes an upward coarsening, 3–5-km-thick Neogene section composed of mudstone, sandstone, and conglomerate of the Carbonera, León, and Guayabo Formations (Fig. 3).

Fig. 3Figure 3

Generalized Mesozoic-Cenozoic stratigraphy of the Eastern Cordillera and Llanos basin, Colombia (after Parra et al., 2009a), showing approximate stratigraphic levels of 12 samples (circles).

U-Pb Geochronology

Methods

Zircon grains from 12 samples of Cenozoic sandstones were separated by standard heavy liquid techniques, selected randomly, and analyzed by laser-ablation–inductively coupled plasma–mass spectrometry. Analyses and associated age calculations followed methods outlined by Chang et al. (2006), utilizing results for zircon standards Peixe (564 ± 4 Ma) and Temora (416.8 ± 1.1 Ma).

We report a total of 1107 U-Pb ages (see supplemental data Table DR11) obtained by analyses that generally yielded <10% discordance, <5% reverse discordance, and <10% uncertainty. Interpreted ages represent 206Pb/238U ages for grains younger than 900 Ma and 207Pb/206Pb ages for grains older than 900 Ma. Results are plotted on relative age probability diagrams (Fig. 4) and normalized such that age-distribution curves for all samples contain the same area. Interpretations are based on age peaks defined by three or more grains.

Fig. 4Figure 4

Detrital zircon U-Pb ages for Cenozoic strata in the proximal Llanos basin (samples 2–6 and 8–12) and axial Eastern Cordillera (samples 1 and 7). Age probability plots arranged in stratigraphic order.

Results and Interpretations

An Eocene sample from the upper Mirador Formation (Fig. 3) in the proximal Llanos basin is characterized by Precambrian age peaks of 1850–1350 Ma (Fig. 4A). In contrast, a sample from the Mirador-equivalent Regadera Formation in the axial Eastern Cordillera shows principally Phanerozoic ages, with peaks at 85–75, 65–55, and 190–170 Ma (Fig. 4A). We ascribe the Llanos age distributions to a dominant eastern source of Proterozoic rocks in the Guyana shield, consistent with previous studies (Cooper et al., 1995; Villamil, 1999; Roure et al., 2003). Age spectra for the axial Eastern Cordillera, however, signify exhumation of Jurassic-Paleogene magmatic-arc rocks from a western source region in the Central Cordillera or Magdalena Valley region (Fig. 2). During Paleocene to Oligocene time, the Eastern Cordillera and proximal Llanos basin likely formed a single integrated basin with sediment supplied from the east and west (Dengo and Covey, 1993; Cooper et al., 1995; Villamil, 1999) and potentially from localized structural highs (Gómez et al., 2005a; Bayona et al., 2008).

Five Oligocene to lower Miocene samples (Fig. 4B) reveal a major provenance shift. A basal sample of the Oligocene C7 member of the Carbonera Formation (Fig. 3) shows exclusively Precambrian age peaks older than 1500 Ma, indicating a cratonic signature. Upsection, the four lower Miocene samples from the C5, C2, and C1 members and an unnamed equivalent unit in the Eastern Cordillera exhibit Precambrian age peaks (1850–1300, 1050–950 Ma) and a collection of Phanerozoic ages (65–45, 90–80, 155–135, 175–170 Ma) that include the first appearance in the Llanos basin of Jurassic-Paleogene zircons, grains that must originate in the west. Although such Phanerozoic zircons could be first-cycle grains from magmatic-arc rocks of the Central Cordillera and Magdalena Valley, the presence of a Grenville age peak at 1050–950 Ma and Paleoproterozoic to Mesoproterozoic grain ages at 1850–1300 Ma (Fig. 4B) suggests recycling of sediments originally derived from the easternmost Andes and/or Guyana shield (e.g., Horton et al., 2010). We attribute the U-Pb ages to initial unroofing of the Upper Cretaceous–Paleogene stratigraphic succession along the axis and eastern flank of the Eastern Cordillera during latest Oligocene–early Miocene time. Unroofing of this cover section produced composite age spectra reflecting recycled contributions from arc-derived Paleogene strata (e.g., Regadera; Fig. 4A), craton-derived Paleogene strata (e.g., Mirador; Fig. 4A), and cra-ton-derived Cretaceous strata.

Five upper Miocene–Pliocene samples from the lower and upper Guayabo Formation (Fig. 3) exhibit age spectra (1850–1300, 1050–950 Ma) comparable to Proterozoic basement in the Guyana shield, with limited Phanerozoic and few Jurassic-Paleogene ages (Fig. 4C). Paleocurrent data and conglomerate clast compositions show that these deposits were derived from western Andean sources rather than the Guyana shield in the east (Parra et al., 2010). We ascribe the U-Pb results to continued unroofing of the Eastern Cordillera, with recycling of principally Cretaceous strata originally sourced from the Guyana shield. Notably, for the upper Miocene–Pliocene upper Guayabo Formation, the conspicuous absence of Jurassic-Paleogene age populations that typified the underlying lower Miocene section (Fig. 4B) rules out the western magmatic arc as a potential source and suggests nearly complete erosional stripping of the arc-derived Paleogene section in the Eastern Cordillera (e.g., Regadera; Fig. 4A). Further, continued erosional recycling of craton-derived Cretaceous strata accounts for the upsection increase in Grenville-aged (1050–950 Ma) detritus at the expense of Jurassic-Paleogene ages (compare Figs. 4B and 4C). Removal of most of the ~2–3-km-thick Paleogene section during early to middle Miocene time suggests an average one-dimensional exhumation rate of ~0.2–0.3 mm/yr, a rate comparable to values estimated from low-temperature thermochronometry (Parra et al., 2009b).

(U-Th)/He Thermochronometry

Zircon (U-Th)/He thermochronometry is an established technique involving a closure temperature of ~180–200 °C (e.g., Reiners, 2005) and a partial retention zone of ~120–180 °C (Stockli, 2005). Detrital (U-Th)/He age determinations were carried out following laboratory procedures described in Biswas et al. (2007). All ages were calculated using Fish Canyon and Durango zircon age standards and alpha-ejection corrections based on morphometric analyses (Farley et al., 1996). Reported age uncertainties reflect the reproducibility of replicate analyses of the two standards, with estimated analytical uncertainties of ~8% (2s) for zircon (U-Th)/He ages (Reiners, 2005; Biswas et al., 2007).

We report 55 (U-Th)/He ages from seven of the aforementioned samples (see supplemental data Table DR2 [footnote 1]). Results are plotted on normalized relative age probability diagrams (Fig. 5). Detrital zircon (U-Th)/He ages for Eocene to mid-Oligocene sandstones of the Mirador and lowermost Carbonera Formations show a strong component of Neoproterozoic ages (11 of 18 grains in the 850–550 Ma range) with subordinate Paleozoic and Jurassic ages (Fig. 5A). In sharp contrast, Miocene-Pliocene sandstones of the middle-upper Carbonera and Guayabo Formations are dominated by Cretaceous-Cenozoic age signatures (28 of 37 grains younger than 150 Ma), with a minor 1000–850 Ma subpopulation of possible Grenville origin (Fig. 5B).

The (U-Th)/He results indicate a substantial shift in detrital age signatures during lower Carbonera deposition that coincides with the provenance shift expressed in the U-Pb data. However, rather than crystallization ages, the detrital zircon (U-Th)/He results provide insight into the integrated cooling history of the sediment source areas. In this case, we interpret the dominantly Precambrian cooling signatures of the Eocene to mid-Oligocene sandstones as the product of extremely long residence time at shallow burial depths (<5–10 km) in the stable cratonic interior of the Guyana shield. The significantly younger (U-Th)/He ages identified in the Miocene-Pliocene sandstones are considered to be the product of Cretaceous-Cenozoic exhumation in the Andean orogenic belt and/or possible contribution from igneous sources. Similar to the U-Pb data, these (U-Th)/He results reveal a pronounced shift in provenance from the ancient Guyana shield in the east to the Andean orogenic belt in the west.

Fig. 5Figure 5

Detrital zircon (U-Th)/He ages for Cenozoic strata in the proximal Llanos basin. Age probability plots arranged in stratigraphic order.

The (U-Th)/He results indicate a substantial shift in detrital age signatures during lower Carbonera deposition that coincides with the provenance shift expressed in the U-Pb data. However, rather than crystallization ages, the detrital zircon (U-Th)/He results provide insight into the integrated cooling history of the sediment source areas. In this case, we interpret the dominantly Precambrian cooling signatures of the Eocene to mid-Oligocene sandstones as the product of extremely long residence time at shallow burial depths (<5–10 km) in the stable cratonic interior of the Guyana shield. The significantly younger (U-Th)/He ages identified in the Miocene-Pliocene sandstones are considered to be the product of Cretaceous-Cenozoic exhumation in the Andean orogenic belt and/or possible contribution from igneous sources. Similar to the U-Pb data, these (U-Th)/He results reveal a pronounced shift in provenance from the ancient Guyana shield in the east to the Andean orogenic belt in the west.

Clast Compositions and Accumulation Rates

The new detrital zircon data are compatible with the unroofing record suggested by conglomerate clast compositions and sediment accumulation rates in the proximal Llanos basin. Clast composition data for the Eocene to upper Miocene succession (Fig. 6A) record the first appearance of large proportions of sandstone clasts followed by increased amounts of distinctive clasts of glauconite-bearing quartzose sandstones (Fig. 6B). Reworked microfossils (pollen, dinoflagellates, and foraminifera; Bayona et al., 2008) suggest a Paleocene or older age for these sandstone clasts. This history indicates late Oligocene–early Miocene unroofing of Paleocene-Eocene sandstones and widespread middle to late Miocene exposure of glauconitic sandstones that are diagnostic of specific mid- to Upper Cretaceous units (Une, Chipaque, and Guadalupe units).

Fig. 6Figure 6

(A) Stratigraphic log of ~4.7-km succession showing approximate levels of 12 samples (circles); (B) corresponding conglomerate clast compositions; and (C) sediment accumulation rates derived from one-dimensional decompacted thicknesses for Eocene-Miocene deposits of the proximal Llanos basin at ~5°N (after Parra et al., 2009a, 2010).

An Eocene to late Miocene sediment accumulation history (Fig. 6C) for the proximal Llanos basin has been constructed on the basis of stratigraphic sections temporally calibrated by detailed palynological assemblages (Parra et al., 2010). The data show a major increase in accumulation rates, starting ca. 26–23 Ma, from ~100 to ~460 m/m.y. Although this stratigraphic transition shows only modest changes in lithofacies or paleo-currents, it corresponds with the abrupt change in detrital zircon age signatures (Figs. 4 and 5) discussed in previous sections.

Discussion and Implications

Coeval shifts in U-Pb and (U-Th)/He zircon age spectra, clast composition, and sediment accumulation rates in the Llanos foreland basin can be attributed to initial uplift of the eastern flank of the Eastern Cordillera (Fig. 2) during latest Oligocene time, consistent with zircon fission-track cooling histories (Parra et al., 2009b). Eastern Cordillera uplift is considered fundamental to the genesis and/or reorganization of the largest drainages in northern South America (Fig. 1). Therefore, initial deformation by ca. 26–23 Ma suggests that precursors of the Amazon, Orinoco, and Magdalena river systems may have originated >10–20 m.y. earlier than envisioned (e.g., Hoorn et al., 1995; Díaz de Gamero, 1996; Campbell et al., 2006). Evidence for protracted drainage histories should be preserved in the depositional records of the respective river deltas and submarine fans (e.g., Dobson et al., 2001; Harris and Mix, 2002), recognizing that changes in sediment accumulation rates could also be the product of climatic effects (Molnar, 2004). We speculate that reduced sediment accumulation rates in the Amazon fan during the early and middle Miocene (Dobson et al., 2001) may be the product of enhanced loading during Eastern Cordillera uplift and a corresponding increase in flexural accommodation and proximal storage of sediment in the Llanos foreland basin (e.g., Bayona et al., 2008; Parra et al., 2009a).

As the principal orographic barrier to easterly air masses, surface uplift of the Eastern Cordillera has influenced the paleogeography and biodiversity of South America’s neotropical rainforests (e.g., Hooghiemstra and Van der Hammen, 1998; Albert et al., 2006). We hypothesize that aridification in the intermontane Magdalena Valley of Colombia (Fig. 2) represents the leeward orographic response to Eastern Cordillera uplift (e.g., Strecker et al., 2007). Although critical threshold elevations were likely reached in late Miocene to Pliocene time (Mora et al., 2008), paleoprecipitation estimates based on mammal faunal assemblages (Kay and Madden, 1997) are consistent with a sediment provenance reversal (Guerrero, 1997) suggesting that initial aridification due to an emerging orographic barrier may have been under way by middle Miocene time.

A further intriguing implication of latest Oligocene uplift along the eastern flank of the Eastern Cordillera is an apparent long-term reduction in the average rate of thrust front advance over the past ~25 m.y., the period of maximum shortening and surface uplift. This seeming contradiction underscores the potential importance of precipitation-driven erosion and inherited mechanical properties on structural evolution of the orogenic wedge in the humid northern Andes.

ACKNOWLEDGMENTS

Funding was provided by Ecopetrol’s Instituto Colombiano del Petróleo as part of a collaborative research agreement with the University of Texas at Austin. M. Parra, A. Mora, and M.R. Strecker also acknowledge funding from the German Academic Exchange Service and Deutsche Forschungs-gemeinschaft. The manuscript benefited from discussions with many Ecopetrol employees and Colombian researchers along with reviews by Isabelle Coutand, James Mortensen, and Stephen Johnston.

REFERENCES CITED

  1. Albert, J.S., Lovejoy, N.R., and Crampton, W.G.R., 2006, Miocene tectonism and the separation of cis- and trans-Andean river basins: Evidence from Neotropical fishes: Journal of South American Earth Sciences, v. 21, p. 14–27.
  2. Bayona, G., Cortés, M., Jaramillo, C., Ojeda, G., Aristizabal, J.J., and Reyes-Harker, A., 2008, An integrated analysis of an orogen-sedimentary basin pair: Latest Cretaceous-Cenozoic evolution of the linked Eastern Cordillera orogen and the Llanos foreland basin of Colombia: GSA Bulletin, v. 120, p. 1171–1197.
  3. Biswas, S., Coutand, I., Grujie, D., Hager, C., Stockli, D., and Grasemann, B., 2007, Exhumation and uplift of the Shillong plateau and its influence on the eastern Himalayas: New constraints from apatite and zircon (U-Th-[Sm])/He and apatite fission track analysis: Tectonics, v. 26, TC6013, doi: 10.1029/2007TC002125.
  4. Campbell, K.E., Frailey, C.D., and Romero-Pittman, L., 2006, The Pan-Amazonian Ucayali Peneplain, late Neogene sedimentation in Amazonia, and the birth of the modern Amazon River system: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 239, p. 166–219.
  5. Chang, Z., Vervoort, J.D., McClelland, W.C., and Knaack, C., 2006, U-Pb dating of zircon by LA-ICP-MS: Geochemistry Geophysics Geosystems, v. 7, Q05009, doi: 10.1029/2005GC001100.
  6. Chew, D.M., Schaltegger, U., Košler, J., Whitehouse, M.J., Gutjahr, M., Spikings, R.A., and Miškovíc, A., 2007, U-Pb geochronologic evidence for the evolution of the Gondwanan margin of the north-central Andes: GSA Bulletin, v. 119, p. 697–711.
  7. Cooper, M.A., Addison, F.T., Álvarez, R., Coral, M., Graham, R.H., Hayward, A.B., Howe, S., Martínez, J., Naar, J., Peńas, R., Pulham, A.J., and Taborda, A., 1995, Basin development and tectonic history of the Llanos basin, Eastern Cordillera, and Middle Magdalena Valley, Colombia: AAPG Bulletin, v. 79, p. 1421–1443.
  8. Cordani, U.G., Sato, K., Teixeira, W., Tassinari, C.G., and Basei, M.A.S., 2000, Crustal evolution of the South American Platform, in Cordani, U.G., Milani, E.J., Thomaz Filho, A., and Campos Neto, M.C., eds., Tectonic Evolution of South America: Rio de Janeiro, 31st International Geological Congress, Institut de Recherché pour le Développement, p. 19–40.
  9. Dengo, C.A., and Covey, M.C., 1993, Structure of the Eastern Cordillera of Colombia: Implications for trap styles and regional tectonics: AAPG Bulletin, v. 77, p. 1315–1337.
  10. Díaz de Gamero, M.L., 1996, The changing course of the Orinoco River during the Neogene: A review: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 123, p. 385–402.
  11. Dobson, D.M., Dickens, G.R., and Rea, D.K., 2001, Terrigenous sediment on Ceara Rise: A Cenozoic record of South American orogeny and erosion: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 165, p. 215–229.
  12. Farley, K.A., Wolf, R., and Silver, L., 1996, The effects of long alpha-stopping distances on (U-Th)/He ages: Geochimica et Cosmochimica Acta, v. 60, p. 4223–4229.
  13. Gómez, E., Jordan, T.E., Allmendinger, R.W., Hegarty, K., Kelley, S., and Heizler, M., 2003, Controls on architecture of the Late Cretaceous to Cenozoic southern Middle Magdalena Valley Basin, Colombia: GSA Bulletin, v. 115, p. 131–147.
  14. Gómez, E., Jordan, T.E., Allmendinger, R.W., and Cardozo, N., 2005a, Development of the Colombian foreland-basin system as a consequence of diachronous exhumation of the northern Andes: GSA Bulletin, v. 117, p. 1272–1292.
  15. Gómez, E., Jordan, T.E., Allmendinger, R.W., Hegarty, K., and Kelley, S., 2005b, Syntectonic Cenozoic sedimentation in the northern middle Magdalena Valley Basin of Colombia and implications for exhumation of the Northern Andes: GSA Bulletin, v. 117, p. 547–569.
  16. Guerrero, J., 1997, Stratigraphy, sedimentary environments, and the Miocene uplift of the Colombian Andes, in Kay, R.F., Madden, R.H., Cifelli, R.L., and Flynn, J.J., eds., Vertebrate Paleontology in the Neotropics: The Miocene Fauna of La Venta, Colombia: Washington D.C., Smithsonian Institution Press, p. 15–42.
  17. Harris, S.E., and Mix, A.C., 2002, Climate and tectonic influences on continental erosion of tropical South America, 0–13 Ma: Geology, v. 30, p. 447–450.
  18. Hooghiemstra, H., and Van der Hammen, T., 1998, Neogene and Quaternary development of the neotropical rain forest: The forest refugia hypothesis, and a literature overview: Earth-Science Reviews, v. 44, p. 147–183.
  19. Hoorn, C., Guerrero, J., Sarmiento, G.A., and Lorente, M.A., 1995, Andean tectonics as a cause for changing drainage patterns in Miocene northern South America: Geology, v. 23, p. 237–240.
  20. Horton, B.K., Saylor, J.E., Nie, J., Mora, A., Parra, M., Reyes-Harker, A., and Stockli, D.F., 2010, Linking sedimentation in the northern Andes to basement configuration, Mesozoic extension, and Cenozoic shortening: Evidence from detrital zircon U-Pb ages, Eastern Cordi­llera, Colombia: GSA Bulletin, v. 122, p. 1423–1442, doi: 10.1130/B30118.1.
  21. Jaramillo, C., Rueda, M., and Mora, G., 2006, Cenozoic plant diversity in the Neotropics: Science, v. 311, p. 1893–1896.
  22. Johnsson, M.J., Stallard, R.F., and Meade, R.H., 1988, First-cycle quartz arenites in the Orinoco River basin, Venezuela and Colombia: Journal of Geology, v. 96, p. 263–277.
  23. Kay, R.F., and Madden, R.H., 1997, Mammals and rainfall: Paleoecology of the middle Miocene at La Venta (Colombia, South America): Journal of Human Evolution, v. 32, p. 161–199.
  24. Molnar, P., 2004, Late Cenozoic increase in accumulation rates of terrestrial sediment: How might climate change have affected erosion rates?: Annual Review of Earth and Planetary Sciences, v. 32, p. 67–89.
  25. Montes, C., Hatcher, R.D., Jr., and Restrepo-Pace, P.A., 2005, Tectonic reconstruction of the northern Andean blocks: Oblique convergence and rotations derived from the kinematics of the Piedras-Girardot area, Colombia: Tectonophysics, v. 399, p. 221–250.
  26. Mora, A., Parra, M., Strecker, M.R., Kammer, A., Dimaté, C., and Rodríguez, F., 2006, Cenozoic contractional reactivation of Mesozoic extensional structures in the Eastern Cordillera of Colombia: Tectonics, v. 25, TC2010, doi: 10.1029/2005TC001854.
  27. Mora, A., Parra, M., Strecker, M.R., Sobel, E.R., Hooghiemstra, H., Torres, V., and Vallejo-Jaramillo, J., 2008, Climatic forcing of asymmetric orogenic evolution in the Eastern Cordillera of Colombia: GSA Bulletin, v. 120, p. 930–949.
  28. Nie, J., Horton, B.K., Mora, A., Saylor, J.E., Housh, T.B., Rubiano, J., and Naranjo, J., 2010, Tracking exhumation of Andean ranges bounding the Middle Magdalena Valley basin, Colombia: Geology, v. 5, p. 451–454.
  29. Parra, M., Mora, A., Jaramillo, C., Strecker, M.R., Sobel, E.R., Quiroz, L., Rueda, M., and Torres, V., 2009a, Orogenic wedge advance in the northern Andes: Evidence from the Oligocene-Miocene sedimentary record of the Medina Basin, Eastern Cordillera, Colombia: GSA Bulletin, v. 121, p. 780–800.
  30. Parra, M., Mora, A., Sobel, E.R., Strecker, M.R., and González, R., 2009b, Episodic orogenic front migration in the northern Andes: Constraints from low-temperature thermochronology in the Eastern Cordillera, Colombia: Tectonics, v. 28, TC4004, doi: 10.1029/2008TC002423.
  31. Parra, M., Mora, A., Jaramillo, C., Torres, V., Zeilinger, G., and Strecker, M.R., 2010, Tectonic controls on Cenozoic foreland basin development in the northeastern Andes, Colombia: Basin Research, doi: 10.1111/j.1365-2117.2009.00459.x.
  32. Reiners, P., 2005, Zircon (U-Th)/He thermochronometry, in Reiners, P.W., and Ehlers, T.A., eds., Low-temperature thermochronology: Techniques, interpretations, and applications: Reviews in Mineralogy and Geochemistry, v. 58, p. 151–179.
  33. Restrepo-Pace, P.A., Colmenares, F., Higuera, C., and Mayorga, M., 2004, A fold-and-thrust belt along the western flank of the Eastern Cordi­llera of Colombia. Style, kinematics, and timing constraints derived from seismic data and detailed surface mapping, in McClay, K.R., ed., Thrust Tectonics and Hydrocarbon Systems: American Association of Petroleum Geologists Memoir 82, p. 598–613.
  34. Roure, F., Bordas-Le Floch, N., Toro, J., Aubourg, C., Guilhaumou, N., Hernández, H., Lecornec-Lance, S., Rivero, C., Robion, P., and Sassi, W., 2003, Petroleum systems and reservoir appraisal in the sub-Andean basins (eastern Venezuela and eastern Colombian foothills), in Bartolini, C., Buffler, R.T., and Blickwede, J., eds., The Circum-Gulf of Mexico and the Caribbean: Hydrocarbon Habitats, Basin Formation, and Plate Tectonics: American Association of Petroleum Geologists Memoir 79, p. 750–775.
  35. Stockli, D.F., 2005, Application of low-temperature thermochronometry to extensional tectonic settings, in Reiners, P.W., and Ehlers, T.A., eds., Low-Temperature Thermochronology: Techniques, Interpretations, and Applications: Reviews in Mineralogy and Geochemistry, v. 58, p. 411–448.
  36. Strecker, M.R., Alonso, R.N., Bookhagen, B., Carrapa, B., Hilley, G.E., Sobel, E.R., and Trauth, M.H., 2007, Tectonics and climate of the southern central Andes: Annual Review of Earth and Planetary Sciences, v. 35, p. 747–787.
  37. Toro, J., Roure, F., Bordas-Le Floch, N., Le Cornec-Lance, S., and Sassi, W., 2004, Thermal and kinematic evolution of the Eastern Cordi­llera fold-and-thrust-belt, Colombia, in Swennen, R., Roure, F., and Granath, J.W., eds., Deformation, Fluid Flow, and Reservoir Appraisal in Foreland Fold and Thrust Belts: American Association of Petroleum Geologists Hedberg Series, v. 1, p. 79–115.
  38. Van der Hammen, T., Werner, J.H., and Van Dommelen, H., 1973, Palynological record of the upheaval of the northern Andes: A study of the Pliocene and lower Quaternary of the Colombian Eastern Cordillera and the early evolution of its high-Andean biota: Reviews of Palaeobotany and Palynology, v. 16, p. 1–122.
  39. Villamil, T., 1999, Campanian-Miocene tectonostratigraphy, depocenter evolution and basin development of Colombia and western Venezuela: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 153, p. 239–275.

top