You are on page 1of 23

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/283723023

Application of detrital zircon U-Pb geochronology to surface and subsurface


correlations of provenance, paleodrainage, and tectonics of the Middle
Magdalena Valley Basin of Colombi...

Article  in  Geosphere · December 2015


DOI: 10.1130/GES01251.1

CITATIONS READS

79 1,699

7 authors, including:

Brian K. Horton Victor M Caballero


University of Texas at Austin Instituto Colombiano del Petroleo
253 PUBLICATIONS   10,434 CITATIONS    44 PUBLICATIONS   987 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Permo-Triassic sedimentation and salt tectonics in the northern Peruvian fold-and-thrust belt View project

Petroleum Tectonic of Fold and Thrust Belts View project

All content following this page was uploaded by Brian K. Horton on 15 December 2018.

The user has requested enhancement of the downloaded file.


Research Paper

GEOSPHERE Application of detrital zircon U-Pb geochronology to surface


and subsurface correlations of provenance, paleodrainage, and
GEOSPHERE; v. 11, no. 6
tectonics of the Middle Magdalena Valley Basin of Colombia
doi:10.1130/GES01251.1
Brian K. Horton1, Veronica J. Anderson1, Victor Caballero2, Joel E. Saylor 3, Junsheng Nie4, Mauricio Parra2,5, and Andrés Mora2
10 figures; 2 supplemental files 1
Institute for Geophysics and Department of Geological Sciences, Jackson School of Geosciences, University of Texas at Austin, Austin, Texas 78712, USA
2
Ecopetrol, Instituto Colombiano del Petróleo, Bucaramanga, Colombia
3
Department of Earth and Atmospheric Sciences, University of Houston, Houston, Texas 77204, USA
CORRESPONDENCE: horton@jsg.utexas.edu 4
MOE Key Laboratory of Western China’s Environmental Systems, College of Earth and Environmental Sciences, Lanzhou University, Lanzhou 73000, China
5
Institute of Energy and Environment, University of São Paulo, 05508-010 São Paulo, SP, Brazil
CITATION:  Horton, B.K., Anderson, V.J., Caballero,
V., Saylor, J.E., Nie, J., Parra, M., and Mora, A.,
2015, Application of detrital zircon U-Pb geochronol-
ogy to surface and subsurface correlations of prov- ABSTRACT Provenance shifts of mid-Paleocene and latest Eocene–earliest Oligocene
enance, paleodrainage, and tectonics of the Middle
age are consistent with incipient uplift of the flanking Central Cordillera and
Magdalena Valley Basin of Colombia: Geosphere,
v. 11, no. 6, p. 1790–1811, doi:10.1130/GES01251.1. Detrital zircon U-Pb geochronology has been used extensively to develop Eastern Cordillera, respectively. However, a well-documented phase of latest
provenance histories for surface outcrops of key stratigraphic localities within Paleocene–middle Eocene beveling of basement uplifts in the Middle Magda-
Received 17 August 2015 sedimentary basins. However, many basins lack sufficiently continuous and lena Valley Basin appears to be largely aliased in the detrital record. Moreover,
Revision received 13 September 2015 widespread exposures of complete successions to evaluate proposed long- despite the proximity of the magmatic arc, there is insufficient syndeposi-
Accepted 2 October 2015
term tectonic histories, stratigraphic correlations, and paleodrainage patterns tional evidence for a proposed Paleogene pulse of magmatism and, in this
Published online 12 November 2015
within individual basins. Here, we demonstrate the utility of subsurface detri- case, limited utility of U-Pb ages in pinpointing precise depositional (strati-
tal zircon U-Pb analysis by integrating ages from three key wells (21 subsur- graphic) ages.
face samples) with previously reported data from six exposed intervals (90 U-Pb age spectra for Oligocene through Pliocene basin fill underscore
surface samples) within a single basin. Samples from the 5–10-km-thick clastic complex along-strike (north-south) and cross-strike (east-west) variations re-
successions span several structural blocks over an ~300 × 50 km swath of flective of compartmentalized transverse deposystems demarcated by point-
the Middle Magdalena Valley Basin, a north-trending intermontane basin in source contributions from the Central Cordillera and Eastern Cordillera. The
the northern Andes of Colombia. Available U-Pb age distributions for mod- late Miocene appearance of 100–0 Ma grains and a regional switch to broad,
ern rivers highlight the distinctive signatures of several competing sediment multimodal age distributions suggest the initial integration of the longitu-
sources, including two major contiguous ranges (Central Cordillera and East- dinal proto–Magdalena River, linking the Middle Magdalena Valley Basin
ern Cordillera) and two localized block uplifts (Santander Massif and San with southern headwaters in the Upper Magdalena Valley and likely driving
­Lucas range). U-Pb results from Jurassic through Neogene stratigraphic units increased sedimentation rates farther north in the offshore Magdalena sub­
spanning the nine surface and subsurface sites, including several type locali- marine fan of the southern Caribbean margin.
ties, enable comparisons of provenance shifts at specific sites and spatial vari-
ations among key stratigraphic intervals across multiple sites.
Distinctive age populations for the Andean magmatic arc, retroarc fold- INTRODUCTION
thrust belt, and South American craton facilitate correlation of stratigraphic
units and reconstruction of the long-term provenance and tectonic evolution Although advances in detrital zircon U-Pb geochronology have fueled a re-
of the Middle Magdalena Valley Basin. Nearly all surface and subsurface lo- surgence in sediment provenance studies, subsurface and intrabasinal applica-
calities show up-section changes in age spectra consistent with (1) Jurassic tions remain limited. Rapid data acquisition through laser-ablation–­inductively
growth of extensional subbasins fed by local igneous sources, (2) Cretaceous coupled plasma–mass spectrometry (LA-ICP-MS) has propelled U-Pb analyses
deposition in an extensive postrift setting, and (3) protracted Cenozoic growth of sand-sized zircon grains to the forefront of provenance studies seeking to
of basin-bounding ranges during Andean crustal shortening. Subsurface sam- discriminate among potential source regions (e.g., Dickinson and Gehrels,
ples augment surface samples, highlighting their utility in developing regional 2008; Nie et al., 2012; Gehrels, 2014). Further applications for sedimentary
For permission to copy, contact Copyright source-to-sink relationships, the timing of paleodrainage integration, and tec- ­basin analysis include assessments of: (1) the relative volumetric contribu-
Permissions, GSA, or editing@geosociety.org. tonic reconstructions. tions from known sources in modern watersheds (Saylor et al., 2013); (2) the
© 2015 Geological Society of America

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1790
Research Paper

c­ omplex mixing and downstream dilution of source signals within modern Silva et al., 2013; Bayona et al., 2013; Reyes-Harker et al., 2015). Alternative sce-
­rivers (Amidon et al., 2005; Link et al., 2005; Zhang et al., 2012; He et al., 2014); narios propose a north-flowing proto–Magdalena River developing in middle
(3) large-scale paleodrainage patterns, including the onset and evolution of to late Miocene time (Hoorn et al., 1995, 2010). Evolution of the Middle Magda­
major paleorivers (Davis et al., 2010; Dickinson et al., 2012; Mackey et al., 2012; lena Valley Basin drainage systems and establishment of a through-going
Blum and Pecha, 2014); and (4) absolute ages of stratigraphic units (Fildani Magdalena River also directly affected sediment accumulation and offshore
et al., 2003; DeCelles et al., 2007; Dickinson and Gehrels, 2009; Horton et al., hydrocarbon prospectivity of the Magdalena delta in the Caribbean Sea.
2015). Although accurate ages for syndepositional volcanic zircons clearly im- Further uncertainties center on potential basin responses to the important re-
prove chronostratigraphic correlations, U-Pb geochronology may provide a gional phases of Mesozoic extension and Cenozoic shortening in the northern-
viable correlation tool on the basis of comparable age distributions (e.g., Rain- most Andes, as well as critical transitions involving focused basement uplift
bird et al., 2007; Lawton et al., 2010; Beranek et al., 2013; Lewis and Sircombe, and punctuated magmatism (e.g., Colletta et al., 1990; Dengo and Covey, 1993;
2013), an approach that could considerably enhance subsurface investigations. Gómez et al., 2003, 2005a, 2005b; Horton et al., 2010a; Bayona et al., 2012; Parra
Despite the wide range of applications, complexities over multiple scales et al., 2012; Nie et al., 2012; Saylor et al., 2012a, 2012b; Caballero et al., 2013a,
highlight various difficulties in pinpointing sediment source regions, recon- 2013b; Reyes-Harker et al., 2015). Here, we present U-Pb ages of detrital zircons for
structing paleodrainage patterns, and correlating stratigraphic units. Whereas 21 subsurface samples of Mesozoic–Cenozoic basin fill, and we integrate these
most basin-scale studies focus on temporal provenance shifts registered within data with results from 83 surface outcrop samples and seven modern river
key stratigraphic sections, uncertainties persist over the effectiveness of signal samples to help correlate the provenance and tectonic histories of surface and
transmission from source to sink. These issues can include (1) uneven con- sub­surface basin fill.
tributions from source areas, (2) storage, buffering, or recycling within drain-
age systems, and (3) potential sequestration due to intrabasinal variations in
hydrodynamics or depositional environments (e.g., Métivier and Gaudemer, GEOLOGIC AND STRATIGRAPHIC FRAMEWORK
1999; DeGraaff-Surpless et al., 2003; Allen, 2008; Lawrence et al., 2011; Saylor
et al., 2013; Latrubesse, 2015). In river-dominated basins, many of these po- The Middle Magdalena Valley Basin is a long-lived sedimentary basin that
tential problems can be minimized by sampling a range of depositional sub­ has recorded uplift and exhumation of the major ranges and block uplifts of the
envi­ron­ments over a sufficiently large region (e.g., Potter, 1978; Ingersoll, 1990; northernmost Andes at 4°N–7°N (Fig. 1). Modern sediment is delivered to
Ingersoll et al., 1993). Nevertheless, few investigations have explored both the the narrow intermontane Middle Magdalena Valley Basin by transverse rivers
temporal and spatial variations in U-Pb provenance signatures for both surface and alluvial fans from the flanking Central Cordillera and Eastern Cordillera,
and subsurface deposits within a single sedimentary basin. with the Magdalena River flowing longitudinally northward along the basin
In this study, we explore an ~150 m.y. provenance history for surface and axis into the Caribbean Sea. Additional sediment sources include the isolated
subsurface samples from a single elongate basin exhibiting important along- San Lucas range and Santander Massif, which form block uplifts at the north-
strike, cross-strike, and proximal-to-distal variations. The Middle Magdalena ern terminations of the Central and Eastern Cordilleras, respectively.
Valley Basin of Colombia is a narrow intermontane basin spanning an area of Although the 5–10-km-thick clastic fill of the elongate Middle Magdalena
30,000 km2 within the northernmost Andes (Fig. 1). Depositional systems in the Valley Basin now defines a 20–100-km-wide by 500-km-long swath, the Meso­
north-trending basin are dictated locally by transverse sources derived from zoic–Cenozoic succession contains a composite history involving a more-­
the two major flanking mountain ranges, the Central Cordillera and Eastern expansive early Cenozoic foreland basin (Gómez et al., 2003, 2005b; Caballero
Cordillera, and regionally by the longitudinal Magdalena River, which today et al., 2010, 2013a, 2013b; Moreno et al., 2011) and isolated Mesozoic exten-
flows ~1500 km along the basin axis to the Caribbean coast. The Middle Mag- sional subbasins (Cooper et al., 1995; Sarmiento-Rojas et al., 2006; Kammer
dalena Valley Basin is a prolific hydrocarbon basin, with ~100 yr of explora- and Sánchez, 2006; Mora et al., 2013; Tesón et al., 2013). To the west, the
tion efforts. Nevertheless, despite hundreds of wells, problems of stratigraphic north-trending Central Cordillera is mostly composed of Mesozoic–early Ceno­
correlation are highlighted by complex spatial variations in Mesozoic–Ceno- zoic igneous rocks of the Andean magmatic arc unroofed during Mesozoic
zoic depositional systems (Morales, 1958; Van Houten and Travis, 1968; Van extension and Cenozoic shortening. In contrast, to the east, the north-north-
Houten, 1976; Butler and Schamel, 1988; Schamel, 1991; Cooper et al., 1995; east–trending Eastern Cordillera consists of the Andean retroarc fold-thrust
Ramon and Rosero, 2006; Naranjo-Vesga et al., 2013). belt, dominated by Cretaceous sedimentary rocks that, along with localized
These complexities are reflected in competing models for paleodrainage basement uplifts, were erosionally recycled as the Middle Magdalena Valley
and basin evolution in the Middle Magdalena Valley Basin. The Paleogene Basin transitioned from a broad foreland basin to an intermontane hinterland
­Middle Magdalena Valley Basin has been envisioned as the locus of either basin (Caballero et al., 2010, 2013a; Moreno et al., 2011; Saylor et al., 2011; Hor-
northward axial paleodrainages or eastward transverse paleorivers (Nie et al., ton, 2012; Parra et al., 2012; Sánchez et al., 2012; Silva et al., 2013; Moreno et al.,
2010, 2012; Moreno et al., 2011; Saylor et al., 2011; Caballero et al., 2013a, 2013b; 2013; Tesón et al., 2013).

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1791
Research Paper

modern river
77°W 75°W 73°W

in
sand sample
Santa Marta

o Bas
Lower subsurface
sandstone
CARIBBEAN Magdalena Massif sample

caib
PLATE Valley surface
9°N sandstone
sample

Mara
syncline
anticline
thrust fault
strike-slip fault

San Santander normal fault


city
Lucas Massif
range Cagui well
stratigraphic
profile
Rio Santo
Domingo
Nuevo Mundo syncline (W) watershed
boundary
Rio Nechi

Cocuyo well Nuevo Mundo syncline (E)

Guane well

EY
ILLER
Rio Umbala Rio Manco Figure 1. Geologic map of the northern-

CORDILLERA

LL
La Salina footwall most Andes of Colombia (after Gómez
Tapias et al., 2007) showing nine study

2
VA
ure
localities (six surface, three subsurface) of

CORD
the Middle Magdalena Valley Basin and

Fig
A
seven modern sand samples from river

EN
drainage outlets of watersheds that span
Rio Nare Opón syncline diverse geologic units in highland portions

RA
AL
6°N of the Central Cordillera, San Lucas range,

LE
Rio Ermitaño syncline

GD
RAL
Santander Massif, and Eastern Cordillera.

IL
MA

RD
CENT
WESTERN

CO
PLATE

Rio Cravo Sur

SIN
Guaduas syncline
N

N
Rio Cusiana

BA
ER
NAZCA

OS
ST

AN
EA

LL
0 50 100 km

Sedimentary Igneous
Upper
f
ssi
Quaternary Cenozoic
Magdalena
Ma

Pliocene Cretaceous
Valley Miocene Jurassic
me

Paleogene Triassic
eta

Upper Cretaceous Paleozoic


Lower Cretaceous
Qu

Metamorphic
Jurassic Mesozoic
Triassic Paleozoic
Paleozoic Proterozoic

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1792
Research Paper

The Jurassic to Neogene succession in the Middle Magdalena Valley was governed by fluvial deposition involving alternating phases of channel
­ asin (Fig. 2) recorded relatively continuous accumulation of 5–10 km of clas-
B and overbank accumulation, as well as alluvial-fan and possible fluvial mega-
tic sediment in marine and nonmarine settings. Age control is provided by fan sedimentation (La Paz, Cantagallo, San Juan de Rio Seco, Esmeraldas,
­marine fossil assemblages for the lower succession (Etayo-Serna et al., 1983; ­Mugrosa, Colorado, Real, and Mesa Formations).
­Etayo-Serna, 1985) and a combination of pollen, freshwater mollusks, and vol- Mesozoic–Cenozoic stratigraphic units of the Middle Magdalena Valley
canic ash geochronology for the upper succession (Gómez et al., 2003, 2005b; ­Basin show complex spatial variations, leading to ambiguous stratigraphic
Jaramillo et al., 2011). Initial basin accumulation during Middle–Late Jurassic correlations (Morales, 1958; Sarmiento Rojas, 2001; Sarmiento-Rojas et al.,
(Girón Formation) to Early Cretaceous time (Los Santos, Rosa Blanca, and 2006). There are no field localities with complete exposures of the Jurassic
Paja Formations) involved proximal deposition of clastic facies in alluvial-fan, through Neogene succession, and many units are defined only in the sub­
fluvial, and shallow-marine settings. Subsequent shallow- to deep-marine surface. The lack of uninterrupted exposures and the subsurface presence of
deposition of mixed sand, mud, and limited carbonate persisted throughout intra­basinal structural highs (e.g., the subsurface Infantas, La Cira, and Cáchira
late Early Cretaceous to early Late Cretaceous time (Tablazo, Simití, and Simi- highs) and associated unconformities (Morales, 1958; Gómez et al., 2003,
jaca Formations). The latest Cretaceous–Paleocene marked a shift to sandier 2005b; Caballero et al., 2013a) have precluded delineation of systematic cross-
sedi­menta­tion, with significant progradation of fluviodeltaic systems (Umir, strike (east-west) trends in lithofacies and associated depositional systems.
­Lisama, Seca, and Hoyón Formations). The Eocene through Pliocene history Along-strike (north-south) variations in the aforementioned stratigraphic units

Guane Cocuyo Cagui


well well well
Quaternary
Pliocene SSW Mesa NNE
l Real Real
Mio m Honda Colorado
e Colorado
Santa Teresa
CENOZOIC

l Barzaloza Mugrosa Mugrosa


Olig e
l San Juan de Rioseco La Paz Esmeraldas Esmeraldas
Cantagallo
Eoc m upper Hoyon La Paz
e MMVB unconformity Infantas - La Cira high
Pal el lower/middle Hoyon Seca Lisama
Lisama
Seca Umir
Cimarrona Labor & Tierna Umir
Late Buscavides La Luna
Olini Salto
CRETACEOUS

Villeta
Simijaca Simití
Hilo Tablazo
Capotes Paja
Socotá
Early
Utica Paja Rosablanca
Naveta Murca Los Santos-Tambor

Guaduas Rio Ermitaño Opón La Salina Nuevo Mundo


syncline syncline syncline footwall syncline
– west – east

0 50 100 km

Figure 2. Generalized stratigraphic profile along the SSW-NNE axis of the Middle Magdalena Valley Basin at 4°N–7.7°N (after Caballero et al., 2013b; Sarmiento Rojas,
2001) showing approximate spatial and chronological information for Mesozoic–Cenozoic stratigraphic units from the nine study localities. MMVB—Middle Magda-
lena Valley Basin.

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1793
Research Paper

Supplemental Table 1. Sample information for the Middle Magdalena Valley Basin (MMVB) and adjacent tectonic provinces of the northernmost Andes of Colombia
Site Sample Age Stratigraphic Unit n = Latitude (N) Longitude (W) Reference

(Fig. 2) ­reveal more straightforward lithostratigraphic trends along the basin et al., 2012, 2013; Caballero et al., 2013a, 2013b). To enable reasonable com-
Cocuyo Well Cocuyo1-7 3. Late Cretaceous early Paleocene lower Umir Fm. 16 7.39530 73.85025 This study
Cocuyo Well Cocuyo1-5 5. early – late Eocene upper Cantagallo Fm. 101 7.39530 73.85025 This study
Cocuyo Well Cocuyo1-4 6. late Eocene – early Oligocene Esmeraldas Fm. 52 7.39530 73.85025 This study
Cocuyo Well Cocuyo1-3 7. Oligocene lower Mugrosa Fm. 48 7.39530 73.85025 This study
Cocuyo Well Cocuyo1-2 8. early – middle Miocene lower Colorado Fm. 102 7.39530 73.85025 This study
Guane Well
Guane Well
Guane Well
Guane1-14
Guane1-13
Guane1-12
1. Jurassic
2. Early Cretaceous
2. Early Cretaceous
Giron Fm.
Los Santos Fm.
Rosa Blanca Fm.
112
55
52
7.23360
7.23360
7.23360
73.79962
73.79962
73.79962
This study
This study
This study
axis, likely due to more-continuous north-trending structures within the re- parisons across these recent data sets from different researchers, we em-
Guane Well Guane1-9 6. late Eocene – early Oligocene Esmeraldas Fm. 112 7.23360 73.79962 This study

gional structural framework. ployed the aforementioned filters for age errors and age discordance across
Guane Well Guane1-8 7. Oligocene Mugrosa Fm. 107 7.23360 73.79962 This study
Guane Well Guane1-2 9. late Miocene Real Fm. 82 7.23360 73.79962 This study
Cagui Well CAG1-14 1. Jurassic Giron Fm. 99 7.68098 73.57646 This study
Cagui Well CAG1-12 2. Early Cretaceous Paja Fm. 96 7.68098 73.57646 This study

For the purpose of this study, we systematically consider results from an all s­ ample results.
Cagui Well CAG1-11 6. late Eocene – early Oligocene Esmeraldas Fm. 92 7.68098 73.57646 This study
Cagui Well CAG1-10 7. Oligocene lower Mugrosa Fm. 124 7.68098 73.57646 This study
Cagui Well CAG1-9 7. Oligocene upper Mugrosa Fm. 77 7.68098 73.57646 This study
Cagui Well CAG1-8 8. early – middle Miocene lower Colorado Fm. 94 7.68098 73.57646 This study
Cagui Well CAG1-7 8. early – middle Miocene middle Colorado Fm. 115 7.68098 73.57646 This study
Cagui Well
Cagui Well
Cagui Well
Nuevo Mundo syncline - East
CAG1-6
CAG1-5
CAG1-2
RS0114091
8. early – middle Miocene
9. late Miocene
9. late Miocene
3. Late Cretaceous early Paleocene
upper Colorado Fm.
lower Real Fm.
upper Real Fm.
Lisama Fm.
97
84
81
39
7.68098
7.68098
7.68098
7.21700
73.57646
73.57646
73.57646
73.32700
This study
This study
This study
Nie et al. (2010)
~300 × 50 km swath of the Middle Magdalena Valley Basin (Fig. 1), including
three wells from the western to far northeastern basin (Cocuyo, Guane, and
Nuevo Mundo syncline - East U821 4. mid-Paleocene early Eocene Lisama Fm. 56 7.22500 73.32901 Nie et al. (2010)
Nuevo Mundo syncline - East U08022 4. mid-Paleocene early Eocene La Paz Fm. 27 7.22900 73.33302 Nie et al. (2010)
Nuevo Mundo syncline - East CU612P 5. early – late Eocene La Paz Fm. 46 7.23900 73.34202 Nie et al. (2010)

MODERN RIVERS AND SEDIMENT SOURCE REGIONS


Nuevo Mundo syncline - East U08024 5. early – late Eocene Esmeraldas Fm. 67 7.23497 73.35015 Nie et al. 2012)
Nuevo Mundo syncline - East VC062 6. late Eocene – early Oligocene Esmeraldas Fm. 98 7.23894 73.35805 Nie et al. (2012)
Nuevo Mundo syncline - East
Nuevo Mundo syncline - East
Nuevo Mundo syncline - East
VC063
U08025
VC066
6. late Eocene – early Oligocene
7. Oligocene
7. Oligocene
Esmeraldas Fm.
Mugrosa Fm.
Mugrosa Fm.
95
72
91
7.23672
7.25290
7.24389
73.35585
73.37502
73.35800
Nie et al. (2012)
Nie et al. (2010)
Nie et al. (2012)
Cagui wells), three surface sites from the eastern basin (Nuevo Mundo syn-
Nuevo Mundo syncline - East VC067 7. Oligocene Mugrosa Fm. 88 7.24664 73.35909 Nie et al. (2012)

cline west limb, Nuevo Mundo syncline east limb, and La Salina footwall), and
Nuevo Mundo syncline - East M09 8. early – middle Miocene Colorado Fm. 43 7.25500 73.38401 Nie et al. (2010)
Nuevo Mundo syncline - East U08027 8. early – middle Miocene Colorado Fm. 67 7.26200 73.38902 Nie et al. (2010)
Nuevo Mundo syncline - East U08028 9. late Miocene Real Gr. 49 7.26200 73.41002 Nie et al. (2010)
Nuevo Mundo syncline - West LM1505097 3. Late Cretaceous early Paleocene Lisama Fm. 53 7.13816 73.53906 Caballero et al. (2013a)

three surface sites from the central-southern basin (Opón syncline, Rio Ermi- Detrital zircon U-Pb age distributions for modern river sands in the north-
Nuevo Mundo syncline - West NM1-2 4. mid-Paleocene early Eocene La Paz Fm. 0 7.13365 73.52729 Reyes-Harker et al. (2015)
Nuevo Mundo syncline - West NM2A 4. mid-Paleocene early Eocene La Paz Fm. 88 7.13365 73.52729 Reyes-Harker et al. (2015)
Nuevo Mundo syncline - West NM2B 4. mid-Paleocene early Eocene La Paz Fm. 81 7.13365 73.52729 Reyes-Harker et al. (2015)
Nuevo Mundo syncline - West NM3A 4. mid-Paleocene early Eocene Toro shale Fm. 78 7.13330 73.52692 Reyes-Harker et al. (2015)
Nuevo Mundo syncline - West NM3B 4. mid-Paleocene early Eocene Toro shale Fm. 88 7.13330 73.52692 Reyes-Harker et al. (2015)
Nuevo Mundo syncline - West
Nuevo Mundo syncline - West
Nuevo Mundo syncline - West
Nuevo Mundo syncline - West
NM1
WS0110097
WS0107094
NM4
5. early – late Eocene
5. early – late Eocene
5. early – late Eocene
5. early – late Eocene
La Paz Fm.
La Paz / Esmeraldas Fm.
Esmeraldas Fm.
Esmeraldas Fm.
86
52
83
93
7.13364
7.25296
7.17364
7.13309
73.52727
73.52739
73.55413
73.52624
Reyes-Harker et al. (2015)
This study
This study
Caballero et al. (2013a)
taño syncline, and Guaduas syncline). ernmost Andes (Nie et al., 2012; Bande et al., 2012; Saylor et al., 2013) demon-
strate the distinctive age signatures for several competing sediment sources:
Nuevo Mundo syncline - West LM1505093 6. late Eocene – early Oligocene Esmeraldas Fm. 106 7.17364 73.55413 Caballero et al. (2013a)
Nuevo Mundo syncline - West NM6A 6. late Eocene – early Oligocene Esmeraldas Fm. 75 7.17378 73.55220 Reyes-Harker et al. (2015)
Nuevo Mundo syncline - West NM6B 6. late Eocene – early Oligocene Esmeraldas Fm. 104 7.17378 73.55220 Reyes-Harker et al. (2015)
Nuevo Mundo syncline - West Santos111A 6. late Eocene – early Oligocene Esmeraldas Fm. 88 7.35276 73.45267 Caballero et al. (2013a)
Nuevo Mundo syncline - West Santos111B 6. late Eocene – early Oligocene Esmeraldas Fm. 74 7.35276 73.45265 Reyes-Harker et al. (2015)
Nuevo Mundo syncline - West
Nuevo Mundo syncline - West
Nuevo Mundo syncline - West
WS0109095
Lisama146A
LM1505094
6. late Eocene – early Oligocene
7. Oligocene
7. Oligocene
Esmeraldas Fm.
Mugrosa Fm.
Mugrosa Fm.
51
92
99
7.25555
7.12621
7.17369
73.52587
73.55093
73.55208
This study
Caballero et al. (2013a)
Caballero et al. (2013a)
principally two contiguous ranges, the Central Cordillera and Eastern Cor­di­
U-Pb GEOCHRONOLOGICAL METHODS
Nuevo Mundo syncline - West LM1505095 7. Oligocene Mugrosa Fm. 94 7.17481 73.54842 Nie et al. (2010)

llera, but also two localized block uplifts, the Santander Massif and San Lucas
Nuevo Mundo syncline - West NM7 7. Oligocene Mugrosa Fm. 79 7.17462 73.54541 Bayona et al. (2012)
Nuevo Mundo syncline - West NM8A 7. Oligocene Mugrosa Fm. 92 7.17443 73.54496 Reyes-Harker et al. (2015)
Nuevo Mundo syncline - West NM8B 7. Oligocene Mugrosa Fm. 89 7.17443 73.54496 Reyes-Harker et al. (2015)
Nuevo Mundo syncline - West LM1505096 8. early – middle Miocene Colorado Fm. 101 7.16872 73.53532 Caballero et al. (2013a)

range (Fig. 1).


La Salina footwall LC08031 7. Oligocene Mugrosa Fm. 68 6.85400 73.77102 Caballero et al. (2013a)
La Salina footwall BEL1-2 8. early – middle Miocene Colorado Fm. 85 7.10662 73.57171 This study
La Salina footwall LC08033 8. early – middle Miocene Colorado Fm. 41 6.88500 73.75202 Caballero et al. (2013a)
La Salina footwall LC08035 9. late Miocene Real Gr. 89 7.09700 73.62702 Caballero et al. (2013a)
Opón Syncline 1072-014 3. Late Cretaceous early Paleocene Umir Fm. 102 6.24721 73.79026 Caballero et al. (2013b)
Opón Syncline
Opón Syncline
Opón Syncline
Opón Syncline
1072-028
1072-029
1072-030
1072-031
3. Late Cretaceous early Paleocene
4. mid-Paleocene early Eocene
4. mid-Paleocene early Eocene
4. mid-Paleocene early Eocene
Umir Fm.
Lisama Fm.
Lisama Fm.
Lisama Fm.
92
72
90
85
6.39791
6.37974
6.37288
6.35323
73.74854
73.78066
73.78633
73.85636
Caballero et al. (2013b)
Caballero et al. (2103b)
Caballero et al. (2103b)
Caballero et al. (2103b)
Detrital zircon U-Pb geochronological results for samples of Mesozoic- (1) Along the western flank of the Middle Magdalena Valley Basin, the
Ceno­zoic sandstones and modern river sands (Supplemental Tables 11 and 22) Central Cordillera forms a contiguous north-trending barrier constructed of
Opón Syncline 1072-017 7. Oligocene Mugrosa Fm. 76 6.36697 73.86109 Caballero et al. (2013b)
Opón Syncline 1072-018 7. Oligocene Mugrosa Fm. 95 6.38162 73.88144 Caballero et al. (2013b)
Opón Syncline 1072-025 7. Oligocene Mugrosa Fm. 89 6.41103 73.87002 Caballero et al. (2013b)
Opón Syncline RO3101092 8. early – middle Miocene Colorado Fm. 94 6.38095 73.89758 Caballero et al. (2103b)
Opón Syncline 1072-013 8. early – middle Miocene Colorado Fm. 95 6.40813 73.85950 Caballero et al. (2013b)
Opón Syncline
Opón Syncline
Opón Syncline
1072-020
1072-026
RO290109
8. early – middle Miocene
8. early – middle Miocene
9. late Miocene
Colorado Fm.
Colorado Fm.
Real Gr.
99
95
37
6.38096
6.44892
6.38984
73.89528
73.83852
73.93400
Caballero et al. (2013b)
Caballero et al. (2013b)
Caballero et al. (2103b)
were acquired through LA-ICP-MS, principally at the University of Arizona upper Paleozoic to lower Cenozoic igneous and metamorphic rocks (Fig. 1).
Opón Syncline 1072-016 9. late Miocene Real Gr. 86 6.45752 73.80868 Caballero et al. (2013b)

LaserChron Center following procedures established for the facility (Gehrels, Modern river sands collected from the mouths of the two largest drainage
1
Supplemental Table 1. Sample information for the 2000, 2014; Gehrels et al., 2008). New U-Pb results, consisting of a single U-Pb networks, the Rio Nare and Rio Nechi (Fig. 1), are both governed by zircon
Middle Magdalena Valley Basin and adjacent tectonic
date for each analyzed zircon grain, are reported for 21 subsurface samples U-Pb age peaks of 95–70 Ma (Figs. 3A and 3B), consistent with consider-
provinces of the northernmost Andes of Colombia.
Please visit http://​dx​.doi​.org​/10​.1130​/GES01251​.S1 acquired from three wells. Random selections of >100 inclusion-free zircon able exposure of the Upper Cretaceous Antioquia (Antioqueño) batholith
or the full-text article on www​.gsapubs​.org to view grains were analyzed for each sample, along with known standards to cor- (88–83 Ma) over roughly half the cumulative drainage area in the Central
Supplemental Table 1.
rect for inter- and intra-element fractionation (Gehrels et al., 2008), resulting Cordillera. Additional sources represented in the river sand U-Pb age spectra
Supplemental Table 2. U-Pb geochronological results for the Middle Magdalena Valley Basin (MMVB) and adjacent tectonic provinces of the northernmost Andes of Colombia
Analysis ID U ppm U/Th
207Pb / 1 error 206Pb / 1 error 207Pb/235U 1 error 206Pb/238U 1 error 207Pb/206Pb 1 error
Best Age
1 error
in a 1–2% (2s) age uncertainty for each analysis. Common Pb corrections include Paleogene (65–50 Ma) and Permian–Triassic (300–200 Ma) granites,
235U (%) 238U (%) Age (Ma) Age (Ma) Age (Ma) (Ma)

COCUYO WELL

Cocuyo1-7: lower Umir Fm., Maastrichtian (7.39530°N, 73.85025°W)


were employed using measured 204Pb and an assumed initial Pb composi- and Paleozoic metasedimentary and meta-igneous rocks of the Cajamarca
Lower_Umir-10 484 1.9 0.1062 8.6 0.0160 0.6 102.5 8.4 102.6 0.6 99.9 204.5 102.6 0.6
Lower_Umir-9
Lower_Umir-14
Lower_Umir-5
Lower_Umir-2
84
293
36
41
2.5
1.9
1.1
1.0
0.0841
0.1710
0.2867
1.6805
90.7
17.7
116.3
846.4
0.0232
0.0242
0.0265
0.0271
5.0
2.2
2.6
3.1
82.0
160.3
256.0
1001.2
71.6
26.2
269.2

148.0
154.3
168.4
172.3
7.4
3.4
4.3
5.3
-1587.5
250.4
1160.6
4084.8
1677.0
406.0
303.4
38.1
148.0
154.3
168.4
172.3
7.4
3.4
4.3
5.3
tion (Stacey and Kramers, 1975). Errors in determining 206Pb/ 238U, 206Pb/ 207Pb, complex, defined chiefly by 300–250 Ma metamorphic ages with minor
Lower_Umir-15 20 9.7 0.4320 35.1 0.0286 5.8 364.6 108.0 181.9 10.4 1790.7 651.8 181.9 10.4
Lower_Umir-16
Lower_Umir-4
Lower_Umir-7
Lower_Umir-17
44
134
1432
154
2.3
1.6
32.1
4.9
0.2574
0.3089
2.3217
2.1996
40.3
10.2
1.1
3.7
0.0432
0.0438
0.2110
0.1994
4.2
1.6
1.1
3.5
232.6
273.4
1219.0
1180.9
83.9
24.5
8.1
26.1
272.3
276.2
1234.1
1171.9
11.2
4.3
12.4
37.6
-152.4
248.7
1192.2
1197.6
1030.5
232.7
5.9
25.0
272.3
276.2
1192.2
1197.6
11.2
4.3
5.9
25.0
and 206Pb/ 204Pb ratios yielded total measurement errors of ~1–2% (2s) for Protero­zoic to early Paleozoic (roughly 1000–400 Ma) inheritance (McCourt
Lower_Umir-12 256 5.0 2.3377 2.8 0.2073 2.6 1223.8 20.0 1214.4 29.0 1240.5 19.6 1240.5 19.6
Lower_Umir-3
Lower_Umir-11
Lower_Umir-1
Lower_Umir-13
49
282
213
225
3.1
2.7
4.1
2.3
2.6603
2.7268
2.8739
2.7136
4.9
2.0
2.0
2.1
0.2287
0.2294
0.2417
0.2279
1.5
1.9
1.7
1.9
1317.5
1335.8
1375.1
1332.2
36.5
15.2
15.1
15.3
1327.9
1331.3
1395.8
1323.4
18.6
22.9
21.0
22.9
1300.6
1343.0
1343.1
1346.2
91.1
14.0
21.5
14.5
1300.6
1343.0
1343.1
1346.2
91.1
14.0
21.5
14.5
each preferred age, which is reported as either the 206Pb/ 238U age (grains et al., 1984; Aspden et al., 1987; Aleman and Ramos, 2000; Cordani et al.,
Lower_Umir-8 153 1.8 2.9277 2.1 0.2447 1.7 1389.1 16.2 1411.1 21.2 1355.5 25.6 1355.5 25.6

Cocuyo1-5: upper Cantagallo Fm., middle-late Eocene (7.39530°N, 73.85025°W)


Lower_Esmeraldas-114
Lower_Esmeraldas-5
260
405
1.4
2.6
0.0743
0.0518
18.6
32.4
0.0085
0.0108
7.8
3.2
72.8
51.2
13.0
16.2
54.8
69.2
4.2
2.2
713.2
-726.0
361.2
919.0
54.8
69.2
4.2
2.2
younger than 900 Ma) or 206Pb/ 207Pb age (grains older than 900 Ma). Age 2005; Ordóñez-Carmona et al., 2006; Vinasco et al., 2006; Cardona et al., 2010;
discordance filters were employed following previous studies in the region, Villagómez et al., 2011).
Lower_Esmeraldas-84 100 1.5 0.1524 45.7 0.0237 4.4 144.0 61.4 151.3 6.6 25.9 1144.6 151.3 6.6
Lower_Esmeraldas-3 237 1.6 0.1181 13.5 0.0239 1.4 113.3 14.5 152.0 2.1 -639.5 369.3 152.0 2.1
Lower_Esmeraldas-78 187 2.7 0.1663 36.5 0.0241 2.1 156.2 52.9 153.7 3.1 195.2 873.4 153.7 3.1
Lower_Esmeraldas-105 149 2.1 0.1803 33.2 0.0242 2.8 168.3 51.5 154.1 4.2 373.1 763.1 154.1 4.2
Lower_Esmeraldas-7 185 1.8 0.1283 37.4 0.0244 3.7 122.6 43.2 155.1 5.7 -471.1 1014.3 155.1 5.7

with all single-grain analyses exhibiting >10% uncertainty and all 206Pb/ 238U (2) At the northern termination of the Central Cordillera, the San Lucas
Lower_Esmeraldas-58 439 1.7 0.1727 8.8 0.0248 1.9 161.7 13.2 157.6 3.0 222.4 199.1 157.6 3.0
Lower_Esmeraldas-107 245 2.2 0.1715 15.5 0.0251 1.3 160.7 23.0 159.9 2.1 173.5 362.1 159.9 2.1
Lower_Esmeraldas-1 178 1.0 0.1807 25.9 0.0253 1.2 168.7 40.3 161.1 2.0 277.0 601.5 161.1 2.0
Lower_Esmeraldas-70 407 2.4 0.1672 7.0 0.0260 1.4 157.0 10.1 165.3 2.3 33.0 163.4 165.3 2.3
Lower_Esmeraldas-4 62 1.4 0.2131 53.2 0.0260 7.0 196.2 95.2 165.5 11.5 584.0 1231.4 165.5 11.5

ages displaying >20% discordance or >5% reverse discordance omitted from range (Fig. 1) forms an isolated block uplift composed of Jurassic volcaniclas-
Lower_Esmeraldas-87 297 2.4 0.1871 26.7 0.0266 2.9 174.1 42.8 169.5 4.8 237.5 622.7 169.5 4.8
Lower_Esmeraldas-116 124 1.3 0.2055 28.0 0.0274 3.0 189.8 48.5 174.4 5.2 384.9 636.3 174.4 5.2
Lower_Esmeraldas-68 123 1.3 0.1808 19.0 0.0282 2.0 168.7 29.6 179.5 3.6 20.5 458.0 179.5 3.6
Lower_Esmeraldas-88 162 1.1 0.2025 19.7 0.0283 2.2 187.3 33.7 180.2 3.8 277.8 452.6 180.2 3.8
Lower_Esmeraldas-74 139 0.9 0.2044 30.1 0.0286 3.3 188.9 51.9 181.9 5.9 276.9 698.7 181.9 5.9

consideration. Given the potential for minor discordance, all provenance and tic rocks capping Triassic–Jurassic granodiorite and Grenvillian-age metamor-
Lower_Esmeraldas-91 91 1.5 0.1560 54.1 0.0288 1.7 147.2 74.3 183.1 3.0 -397.9 1509.6 183.1 3.0
Lower_Esmeraldas-81 157 1.5 0.1980 20.2 0.0288 1.5 183.4 33.9 183.3 2.8 185.4 473.3 183.3 2.8
Lower_Esmeraldas-97 672 0.6 0.2074 3.8 0.0291 1.1 191.3 6.6 185.0 2.0 271.0 83.1 185.0 2.0
Lower_Esmeraldas-80 110 1.7 0.2078 19.2 0.0293 7.3 191.7 33.5 186.0 13.4 262.0 409.8 186.0 13.4
Lower_Esmeraldas-118 74 1.6 0.2523 24.8 0.0294 3.6 228.5 50.8 186.8 6.6 681.8 531.5 186.8 6.6

chronostratigraphic interpretations were developed on the basis of age popu- phic basement developed on a Mesoproterozoic protolith (Gómez Tapias et al.,
Lower_Esmeraldas-86 50 0.7 0.2191 34.2 0.0296 3.9 201.2 62.6 187.8 7.2 360.2 788.6 187.8 7.2
Lower_Esmeraldas-79 284 1.2 0.1955 10.3 0.0300 1.1 181.3 17.1 190.6 2.1 62.1 244.4 190.6 2.1
Lower_Esmeraldas-82 119 2.3 0.2705 17.1 0.0374 9.7 243.1 37.0 236.4 22.6 307.9 322.3 236.4 22.6
Lower_Esmeraldas-119 378 4.0 0.2419 7.8 0.0374 2.0 220.0 15.4 236.7 4.6 43.9 180.2 236.7 4.6
Lower_Esmeraldas-39 622 2.9 0.2555 6.7 0.0381 2.4 231.0 13.9 240.8 5.6 132.6 148.1 240.8 5.6

lations defined by results for three or more zircon grains. U-Pb data are plotted 2007; Clavijo et al., 2008; Cuadros et al., 2014; Blanco-Quintero et al., 2014).
Lower_Esmeraldas-24 95 1.8 0.2827 20.8 0.0383 1.8 252.8 46.5 242.6 4.3 348.6 471.9 242.6 4.3
Lower_Esmeraldas-110 149 1.8 0.2748 9.2 0.0398 2.1 246.5 20.2 251.8 5.3 196.3 209.3 251.8 5.3
Lower_Esmeraldas-9 65 1.5 0.2947 28.0 0.0402 4.3 262.2 64.9 254.3 10.8 333.8 639.0 254.3 10.8
Lower_Esmeraldas-57 102 1.5 0.2722 21.8 0.0411 3.0 244.4 47.5 259.7 7.6 100.8 517.0 259.7 7.6
Lower_Esmeraldas-10 336 1.7 0.2947 6.3 0.0415 1.5 262.2 14.6 261.9 3.8 265.4 141.4 261.9 3.8

as age probability distribution functions and age histograms. Significant age U-Pb results from modern river sands collected at the mouth of the Rio Santo
Lower_Esmeraldas-25 179 1.8 0.2679 13.9 0.0420 1.3 241.0 29.8 265.0 3.3 13.3 334.3 265.0 3.3
Lower_Esmeraldas-32 380 2.0 0.3059 6.8 0.0420 1.6 271.0 16.2 265.3 4.1 320.6 150.5 265.3 4.1
Lower_Esmeraldas-31 206 1.8 0.2967 8.2 0.0422 1.4 263.8 19.0 266.2 3.6 242.7 185.9 266.2 3.6
Lower_Esmeraldas-18 235 1.7 0.2929 7.7 0.0426 2.8 260.8 17.6 268.7 7.3 190.2 166.3 268.7 7.3
Lower_Esmeraldas-19 650 1.7 0.3097 4.9 0.0430 2.2 274.0 11.7 271.2 6.0 297.6 98.3 271.2 6.0

populations are generally reported as age ranges rather than individual age Domingo along the northwestern Middle Magdalena Valley Basin margin ex-
Lower_Esmeraldas-8 270 1.2 0.2975 5.3 0.0430 3.1 264.5 12.4 271.5 8.4 202.4 99.3 271.5 8.4
Lower_Esmeraldas-29 160 1.6 0.3061 15.6 0.0432 2.2 271.2 37.0 272.8 6.0 257.7 355.5 272.8 6.0
Lower_Esmeraldas-37 141 1.7 0.2789 21.1 0.0434 1.4 249.8 46.8 274.0 3.7 27.8 510.0 274.0 3.7
Lower_Esmeraldas-103 76 1.2 0.3741 19.5 0.0434 2.6 322.7 54.0 274.1 7.0 689.6 415.8 274.1 7.0
Lower_Esmeraldas-94 386 1.6 0.3178 5.5 0.0435 0.9 280.2 13.5 274.6 2.5 327.2 123.6 274.6 2.5

peaks (i.e., the probability maxima of distribution functions), in order to em- hibit a dominant Permian–Triassic (300–200 Ma) population within a broad
Lower_Esmeraldas-104 481 1.8 0.3030 4.4 0.0436 2.0 268.8 10.3 275.2 5.4 213.2 90.1 275.2 5.4
Lower_Esmeraldas-33 759 1.3 0.3073 4.8 0.0438 2.1 272.1 11.5 276.3 5.7 236.6 99.9 276.3 5.7
Lower_Esmeraldas-52 158 1.2 0.3074 12.5 0.0440 1.8 272.2 29.9 277.4 4.9 227.4 286.9 277.4 4.9
Lower_Esmeraldas-67 539 1.6 0.3152 3.4 0.0440 2.2 278.2 8.2 277.4 5.9 285.1 58.9 277.4 5.9
Lower_Esmeraldas-120 260 1.6 0.3037 8.2 0.0440 1.3 269.3 19.4 277.4 3.6 198.8 188.3 277.4 3.6

phasize and correlate age trends across multiple samples. population of Carboniferous–Jurassic (350–150 Ma) ages, along with a minor
Lower_Esmeraldas-65 290 1.8 0.3153 5.3 0.0440 0.8 278.3 12.8 277.5 2.1 284.7 119.2 277.5 2.1
Lower_Esmeraldas-64 293 2.2 0.3082 8.7 0.0441 4.4 272.8 20.9 277.9 12.1 229.3 173.6 277.9 12.1
Lower_Esmeraldas-109 250 2.5 0.3433 6.4 0.0455 2.2 299.7 16.6 286.6 6.0 402.6 134.9 286.6 6.0
Lower_Esmeraldas-69 210 1.2 0.3663 5.7 0.0516 2.1 316.9 15.6 324.6 6.8 261.1 122.0 324.6 6.8
Lower_Esmeraldas-48 562 5.0 0.5563 2.6 0.0736 1.8 449.1 9.5 457.9 8.0 404.0 42.5 457.9 8.0

Previously published U-Pb data are incorporated into our regional analy- early Paleozoic (500–400 Ma) signal (Fig. 3C).
Lower_Esmeraldas-40 115 1.5 0.5468 10.5 0.0745 1.7 442.9 37.7 463.5 7.7 337.0 235.3 463.5 7.7
Lower_Esmeraldas-13 299 5.0 0.6291 3.3 0.0801 1.8 495.5 13.1 496.8 8.4 489.8 62.8 496.8 8.4
Lower_Esmeraldas-53 143 1.8 1.4423 6.4 0.1468 4.8 906.7 38.3 883.2 39.8 964.3 85.4 883.2 39.8
Lower_Esmeraldas-30 103 3.0 1.4313 3.6 0.1500 1.8 902.1 21.3 900.8 15.4 905.1 63.0 905.1 63.0
Lower_Esmeraldas-27 235 5.1 1.5224 2.3 0.1569 1.7 939.5 14.0 939.8 14.7 938.7 31.9 938.7 31.9

sis, including results for 83 samples from six exposed stratigraphic sections (3) In the Santander Massif, along the northern margin of the Eastern Cor-
Lower_Esmeraldas-102 165 3.0 1.5458 3.3 0.1576 1.6 948.8 20.4 943.2 14.1 961.9 59.1 961.9 59.1
Lower_Esmeraldas-55 210 5.1 1.5599 2.7 0.1578 2.1 954.4 17.0 944.4 18.2 977.6 36.7 977.6 36.7
Lower_Esmeraldas-23 156 5.4 1.7177 3.8 0.1705 2.7 1015.2 24.5 1015.0 25.3 1015.6 54.9 1015.6 54.9
Lower_Esmeraldas-11 180 2.2 1.7540 4.2 0.1735 3.7 1028.6 27.2 1031.6 35.5 1022.4 39.4 1022.4 39.4
Lower_Esmeraldas-85 246 4.1 1.8246 1.9 0.1767 1.4 1054.3 12.6 1049.0 13.2 1065.4 27.3 1065.4 27.3

and seven samples of modern river sands (Nie et al., 2010, 2012; Ibañez-Mejia dillera, two modern drainage networks encompass Triassic–Jurassic granites
Lower_Esmeraldas-34 477 3.4 1.8303 2.4 0.1751 2.3 1056.4 16.1 1040.2 21.9 1090.0 17.8 1090.0 17.8
Lower_Esmeraldas-2 64 1.2 1.9088 6.2 0.1808 3.4 1084.2 41.6 1071.4 33.1 1109.9 105.2 1109.9 105.2
Lower_Esmeraldas-71 637 3.6 2.1331 2.0 0.1980 2.0 1159.6 14.2 1164.3 20.8 1150.8 12.0 1150.8 12.0
Lower_Esmeraldas-15 142 3.6 1.7528 5.8 0.1621 4.9 1028.2 37.2 968.2 44.2 1158.3 59.4 1158.3 59.4
Lower_Esmeraldas-51 50 2.1 2.2096 3.4 0.2040 2.0 1184.1 23.8 1196.6 21.8 1161.3 54.6 1161.3 54.6

et al., 2011; Bayona et al., 2012, 2013; Caballero et al., 2013a, 2013b). These and subordinate Ordovician–Silurian metasedimentary rocks and Precambrian
Lower_Esmeraldas-106 74 2.1 2.2278 4.6 0.2047 1.8 1189.8 32.1 1200.5 20.1 1170.5 83.0 1170.5 83.0
Lower_Esmeraldas-77 406 3.9 2.2606 1.9 0.2076 1.5 1200.1 13.1 1216.2 17.1 1171.1 20.6 1171.1 20.6

2
Supplemental Table 2. U-Pb geochronological re- data were acquired by LA-ICP-MS in the same or similar laboratory settings basement (Fig. 1). These distributions are faithfully represented by U-Pb age
sults for the Middle Magdalena Valley Basin and ad- (in laboratories at the University of Arizona, University of Texas at Austin, spectra for modern river sands from the Rio Umbala and Rio Manco (Figs.
jacent tectonic provinces of the northernmost Andes
and Washington State University) using comparable methods, as detailed in 3D and 3E). The rivers are governed by 300–150 Ma ages, with clustering of
of Colombia. Please visit http://​dx​.doi​.org​/10​.1130​
/GES01251​.S2 or the full-text article on www​.gsapubs​ the original studies (Nie et al., 2010, 2012; Ibañez-Mejia et al., 2011; B
­ ayona 210–190 Ma ages, consistent with Late Triassic–Early Jurassic emplacement
.org to view Supplemental Table 2.

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1794
Research Paper

100

50
A Rio Nare (n = 110)
Central Cordillera
0
40

20
B Rio Nechi (n = 92)
Central Cordillera
0
40

20
C Rio Santo Domingo (n = 67)
San Lucas range
0
40
D Rio Umbala (n = 83)

Relative Probability
number of analyses
20 Figure 3. (A–G) Detrital zircon U-Pb age
Santander Massif distributions for individual samples of
modern rivers in the northernmost Andes.
0
40 Age spectra are depicted as age probabil-

20
E Rio Manco (n = 80)
ity functions (thick black lines) and age
histograms (open rectangles). (H) Plot of
Santander Massif cumulative results, with horizontal arrows
0 highlighting diagnostic age populations
20 and gray shaded zones denoting the Cen-

10
F Rio Cravo Sur (n = 67))
tral Cordillera and Eastern Cordillera.

Eastern Cordillera
0
20

10
G Rio Cusiana (n = 79)
Eastern Cordillera
0
200 Central
al Cordillera
Cordi
H All Rivers (N = 7 samples)
ples) (n = 578)
p
100 S
SMSL
SMSL – Santander Massif, Eastern Cordillera (cratonic sources)
0 San Lucas range
0 200 400 600 800 1000 1200 1400 1600 1800
Age (Ma)

ages for the Santa Barbara granodiorite, Pescadero granite, and associated 400 Ma; Figs. 3F and 3G), which are clear indicators of recycling from exhumed
intrusions (Goldsmith et al., 1971; Irving, 1975; Forero-Suarez, 1990; Dörr et al., Cretaceous sedimentary rocks (Bande et al., 2012; Saylor et al., 2013).
1995; Mantilla Figueroa et al., 2013). Populations of 500–400 Ma and 1200– Although there is modest overlap in ages among these different tectonic
550 Ma grains record input from less-extensive igneous and metamorphic provinces, several essential distinctions are recognized in the U-Pb age distribu-
rocks of Ordovician–Silurian and Proterozoic age, respectively. tions for modern rivers (Fig. 3H). First, the Eastern Cordillera is unambiguously
(4) To the east, the Middle Magdalena Valley Basin is flanked by the broad dominated by 1800–900 Ma zircons, which are diagnostic of voluminous Cre-
Eastern Cordillera, a north-northeast–trending fold-thrust system composed taceous clastic deposits originally derived from Precambrian cratonic sources
mainly of Cretaceous clastic deposits (Fig. 1) originally derived from source to the east in the Guyana Shield and possible subsurface basement sources
areas in the South American craton (Guyana Shield) to the east (Horton et al., of the Llanos Basin (Figs. 3F, 3G, and 3H). Second, the Central Cor­di­llera is
2010b). U-Pb results from the Rio Cravo Sur and Rio Cusiana show primar- uniquely characterized by Late Cretaceous–early Eocene (100–50 Ma) zircons,
ily Proterozoic grains (1800–900 Ma), with peak ages of 1100–900, 1600–1500, with a limited population of Permian–Triassic (300–200 Ma) zircons (Figs. 3A,
1850–1750, and 1250–1200 Ma, and secondary early Paleozoic ages (500– 3B, and 3H). Third, intrabasinal Middle Magdalena Valley and m ­ arginal Middle

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1795
Research Paper

Magdalena Valley uplifts such as the San Lucas range and Santander Massif Jurassic
(and by inference, additional subsurface structural highs such as the I­nfantas,
La Cira, and Cáchira highs) share the 300–200 Ma signature but are differenti- The oldest stratigraphic units of the Middle Magdalena Valley Basin, which
ated by Jurassic (200–150 Ma), Carboniferous (350–300 Ma), and early Paleo- are not exposed at the surface, consist of coarse-grained deposits of Middle–
zoic (500–400 Ma) populations (Figs. 3C–3E and 3H). Late Jurassic to earliest Cretaceous age. At the base of the succession, U-Pb
analyses of two subsurface samples of the coarse-grained nonmarine Girón
Formation yield similarly unimodal age populations composed of Jurassic
SUBSURFACE MIDDLE MAGDALENA VALLEY BASIN RESULTS grains dated at 200–165 Ma (Figs. 4 and 5). Over 70% of analyzed zircon grains
fall in the 190–170 Ma age range, with strong Early to Middle Jurassic peaks
Detrital zircon U-Pb ages for 21 sandstone samples from three wells define at 185–175 Ma in the southwest (Guane well) and 180–170 Ma in the northeast
several key populations and provenance shifts for subsurface Jurassic through (Cagui well). Results from the three youngest grains define a population at
Neogene clastic deposits of the Middle Magdalena Valley Basin (Supplemental ca. 166 Ma, providing a maximum depositional (stratigraphic) age constraint
Tables 1 and 2; Figs. 4 and 5). Whereas the Guane and Cocuyo wells are closer for the subsurface Girón Formation. The only other age signature involves a
to the western margin, the Cagui well is situated near the far northeastern limited occurrence of Precambrian (principally 1800–900 Ma) grains, including
basin margin (Figs. 1 and 2). an 1100–1000 Ma signal.

20
Cocuyo well
10 Late Miocene – Pliocene (n = 80) Guane well

0
40

20 Early − Middle Miocene (n = 101)

0
40

20 Oligocene (N = 2 samples) (n = 152)

0
40
number of analyses

Relative Probability
Late Eocene − Early Oligocene (N = 2 samples) (n = 163)
20
Figure 4. Comparative plots of detrital zir-
con U-Pb age distributions for 11 individual
0
40 subsurface samples from the northwest-
ern Middle Magdalena Valley Basin (Guane
20 Early − Late Eocene (n = 101) and Cocuyo wells), arranged in strati-
graphic order and color coded according to
0 mapped geologic units (Fig. 1).
4

2 Late Cretaceous − Early Paleocene (n = 16)

0
40

20 Early Cretaceous (N = 2 samples) (n = 104)

0
100

50 Jurassic (n = 112)

0
0 200 400 600 800 1000 1200 1400 1600 1800
Age (Ma)

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1796
Research Paper

40 Cagui well
20 Late Miocene – Pliocene (N = 2 samples) (n = 164)

0
40
Early − Middle Miocene (N = 3 samples) (n = 305)
20

0
40

number of analyses

Relative Probability
20 Oligocene (N = 2 samples) (n = 199)
Figure 5. Comparative plots of detrital zir-
con U-Pb age distributions for 10 individual
0
20 subsurface samples from the northeastern
Middle Magdalena Valley Basin (Cagui
10 Late Eocene − Early Oligocene (n = 92) well), arranged in stratigraphic ­order and
color coded according to mapped geologic
0 units (Fig. 1).
20

10 Early Cretaceous (n = 96)

0
100

50 Jurassic (n = 98)

0
0 200 400 600 800 1000 1200 1400 1600 1800
Age (Ma)

Cretaceous (170–150 Ma) ages (Figs. 4 and 5). The continued broad distribution of Precam-
brian ages (principally 1400–900 Ma), which accounts for roughly half of ana-
For the relatively finer-grained marine deposits of the Lower Cretaceous lyzed grains, includes a substantial Mesoproterozoic component with notable
succession, U-Pb ages for three samples from the Los Santos, Paja, and Rosa 1400–1300 Ma and 1250–1150 Ma peaks.
Blanca Formations (Guane and Cagui wells) show the continuation of the
prominent Jurassic (200–150 Ma) age signature, with an Early Jurassic concen-
tration at 190–175 Ma (Figs. 4 and 5). The establishment of a significant popula- Latest Eocene–Earliest Oligocene
tion of Precambrian (1800–900 Ma) zircons, composing roughly half of the total
analyzed grains, is expressed in subpopulations of late Mesoproterozoic–early Three samples of the uppermost Eocene–lowermost Oligocene Esmeral-
Neoproterozoic (1300–900 Ma), early Mesoproterozoic (1600–1500 Ma), and das Formation, collected across the three wells, point to spatial and temporal
Paleoproterozoic (1800–1700 Ma) age. The U-Pb results also reveal the first variations among several detrital age populations within the Middle Magda-
appearance of several minor signals in the Middle Magdalena Valley Basin, lena Valley Basin. In the southwest (Guane and Cocuyo wells), an up-section
including late Paleozoic (300–250 Ma) ages in the northeast (Cagui well) and reduction of the Jurassic (200–150 Ma) signal is accompanied by increased
early Paleozoic (400–500 Ma) ages in the southwest (Guane well). late Paleozoic (300–250 Ma) and continued Precambrian (1800–900 Ma) pop-
ulations, including Mesoproterozoic subpopulations matching those of subja-
Latest Cretaceous–Eocene cent units (Fig. 4). In the northeast, the Cagui well displays the persistence of
a strong Jurassic (200–150 Ma) signal and a broader distribution of Precam-
U-Pb age spectra for two sandstone samples from the uppermost Cre- brian (1800–900 Ma) ages, including a sizeable Paleoproterozoic–early Meso­
taceous Umir Formation and Eocene Cantagallo Formation indicate an up-­ protero­zoic (1800–1400 Ma) population (Fig. 5). In addition, early Paleozoic
section rise in the late Paleozoic (300–250 Ma) population and, within the major (500–400 Ma) and Eocene (40–30 Ma) peaks occur at the expense of the late
Jurassic (200–150 Ma) signal, a shift toward younger Middle to Late Jurassic Paleozoic (300–250 Ma) population.

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1797
Research Paper

Oligocene syncline east limb (Fig. 6). The central-southern Middle Magdalena Valley data
set consists of results from 43 samples of the Opón syncline, Rio Ermitaño
U-Pb results from four samples of the Oligocene Mugrosa Formation sig- syncline, and Guaduas syncline (Fig. 7).
nify a considerable reduction in the formerly dominant Jurassic (200–150 Ma)
signal across the three wells, accounting for only ~5% of analyzed zircon grains Cretaceous
(Figs. 4 and 5). This drop is accompanied by the sustained presence of a re-
gionally significant late Paleozoic (300–250 Ma) population along with an in- Exposures of Mesozoic strata are extremely limited within the basin. A
creased Precambrian (1800–900 Ma) signal. Both margins of the Middle Mag- single mid-Cretaceous sample from the Simijaca Formation in the southern
dalena Valley Basin exhibit high amounts of Precambrian detritus, constituting Middle Magdalena Valley Basin (Guaduas syncline) yields a U-Pb age signa-
roughly 60–80% of total grains analyzed. For the western basin margin, the ture (Fig. 7) governed by Precambrian ages (1800–1450, 1350–1250, and 1150–
Oligocene Mugrosa Formation represents the greatest proportion of Precam- 1000 Ma). This broad Proterozoic age distribution can be correlated with the
brian grains for any part of the Middle Magdalena Valley succession within the aforementioned detrital zircon age spectra for subsurface Cretaceous strata,
Guane and Cocuyo wells (Fig. 4). and it can be readily discriminated from the unimodal Jurassic (200–150 Ma)
age signature for Middle–Upper Jurassic deposits at the base of the basin fill
succession (Fig. 4).
Miocene–Pliocene The strong Proterozoic signal is compatible with crystalline basement
sources such as the expansive Precambrian exposures of the South Ameri-
For the uppermost basin fill, U-Pb results for seven subsurface samples of can craton (Guyana Shield) to the east. However, given the similarity of U-Pb
the Miocene–Pliocene Colorado and Real Formations define key provenance age distributions, it is likely that localized highs composed of basement and
shifts and a sharp spatial contrast across the Middle Magdalena Valley Basin. Paleozoic cover strata within the Eastern Cordillera and/or Llanos Basin also
In the northeast, five samples from the Cagui well display nearly exclusively provided Proterozoic detritus. Such local sources, however, were largely elim-
(~85–95%) Precambrian ages, with prominent age peaks of 1700–1300 and inated by late Early Cretaceous time, when marine depositional conditions
1100–900 Ma, along with subordinate Triassic–Jurassic (250–150 Ma) and early characterized the Eastern Cordillera and flanking Llanos and Magdalena re-
Paleozoic (500–400 Ma) populations (Fig. 5). In contrast, in the southwest, sam- gions (Cooper et al., 1995; Sarmiento-Rojas et al., 2006; Mora et al., 2009).
ples from the Guane and Cocuyo wells record much fewer (<50%) Precam- The absence of the 200–150 Ma signal characterizing subsurface Creta-
brian ages, and a relative strengthening of the Jurassic (200–150 Ma) signature ceous equivalents farther north (Fig. 4) suggests that the preferential influence
(Fig. 4). The stratigraphically highest sample, from the upper Miocene Real of isolated uplifts such as the San Lucas range and Santander Massif in the
Formation in the Guane well (Fig. 4), shows not only the loss of a late Paleozoic northern Middle Magdalena Valley Basin does not persist for the central-south-
(300–250 Ma) signal, but also the emergence of the most substantial propor- ern Middle Magdalena Valley Basin.
tion (up to 10%) of 100–0 Ma grains among all stratigraphic units along the
western basin margin; these pre-Miocene (30–100 Ma) grains likely reflect con- Latest Cretaceous–Early Paleocene
tributions from Cretaceous–Paleogene igneous sources.
Variations in U-Pb age populations are expressed in 10 samples from the
Upper Cretaceous–lower Paleocene Umir, Seca, Labor-Tierna, lower Lisama,
SURFACE MIDDLE MAGDALENA VALLEY BASIN RESULTS and lower Hoyón Formations. Most distinctive is the first appearance of Late
AND REGIONAL CORRELATIONS Cretaceous–Paleocene (90–60 Ma) zircons, with 90–70 Ma grains delineat-
ing the central-southern Middle Magdalena Valley Basin (Fig. 7) and a small
A survey of detrital zircon U-Pb ages for outcrop samples enables spatial 80–60 Ma population defined in the eastern Middle Magdalena Valley Basin
comparisons and correlations among surface and subsurface deposits. Cumu- (Fig. 6). These signatures are not detected in correlative subsurface strata far-
lative data sets are presented for the eastern, central, and southern segments ther north. However, similar to the age distributions of older Cretaceous units
of the Middle Magdalena Valley Basin (Figs. 1 and 2), including the type sec- (Figs. 4 and 5), Precambrian zircons (primarily 1800–1200 Ma) are present and
tions of several stratigraphic units, following the analytical and data treatment remain dominant in the eastern Middle Magdalena Valley Basin, with subordi-
procedures of the original studies (Nie et al., 2010, 2012; Ibañez-Mejia et al., nate populations of Paleozoic (500–400 and 300–250 Ma) ages.
2011; Bayona et al., 2012, 2013; Caballero et al., 2013a, 2013b; Silva et al., 2013; Collectively, these results are consistent with continued clastic contribu-
Reyes-Harker et al., 2015). The eastern Middle Magdalena Valley data set in- tions from distal cratonic sources to the east, but also a localized introduction
cludes cumulative results for 40 samples from the La Salina locality (footwall of 90–60 Ma zircons along the western margin of the southern Middle Magda-
of the La Salina thrust), Nuevo Mundo syncline west limb, and Nuevo Mundo lena Valley Basin (as particularly well expressed in the Guaduas syncline; Figs.

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1798
Research Paper

40 Nuevo Mundo Syncline (E)


Late Miocene – Pliocene (N = 2 samples) (n = 135) Nuevo Mundo Syncline (W)
20
La Salina footwall
0
100

50 Early − Middle Miocene (N = 5 samples) (n = 335)

0
200

100 Oligocene (N = 10 samples) (n = 850)

number of analyses

Relative Probability
0 Figure 6. Comparative plots of detrital zir-
200 con U-Pb age distributions for 40 individual
samples (arranged into seven composite
100 Late Eocene − Early Oligocene (N = 8 samples) (n = 684)
plots) from the eastern Middle Magdalena
Valley Basin (Nuevo Mundo syncline east
0 limb, Nuevo Mundo syncline west limb,
100 and La Salina footwall), arranged in strati-
50 Early − Late Eocene (N = 6 samples) (n = 423) graphic order and color coded according to
mapped geologic units (Fig. 1).
0
40

20 Mid-Paleocene − Early Eocene (N = 7 samples) (n = 402)

0
10

5 Late Cretaceous − Early Paleocene (N = 2 samples) (n = 86)

0
0 200 400 600 800 1000 1200 1400 1600 1800
Age (Ma)

1 and 2). This western source can be linked to earliest uplift of Cretaceous– and progressively replaces an extensive Precambrian component in the east-
Paleo­cene magmatic arc rocks of the Central Cordillera, but it is restricted to ern Middle Magdalena Valley Basin (Fig. 6), is absent in nearly all subsurface
the southern segments of the Middle Magdalena Valley Basin. Paleogene deposits of the northern Middle Magdalena Valley Basin (Figs. 4
and 5). These results may point to progressive northward growth of an inte­
Late Paleocene–Eocene grated western source, with local concentrations of these 90–60 Ma grains
suggesting temporally or spatially focused pulses of sediment influx from Cre-
Widespread changes in U-Pb age spectra are expressed in a sizeable set of taceous–Paleocene igneous rocks of the Central Cordillera. Such northward
34 upper Paleocene–lowermost Oligocene samples from the upper Lisama, propagation of Central Cordillera uplift matches paleogeographic reconstruc-
upper Hoyón, La Paz (including Toro Shale), lower Esmeraldas, and lower San tions (Gómez et al., 2003, 2005b; Bayona et al., 2011; Ayala et al., 2012).
Juan de Rio Seco Formations. The most pronounced shift is represented by
the regional mid-Paleocene emergence of a Late Cretaceous–Paleocene (90– Oligocene–Middle Miocene
60 Ma) age population (Figs. 6 and 7), with a subset of probable syndeposi-
tional zircon grains concentrated at 65–55 Ma (Figs. 6 and 7). Additional pop- Complex shifts in U-Pb age distributions characterize 30 samples of Oligo-
ulations include a broad Precambrian (1800–900 Ma) distribution, a Jurassic cene–middle Miocene clastic fill, including the upper Esmeraldas, Mugrosa,
(200–150 Ma) signal, and minor clusters of early Paleozoic (500–400 Ma) and Colorado, and Santa Teresa Formations. The most distinct change involves an
Permian–Triassic (300–200 Ma) ages. up-section reduction of the Cretaceous–Paleogene (100–50 Ma) signature that
The Late Cretaceous–Paleocene population, which dominates the entire dominated the underlying Eocene section. The progressive relative decline of
Paleogene succession of the southern Middle Magdalena Valley Basin (Fig. 7) this population in the central-southern Middle Magdalena Valley Basin (Fig. 7),

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1799
Research Paper

100 Opón Syncline


Late Miocene – Pliocene (N = 8 samples) (n = 618) Rio Ermitaño Syncline
50 Guaduas Syncline
0
100

50 Early − Middle Miocene (N = 6 samples) (n = 519)

0
400

200 Oligocene (N = 9 samples) (n = 821)

0
200

number of analyses

Relative Probability
100 Late Eocene − Early Oligocene (N = 2 samples) (n = 213) Figure 7. Comparative plots of detrital zir-
con U-Pb age distributions for 43 individ-
ual samples (arranged into eight compos-
0
400 ite plots) from the central-southern Middle
Magdalena Valley Basin (Opón syncline,
200 Early − Late Eocene (N = 3 samples) (n = 296) Rio Ermitaño syncline, and Guaduas syn-
cline), arranged in stratigraphic order and
0 color coded according to mapped geologic
200 units (Fig. 1).

100 Mid-Paleocene − Early Eocene (N = 7 samples) (n = 595)

0
200

100 Late Cretaceous − Early Paleocene (N = 7 samples) (n = 665)

0
10

5 Early Cretaceous (n = 61)

0
0 200 400 600 800 1000 1200 1400 1600 1800
Age (Ma)

and its early–middle Miocene disappearance in the eastern Middle Magdalena fied contributions of Paleozoic and Precambrian ages. Although Precambrian
Valley Basin (Fig. 6), is coupled with the joint introduction of Jurassic (200– age distributions are consistent with direct input from a cratonic source, these
150 Ma) and late Paleozoic (300–250 Ma) signals. These trends are accompa- grains were most likely recycled from uplifted Cretaceous sedimentary rocks
nied by Precambrian age distributions (1800–900 Ma) that define an up-section during mid-Cenozoic growth of the Andean fold-thrust belt encompassing the
increase in the central-southern Middle Magdalena Valley Basin but a com- Eastern Cordillera and associated basement uplifts.
plex increase and then subsequent decrease in the eastern Middle Magdalena
Valley Basin. The complex fluctuation of Precambrian detritus in the eastern Late Miocene–Pliocene
Middle Magdalena Valley Basin can be correlated with comparable shifts in
the subsurface record of the Guane and Cocuyo wells (Fig. 4). However, none U-Pb results from uppermost levels of the Middle Magdalena Valley Basin
of the three wells shows the Late Cretaceous–early Paleogene signal, and the show a distinctively cosmopolitan assemblage distinguished by the abrupt
Cagui well is uniquely distinguished by the Miocene dominance (up to 90%) of ­arrival of young zircons of chiefly Cenozoic age. Results from 10 samples of the
Precambrian zircon populations (Fig. 5). upper Miocene–Pliocene Real and Mesa Formations show a pronounced intro-
The results signify a complex interplay among source areas. Phases of rela­ duction of grains younger than 100 Ma, with significant age probability peaks at
tively diminished influx from Cretaceous–Paleogene igneous sources (such as 88–84, 81–78, 43–40, 14–11, and 9–6 Ma (Figs. 6 and 7). The youngest ages cluster
the 100–50 Ma signal) in the Central Cordillera broadly coincide with ampli- in the 12–6 Ma range and reflect contributions from syndepositional ­magmatic

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1800
Research Paper

sources that approximate the depositional ages for the southern-central Middle roofing or drainage modifications. This late Miocene shift is particularly well
Magdalena Valley Basin (Opón syncline). A broad spread of subordi­nate age expressed in a comparison of composite U-Pb age distributions for Cenozoic
populations includes Precambrian (1800–900 Ma), early Paleozoic (500–400 Ma), basin fill (Fig. 8), which shows the introduction of Late Cretaceous–Cenozoic
Permian–Triassic (300–200 Ma), and Jurassic (200–150 Ma) ages comparable to zircons, consistent with the birth of a proto–Magdalena River.
results for the underlying Oligocene–middle Miocene deposits.
The overall age distributions correlate with subsurface age spectra for
the Guane well (Fig. 4), matching both the appearance of young grains and SEDIMENT SOURCES AND TECTONIC RECORDS
the multiple modes of older pre-Cenozoic populations. Neither data set, how-
ever, correlates with the Precambrian-dominated age signatures for Miocene– Key provenance trends summarized from the U-Pb results (Figs. 3–8)
Pliocene strata of the Cagui well (Fig. 5). The internal consistencies among facili­tate spatial comparisons and correlations of Mesozoic–Cenozoic strati-
some localities yet sharp contrasts among others underscore the continued graphic and tectonic histories for surface and subsurface deposits within the
competition among disparate sediment sources in the Central Cordillera and Middle Magdalena Valley Basin. Broader consideration of statistical simi­lari­
Eastern Cordillera. The wholesale emergence of 100–0 Ma zircons indicates ties among U-Pb age spectra is enabled by calculation of cross-correlation co-
that late Miocene paleodrainage systems of the Middle Magdalena Valley efficients between probability density plots of all sample pairs (Saylor et al.,
Basin gained access to a new sediment source, either through bedrock un- 2012b, 2013; Saylor and Sundell, 2014). Following recent studies (e.g., Chap-

Cagui well (N = 8 samples) (n = 760)

Cocuyo well (N = 3 samples) (n = 201)

Figure 8. Comparative plots of detrital


zircon U-Pb age spectra for 64 individual
samples (arranged into composite plots for
Guane well (N = 3 samples) (n = 295) seven different localities) of late Eocene to
Pliocene deposits from the Middle Magda-

Relative Probability
lena Valley Basin. Differences in age distri-
butions reflect relative contributions from
the Central Cordillera and Eastern Cor­
Nuevo Mundo Syncline (N = 25 samples) (n = 2004) di­
llera, as denoted by horizontal arrows
and gray shaded zones. The northernmost
Middle Magdalena Valley Basin (Cagui
well) is distinguished by large Precambrian
(1800–900 Ma) populations, indicative of
Opón Syncline (N = 15 samples) (n = 1248) greater input from the Eastern Cordillera.
The late Miocene introduction of 100–0 Ma
grains (open arrows) reveals a shift in prov­
enance patterns, possibly related to the ap-
pearance of a through-­going proto–Magda­
Rio Ermitaño Syncline (N = 5 samples) (n = 487) lena River. Note that the color scheme
departs from mapped geologic units (Fig. 1)
Central Cordillera Eastern Cordillera and other U-Pb age plots (Figs. 4–7).

Late Miocene – Pliocene


Guaduas Syncline (N = 5 samples) (n = 436) Early – Middle Miocene
Oligocene
Late Eocene-Early Oligocene

0 200 400 600 800 1000 1200 1400 1600 1800


Age (Ma)

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1801
Research Paper

man et al., 2012, 2015; Vermeesch, 2013), these correlation values are displayed though continental crystalline basement underlies the Eastern Cordillera and
graphically in multidimensional scaling (MDS) plots (Fig. 9) in which greater Middle Magdalena Valley Basin, and small populations of inherited Precam-
similarity is shown as closer proximity. The MDS plots for 104 Jurassic through brian grains are detected in Paleozoic–Mesozoic rocks of the Central Cordillera
Pliocene samples (Figs. 9A–9D) and seven modern river sand samples (Fig. 9E) (Figs. 3A and 9E; Cordani et al., 2005; Vinasco et al., 2006), the sizeable fraction
reveal clear quantitative distinctions among samples derived from the major and broad spread of dated Precambrian grains in the absence of major west-
source regions. ern signatures suggest most basement contributions were derived from the
The Eastern Cordillera, San Lucas and Santander block uplifts, and Central South American craton (Guyana Shield) to the east (Figs. 9A and 10B). How-
Cordillera all define diagnostic U-Pb age signatures (Fig. 9). Whereas the East- ever, the coupled reduction in the formerly dominant Jurassic (200–150 Ma)
ern Cordillera is dominated by Cretaceous sedimentary units originally derived populations and emergence of subordinate Paleozoic (300–250 Ma and 500–
from the Precambrian craton, the Central Cordillera is composed largely of 400 Ma) signals attest to deeper unroofing of local intrabasinal Middle Magda-
Cretaceous–Paleogene igneous rocks of the Andean magmatic arc, and the lena Valley uplifts and probable extrabasinal uplifts, principally in the Eastern
San Lucas and Santander block uplifts are characterized mostly by Jurassic Cordillera, but also conceivably in isolated zones of the Central Cordillera. The
igneous rocks and their sedimentary derivatives. These age distinctions, as ex- absence of syndepositional Cretaceous zircons underscores the lack of signifi-
pressed in modern rivers (Figs. 3 and 9E), serve as critical markers in the sedi­ cant contributions from the Andean magmatic arc to the west, possibly due to
mentary record that allow discrimination of evolving sources during Meso­ transport distance or a regional west-dipping topographic gradient.
zoic–Cenozoic basin evolution, thus providing the paleogeographic framework Whereas local signatures are consistent with individualized basement
for tectonic reconstructions of the region (Fig. 10). uplifts, the prevalent basement ages require an integrated drainage system
across a broad region linking the distal eastern craton to the Middle Magdalena
Jurassic Valley Basin (Figs. 10B and 10C). This region includes the present-day Eastern
Cordillera and Llanos Basin, where several subbasins were active during Early
Coarse-grained Upper Jurassic–lowermost Cretaceous units of the lower- Cretaceous extension (Cooper et al., 1995; Sarmiento-Rojas et al., 2006; Mora
most Middle Magdalena Valley Basin are dominated by bedrock signatures of et al., 2009). The coalescence of local fault-related subbasins into larger basin
localized uplifts. The unimodal 200–150 Ma U-Pb age populations (Figs. 4 and systems enabled regional westward transport of basement detritus. Postexten-
5) are consistent with local exposures of Lower to Middle Jurassic granitoid sional thermal subsidence during mid- to Late Cretaceous time induced further
intrusions and correlative volcanic rocks (Cáceres et al., 2003; Gómez ­Tapias basin expansion and depositional overlap of formerly active normal faults and
et al., 2007; Bayona et al., 2011). Unimodal age signatures also define the local sediment sources (Fig. 10C). Regionally, progressive stratigraphic onlap
modern rivers draining intrabasinal and marginal Middle Magdalena Valley eastward across the Llanos Basin (Fig. 1) and onto the South American cra-
uplifts such as the San Lucas range and Santander Massif (Figs. 3C–3E, 3H, 9A, ton (Guyana Shield) resulted in a Late Cretaceous shift to older Precambrian
9E, and 10A). A similar Jurassic age signature is observed in selected Upper sources, with lessened 1200–900 Ma and enhanced 1900–1300 Ma populations
­Jurassic strata (Horton et al., 2010b), suggesting the possibility of a correlation (Horton et al., 2010b; Saylor et al., 2013).
tool between the Middle Magdalena Valley Basin and Eastern Cordillera (e.g.,
Fabre, 1987; Cáceres et al., 2003). Latest Cretaceous–Eocene
Local sources of sediment are consistent with interpretations of a Juras-
sic–earliest Cretaceous extensional setting containing localized coarse-grained A comprehensive reversal of sedimentary polarity is recorded by the intro-
depocenters (Cooper et al., 1995; Sarmiento-Rojas et al., 2006). In this regional duction of a detrital zircon U-Pb age population indicative of a new sediment
extensional system, segregated fault-bounded subbasins (e.g., Mora et al., source along the western margin of the Middle Magdalena Valley Basin (Figs.
2013; Tesón et al., 2013) were fed sediment by small drainage networks erod- 10D and 10E). The Late Cretaceous–Paleocene (90–60 Ma) population (Figs. 6
ing uplifted ranges dominated by Triassic–Jurassic igneous rocks (Fig. 10A). and 7) is unambiguously linked to erosion of igneous rocks in the Central Cor-
These isolated ranges were likely developed in the structural zones presently dillera (Figs. 8, 9B, and 9E), as comparable ages are not observed elsewhere.
occupied by marginal block uplifts such as the San Lucas range and Santander The dominance of this signal is demonstrated by the unimodal 95–70 Ma sig-
Massif (Fig. 1). natures for modern rivers draining the region (Figs. 3A, 3B, and 3H). Less clear
is the precise timing. Although the central-southern Middle Magdalena Valley
Cretaceous Basin shows a roughly Maastrichtian appearance, data sets from the east-
ern-northeastern Middle Magdalena Valley Basin show a striking mid-Paleo-
Cretaceous deposits record a mix of distal cratonic sources and diminished cene shift (e.g., Nie et al., 2010, 2012; Caballero et al., 2013a, 2013b). The intro-
basement uplifts. U-Pb results are most distinguished by the introduction of a duction of this key signature progressively from south to north suggests that
substantial Precambrian (1800–900 Ma) age population (Figs. 4, 5, and 7). Al- uplift of the Central Cordillera was not a single simultaneous event, but rather

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1802
Research Paper

0.6 0.6
C middle–late Oligocene E modern rivers
early Oligocene
0.4 0.4

0.2 0.2

Y
0 0

–0.2 Central San Lucas / –0.2 Central San Lucas /


Cordillera Santander Cordillera Santander

–0.4 –0.4

Eastern Cordillera / craton Eastern Cordillera / craton


0.6 0.6
B middle Eocene–late Eocene D late Miocene–early Pliocene
middle Paleocene–early Eocene early–middle Miocene
0.4 0.4

0.2 0.2
Y

Y
0 0

–0.2 Central San Lucas / –0.2 Central San Lucas /


Cordillera Santander Cordillera Santander

–0.4 –0.4

Eastern Cordillera / craton Eastern Cordillera / craton


0.6 –0.6
A Late Cretaceous–early Paleocene
Early Cretaceous
0.5 0.4 0.3 0.2 0.1
X
0 –0.1 –0.2 –0.3 –0.4

Jurassic
0.4

0.2
Figure 9. Multidimensional scaling (MDS) plots showing relative quantitative
similarity among all U-Pb age distributions for 111 samples, including (A–D)
Jurassic through Pliocene samples (N = 104) and (E) modern river sand sam-
Y

0 ples (N = 7), from the Middle Magdalena Valley Basin and adjacent regions,
arranged in stratigraphic order and color coded according to mapped geologic
units (Fig. 1) and other U-Pb age plots (Figs. 4–7). Similarity was quantified as
–0.2 Central San Lucas / cross-correlation coefficients of probability density plots for all sample pairs
Cordillera and converted to MDS plots in which individual samples (­color-coded dots)
Santander
plotting closer to each other signify a higher degree of similarity. For com-
–0.4 parison, the major source regions identified from modern river samples (the
Central Cordillera, San Lucas/Santander block uplifts, and Eastern Cordillera
[and South American craton]) are plotted as rectangular fields.
Eastern Cordillera / craton
–0.6
0.5 0.4 0.3 0.2 0.1 0 –0.1 –0.2 –0.3 –0.4
X

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1803
Research Paper

A Jurassic B Early Cretaceous C Late Cretaceous D Paleocene 9°N

SM SM SM SM
SL Cagui
SL Cagui
SL Cagui
SL Cagui

Cocuyo Cocuyo Cocuyo Cocuyo


NUE NUE NUE NUE
Guane Guane Guane Guane

SAL SAL SAL SAL


VB

VB

VB
VB
OPO OPO OPO OPO
MM

MM

MM
MM
CC CC CC CC
ERM ERM ERM ERM
EC EC EC EC 6°N

Figure 10. Jurassic through Miocene non-


palinspastically restored map-view paleo-
GUA GUA GUA GUA
geographic reconstructions of the Middle
Magdalena Valley Basin (MMVB) and ad-
QM QM QM QM
LB LB LB LB jacent regions showing major sedi­ ment
dispersal patterns (arrows), zones of active

E Eocene F Oligocene G Miocene 75°W 73°W deposition (stipple pattern), and zones of
exhumation, as defined by exposed out-
n to
Sedimentary Igneous crop regions (color coded accord­ ing to
a
uS Neogene outcrop Cenozoic

Buc
írit mapped geologic units; Fig. 1). Present-day
p

a
Es
Miocene

ram
Cretaceous fault locations are from Gómez Tapias et al.

ang
Paleogene outcrop (2007).
Jurassic

a
Paleogene

Palestina
Triassic
SM SM SM Upper Cretaceous Paleozoic
SL SL El Bagre
SL

Servitá
Cagui Cagui Cagui
Lower Cretaceous Metamorphic
Cocuyo Cocuyo Cocuyo Jurassic Mesozoic
NUE NUE NUE
Guane Guane Guane Triassic Paleozoic
rra
Otú

i ta Paleozoic Proterozoic
im
b e
SAL SAL C SAL
sa

Sample sites
Ca
VB

VB

Cagui well Fault


OPO OPO OPO Cocuyo well name active
MM

MM

CC CC CC Guane well deposition

Boyacá
rez

NUE —Nuevo Mundo syncline


á
Su

ERM ERM ERM


EC EC EC SAL —La Salina footwall
Bi Cambao

OPO —Rio Opón syncline local


ERM —Rio Ermitaño syncline paleoflow
ma

io
ga
tu i

ac

GUA —Guaduas syncline


pa

Ig n
á
So
a
nd

Tectonic provinces
n
Sa
Ho

CC —Central Cordillera
EC —Eastern Cordillera
o

regional
m
ra

GUA GUA GUA


na

SL —San Lucas range


ca

ica

paleoflow
sia
s

a
Pe

SM —Santander Massif
Cu
Gu

QM QM QM QM—Quetame Massif 0
LB LB LB 50 100 km
a

Ibagué
li

LB —Llanos Basin
sa
Te

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1804
Research Paper

progressed northward along strike during Paleogene shortening and trans- sediment from the Eastern Cordillera. Further spatial variability is delineated
pressional deformation. This spatial variability also precludes usage of this by the Cretaceous–Paleogene signature: Although this 100–50 Ma population
provenance signal as a robust tool for stratigraphic correlation. Subsurface is severely reduced to nonexistent in most locations, its local persistence in
samples of Paleogene units from more northwestern regions of the Middle the Opón syncline (Rio Oponcito) locality (central-southern Middle Magda-
Magdalena Valley Basin show none of the Central Cordillera sources detected lena Valley Basin) signifies spatially focused input from the Central Cordillera.
in the southern, central, and eastern Middle Magdalena Valley Basin. Instead, These complex alternations reflect continued competition among eastern and
an up-section increase in the late Paleozoic (300–250 Ma) population (Figs. 4 western sources (Figs. 8 and 9C–9E) and suggest a compartmentalized prov-
and 5) implies derivation from intrabasinal and marginal Middle Magdalena enance record displaying significant spatial variations, both along-strike and
Valley uplifts comparable to the San Lucas range and Santander Massif adja- across-strike (Fig. 10F). Such provenance variations attest to a Cenozoic record
cent to the northern Middle Magdalena Valley Basin (Figs. 1, 3C–3E, and 3H). in the Middle Magdalena Valley Basin dominated by transverse (east-west)
This conspicuous along-strike variability in provenance history indicates a rather than longitudinal (north-south) deposystems, consistent with a closed
­basin dominated by transverse rather than longitudinal deposystems. topographic basin disconnected from the Caribbean Sea (Moreno et al., 2011;
Regionally, the provenance record is consistent with proposed latest Creta- Caballero et al., 2013b; Moreno et al., 2013; Silva et al., 2013; Reyes-Harker
ceous–Paleocene uplift and exhumation in the Central Cordillera (Figs. 10D and et al., 2015).
10E; Gómez et al., 2003; Villagómez and Spikings, 2013). This tectonic episode
appears to represent a fundamental switch from a regional postrift setting to a
zone of shortening and transpression linked to active subduction and possible Miocene–Pliocene
minor collisions along the western margin.
Final modification then establishment of the modern axial drainage config-
uration in the Magdalena Valley was achieved in late Miocene–Pliocene time
Oligocene–Middle Miocene (Fig. 10G). U-Pb results for upper Miocene–Pliocene deposits show the most
cosmopolitan provenance assemblage for the entire Middle Magdalena Val-
Another distinctive reversal in sediment dispersal is expressed by the ley Basin succession, as defined by multimodal Precambrian, Paleozoic, and
emergence of U-Pb age populations revealing growth of a substantial new Mesozoic populations (Figs. 4, 6, and 7). This diverse and very broad distribu­
sediment source along the eastern margin of the Middle Magdalena Valley tion (Fig. 9D) contrasts with the concentrated unimodal and bimodal distri­
Basin. A marked Oligocene amplification of a broadly distributed Precam- butions expressed in most underlying deposits (Figs. 9A–9C). Although minor
brian population (1800–900 Ma) coincides with increased late Paleozoic (300– up-section increases and decreases characterize the various modes, a substan-
250 Ma) to Jurassic (200–150 Ma) ages and an abrupt decline of the formerly tial shift is recorded by the sharp introduction of a 100–0 Ma signal (Fig. 8).
dominant Cretaceous–Paleogene (100–50 Ma) signature (Figs. 8 and 9C). The This critical signal appears across all but one of the surface and subsurface
ubiquitous increase in Precambrian detrital populations across all sample lo- samples (Figs. 4, 6, and 7), and it is only absent from the Cagui well (Fig. 5),
calities (Figs. 4–7) offers a potential correlation tool for surface and subsurface which shows the continued dominance of Precambrian detritus derived from
units. Although Precambrian signatures are consistent with a distal cratonic a point source of sediment eroded from Cretaceous strata in the Eastern Cor-
source (Figs. 3F–3H), independent thermochronological and sedimentary evi- dillera fold-thrust belt. The Cretaceous–Cenozoic (100–0 Ma) signature requires
dence for early uplift in the Eastern Cordillera (Parra et al., 2009a, 2009b, 2012; important contributions from the Central Cordillera (e.g., McCourt et al., 1984;
Mora et al., 2010; Bande et al., 2012; Saylor et al., 2012a, 2012b; Sánchez et al., Aspden et al., 1987; Gómez Tapias et al., 2007; Villagómez et al., 2011; Bayona
2012) requires that Precambrian grains were recycled from uplifted Cretaceous et al., 2012, and references therein), but most probable candidates include the
strata in the Andean fold-thrust belt (Figs. 8 and 9C). Therefore, these old zir- large Antioquia batholith and satellite intrusions (broadly 95–80 Ma) adjacent
con signatures are regarded as a high-fidelity record of early rock uplift and to the central Middle Magdalena Valley Basin and multiple 65–45 Ma batholiths
exhumation of the Eastern Cordillera, with a dominance of eastern over west- and stocks (including the Manizales, El Bosque, and ­Hatillo intrusions) near the
ern sources (Fig. 10F). transition from the Middle Magdalena Valley Basin to the Upper Magdalena
Following this provenance reversal, complex early–middle Miocene shifts Valley Basin.
are defined by fluctuations in various age populations. Most pronounced are The regional emergence of the 100–0 Ma signature (Fig. 8), and the consis-
irregular spatial variations in Precambrian populations, with decreases in most tency of multimodal pre-Cenozoic zircon populations across studied localities
of the northern Middle Magdalena Valley Basin (Figs. 4 and 6) contrasting with reflect drastic modifications in large-scale paleodrainage patterns. Although
increases in the central-southern and far northeastern Middle Magdalena Val- localized input of syndepositional Neogene volcanic material characterized
ley Basin locations (Figs. 5 and 7). A sharp increase in Precambrian detritus the Opón syncline (central-southern Middle Magdalena Valley Basin), the en-
occurs within the Cagui well (Fig. 5), indicating a punctuated point source of hanced Cretaceous–Cenozoic populations point to a drainage expansion, likely

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1805
Research Paper

through capture of upstream catchments of the Upper Magdalena Valley, in mentioned tectonic signatures. Perhaps the most valuable criteria involve the
which the evolving system gained access to new sediment sources (Fig. 10G). identifications of important reversals of sediment polarity (e.g., Coney and
This episode corresponds with a transition from a series of compartmentalized Evenchick, 1994). Such reversals prove to be sufficient in identifying whole-
transverse deposystems, each dominated by sources in either the Central Cor- sale uplift of contiguous barriers but can be complicated by along-strike vari-
dillera or Eastern Cordillera, into an integrated longitudinal system. This late ations. As an example, along-strike propagation of uplift in the Central Cor­di­
Miocene switch constitutes the onset of a through-going proto–Magdalena llera (­Gómez et al., 2003, 2005b; Villagómez and Spikings, 2013) may explain­
River, flowing axially from south to north, and linking the Upper Magdalena ­modest discrepancies from south to north within the flanking Middle Magda-
Valley and Middle Magdalena Valley Basin to the Caribbean coast (Fig. 10G). lena V
­ alley Basin.

DISCUSSION Insufficient Provenance Signatures

Surface and Subsurface Provenance Signatures Despite the utility of detrital provenance archives, several key components
of the geologic and tectonic history appear to be less well expressed, even
Tectonic events of the northernmost Andes (Fig. 1) are robustly expressed missing, from the U-Pb data set. First, a well-documented phase of late Paleo-
in both the surface and subsurface provenance record of the Middle Magda- cene–middle Eocene basement uplift and erosional beveling in the central to
lena Valley Basin (Figs. 4–9). At the broad scale, the detrital zircon U-Pb geo- northern Middle Magdalena Valley Basin (Fig. 10; Gómez et al., 2003, 2005b;
chronological data set faithfully records the three principal Mesozoic–Cenozoic Moreno et al., 2011) appears to be mostly aliased in the detrital record, as there
episodes of the northernmost Andes. First, for the Jurassic–earliest Creta- is no corresponding major shift in U-Pb age distributions (Figs. 4–7 and 9B).
ceous phase of regional extension involving the growth of synrift subbasins Several subsurface basement highs and the associated Middle Magdalena Val-
(e.g., Colletta et al., 1990; Sarmiento Rojas, 2001; Sarmiento-Rojas et al., 2006; ley Basin angular unconformity (Fig. 2) are well imaged by seismic-reflection
­Kammer and Sánchez, 2006; Mora et al., 2009), detrital zircon data show uni- profiles, with supporting borehole data. Although localized, these are large
modal age signatures emblematic of uplifted bedrock sources along localized features (Infantas, La Cira, and Cáchira highs) with over several kilometers
normal faults (Fig. 10A). Second, a Cretaceous phase of regional postexten- of vertical structural relief (Morales, 1958; Moreno et al., 2011; Parra et al.,
sional subsidence (Fabre, 1987; Toussaint and Restrepo, 1994; Cooper et al., 2012; G­ ómez et al., 2003, 2005b; Caballero et al., 2013a, 2013b; Mora et al.,
1995; Roeder and Chamberlain, 1995; Toro et al., 2004) is reflected in the emer- 2013; Tesón et al., 2013). Therefore, it is difficult to envision a case where the
gence of a distal eastern cratonic source replacing localized footwall uplifts provenance signal would be volumetrically eclipsed by other source regions.
(Figs. 10B and 10C). Third, a protracted Cenozoic period of Andean shortening Rather, we suggest that comparable U-Pb age signatures of 300–250 Ma and
(Dengo and Covey, 1993; Taboada et al., 2000; Gómez et al., 2005a; Parra et al., 200–150 Ma for both the uplifted basement and recycled sedimentary units
2009b; Mora et al., 2010) is expressed as a complex mix of proximal sources preclude clear discrimination of the Middle Magdalena Valley Basin base-
dictated by a competition between eastern and western sources during the ment sources. Never­ the­less, low-temperature thermochronological results
growth of basin-bounding ranges (Figs. 10D–10G). from other deposits of this age reveal a sharp influx of material that can be
At finer scales, U-Pb age distributions also provide accurate signatures attributed to Middle Magdalena Valley Basin basement uplift (Saylor et al.,
of two critical transitions during Andean mountain building. First, earliest 2012b), highlighting the utility of complementary techniques.
shortening-related uplift in the flanking Central Cordillera (Gómez et al., 2003, Second, despite the proximity to the western magmatic arc, there is no
2005b) can be tied to a latest Cretaceous–Paleocene reversal in paleodrainage, robust Middle Magdalena Valley Basin record of a proposed Paleogene mag-
as defined by a shift in U-Pb age spectra from an eastern cratonic source to a matic pulse, and U-Pb ages prove to be of limited utility in constraining Paleo-
western orogenic source (Figs. 9A, 9B, 9E, 10C, and 10D). Second, initial uplift gene depositional ages. Bayona et al. (2012) suggested that the clear pulse of
of the Eastern Cordillera (Parra et al., 2009b; Mora et al., 2010; Bande et al., 65–45 Ma magmatism in the Central Cordillera, near the transition from the
2012; Saylor et al., 2012a; Sánchez et al., 2012) is identified by U-Pb age signa- Middle Magdalena Valley Basin to the Upper Magdalena Valley Basin (Mc-
tures (recycled populations of craton-derived Precambrian detritus) recording Court et al., 1984; Aspden et al., 1987; Gómez Tapias et al., 2007; Villagómez
the latest Eocene–Oligocene unroofing of Mesozoic strata in the Andean fold- et al., 2011), may also extend to undiscovered igneous units in the Eastern
thrust belt (Figs. 9B, 9C, 9E, 10E, and 10F). Cordillera. Nevertheless, the U-Pb results illustrate very limited proportions
Of these provenance shifts, nearly all of them provide sufficient fidelity of possible syndepositional zircons within Paleocene–middle Eocene Middle
for correlation of surface and subsurface deposits. Rather than constraining Magdalena Valley Basin strata. Of the 47 samples from the uppermost Creta-
precise matches among U-Pb age spectra of individual samples, such correla- ceous–lowermost Oligocene section, only three samples (all restricted to the
tions are reliant on the up-section appearance or disappearance of the afore- Guaduas syncline in the southernmost basin) contain statistically significant

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1806
Research Paper

populations of grains (n > 2 ages) approaching stratigraphic ages, accounting tering on the late Miocene (Duque-Caro, 1978; Guerrero, 1993; Hoorn et al.,
for a cumulative total of ~1% of all analyzed grains. Given the short transport 1995), Oligocene (Villamil, 1999), or middle Eocene (Escalona and Mann, 2011).
distances (<100–200 km) between the Central Cordillera and the Middle Mag- Most studies propose that growth of the Eastern Cordillera as a topographic
dalena Valley Basin, the limited amount of Cenozoic ages cannot be attributed barrier would have deflected an axial Magdalena River northward into the
to preferential weathering and selective removal of these grains (e.g., Bayona Carib­bean Sea, with the isolated Santa Marta Massif near the northern coast
et al., 2012). Therefore, erosion of the observed Paleogene magmatic centers also playing a role in the drainage shift away from a former eastern outlet
in the Central Cordillera must have fed sediment to only a limited portion of within the Maracaibo Basin (Fig. 1; Duque-Caro, 1978; Ayala et al., 2012). How-
the southernmost Middle Magdalena Valley Basin, with no measurable contri- ever, the U-Pb results show no clear integration of a large-scale fluvial system
butions to downstream localities in the central and northern parts of the basin until late Miocene time (Fig. 8). In fact, an integrated paleodrainage system
(Figs. 10D–10F). is precluded by the highly variable U-Pb age distributions, which show the
appearance and disappearance of local signals in the pre–late Miocene suc-
Recognition of Paleodrainages and Paleorivers cession. Furthermore, a late Miocene onset is consistent with the depositional
record of the Magdalena delta and submarine fan (Flinch, 2003; Flinch et al.,
U-Pb age distributions from the Middle Magdalena Valley Basin illus- 2003a, 2003b; Rincón et al., 2007; Cadena and Slatt, 2013; Cadena Mendoza,
trate a complex series of transverse paleodrainage alternations followed by 2014). This delayed integration of a through-going Magdalena River (Fig. 10G)
a wholesale drainage reorganization involving the establishment of a major may be due to the role of complex intrabasinal Middle Magdalena Valley
axial paleoriver (Fig. 10). After initial topographic growth of the Eastern Cor­di­ structures or localized subsiding depocenters, or upstream processes farther
llera, the mid- to late Cenozoic evolution of the Middle Magdalena Valley Basin south such as uplift along the margins of the Upper Magdalena Valley ­Basin
was governed by competing sediment sources along both the western and (­Butler and Schamel, 1988; Ramon and Rosero, 2006; Montes et al., 2005;
eastern margins. Rather than a simple mix of source signatures, most Oligo- ­Anderson, 2015).
cene–middle Miocene samples of fluvial basin fill indicate a clear dominance
of either the Eastern Cordillera (e.g., Cagui well; Fig. 5) or the Central Cor-
dillera (e.g., southern Middle Magdalena Valley Basin; Fig. 7). In fact, several CONCLUSIONS
episodes of pronounced sediment influx are recorded over a wide range of
sites (in both the north-south and east-west directions) in which the Middle Despite the complex time-space interaction of Mesozoic–Cenozoic tec­tonics
Magdalena Valley Basin was dominated by either a western source or an east- and sediment accumulation in the Middle Magdalena Valley Basin, robust
ern source (Fig. 9). For example, Oligocene deposits from most of the surface provenance trends are revealed by U-Pb age distributions from both surface
and subsurface localities show the originally bimodal eastern versus western and subsurface basin fill. Published U-Pb age distributions for detrital zircons
signals progressively superseded by a complex integrated signature (Figs. 8 collected from the mouths of modern rivers demonstrate that small to large
and 9). At face value, the lengthy Paleocene to middle Miocene record could be watersheds provide clear signals representative of the competing source re-
largely construed as alternating “see-saw” episodes of shifting paleodrainage gions. Therefore, U-Pb ages within the Middle Magdalena Valley Basin strati-
between two opposing transverse margins (Fig. 10). Although this character- graphic succession allow for discrimination of sediment influxes from the
ization may seem overly simplistic, the provenance record does suggest that contiguous Central Cordillera and Eastern Cordillera, as well as the localized
Middle Magdalena Valley Basin deposystems were highly sensitive to sedi- Santander Massif and San Lucas range.
ment flux and subsidence parameters controlled by processes in the flanking The principal Mesozoic–Cenozoic tectonic events of the northernmost A ­ ndes
ranges. Regardless of the driving mechanisms, the U-Pb results require that are well expressed in the sediment provenance record, including (1) ­Jurassic
these fluvial systems were mostly dominated by either an Eastern Cordilleran growth of extensional subbasins fed by local sources, (2) Cretaceous depo-
or Central Cordilleran source, without large amounts of mixing or downstream sition in a regional postrift setting, and (3) Cenozoic crustal shortening and
dilution. This time-space complexity likely reflects relatively short transport surface uplift of contiguous ranges and localized block uplifts. Finer-scale
distances for the Middle Magdalena Valley Basin (tens to hundreds of kilome- transitions in the Andean shortening history are also preserved. Early uplift in
ters), as compared to systems in the modern Andean foreland (thousands of the flanking Central Cordillera is defined by a mid-Paleocene provenance shift
kilometers). The results further suggest isolated deposystems within a com- reflecting the first arrival of material derived from the west, principally from
partmentalized basin system, without an integrated regional drainage network magmatic arc rocks of Cretaceous age. Subsequent uplift and unroofing of the
(Figs. 10D–10F). Eastern Cordillera are indicated by the late Eocene introduction of considerable
The provenance data provide a refined understanding of the evolution of cratonic Precambrian detritus recycled from Cretaceous–Paleogene strata.
the proto–Magdalena River (Fig. 10G). Disagreement persists over the inferred The subsurface data set provides essential access to the older, pre-Eocene
onset of a regional longitudinal Magdalena River, with previous estimates cen- succession and reveals a complex Cenozoic history of transverse and axial

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1807
Research Paper

paleodrainage. However, several components of the provenance record are Bayona, G., Cardona, A., Jaramillo, C., Mora, A., Montes, C., Valencia, V., Ayala-Calvo, C., Monte­
negro, O., and Ibañez-Mejía, M., 2012, Early Paleogene magmatism in the northern Andes:
poorly defined. A well-documented phase of late Paleocene–middle Eocene
Insights on the effects of oceanic plateau–continent convergence: Earth and Planetary Sci-
uplift of basement highs in the Middle Magdalena Valley Basin appears to be ence Letters, v.  331–332, p.  97–111, doi:​10​.1016​/j​.epsl​.2012​.03​.015​.
aliased in the U-Pb age distributions, likely due to similarities with recycled Bayona, G., Cardona, A., Jaramillo, C., Mora, A., Montes, C., Caballero, V., Mahecha, H., Lamus,
sedimentary signatures. In addition, despite the proximity of the Middle Mag- F., Montenegro, O., Jimenez, G., Mesa, A., and Valencia, V., 2013, Onset of fault reactivation
in the Eastern Cordillera of Colombia and proximal Llanos Basin; response to Caribbean–
dalena Valley Basin to the Andean magmatic arc, strong evidence for a pro-
South American convergence in early Palaeogene time, in Nemčok, M., Mora, A., and Cos-
posed Paleogene magmatic pulse is lacking, and U-Pb ages are of relatively grove, J.W., eds., Thick-Skin–Dominated Orogens: From Initial Inversion to Full Accretion:
limited utility in defining precise depositional ages. The time-space complexity Geological Society of London Special Publication 377, p. 285–314, doi:​10​.1144​/SP377​.5​.
Beranek, L.P., Pease, V., Scott, R.A., and Thomsen, T.B., 2013, Detrital zircon geochronology
of U-Pb age distributions during Oligocene–Miocene time shows alternating
of Ediacaran to Cambrian deep-water strata of the Franklinian Basin, northern Ellesmere
periods of sediment influx from the two opposing transverse margins, without Island, Nunavut: Implications for regional stratigraphic correlations: Canadian Journal of
significant source dilution, consistent with a compartmentalized basin system Earth Sciences, v.  50, p.  1007–1018, doi:​10​.1139​/cjes​-2013​-0026​.
characterized by isolated proximal deposystems. The abrupt up-section intro- Blanco-Quintero, I.F., García-Casco, A., Toro, L.M., Moreno, M., Ruiz, E.C., Vinasco, C.J., Cardona,
A., Lázaro, C., and Morata, D., 2014, Late Jurassic terrane collision in the northwestern mar-
duction of youthful (100–0 Ma) zircons signifies the late Miocene establishment
gin of Gondwana (Cajamarca complex, eastern flank of the Central Cordillera, Colombia):
of an integrated longitudinal drainage network and activation of the proto– International Geology Review, v. 56, p. 1852–1872, doi:​10​.1080​/00206814​.2014​.963710​.
Magda­lena River along the axis of the Middle Magdalena Valley Basin, coinci- Blum, M., and Pecha, M., 2014, Mid-Cretaceous to Paleocene North American drainage reorgani-
dent with growth of the Magdalena submarine fan along the Caribbean margin. zation from detrital zircons: Geology, v. 42, p. 607–610, doi:​10​.1130​/G35513​.1​.
Butler, K., and Schamel, S., 1988, Structure along the eastern margin of the Central Cordillera,
Upper Magdalena Valley, Colombia: Journal of South American Earth Sciences, v. 1, p. 109–
ACKNOWLEDGMENTS 120, doi:​10​.1016​/0895​-9811​(88)90019​-3​.
Caballero, V., Parra, M., and Mora, A., 2010, Levantamiento de la Cordillera Oriental de Colombia
This research was funded by the Instituto Colombiano del Petróleo (ICP), a division of Ecopetrol,
durante el Eocene tardió–Oligoceno temprano: Proveniencia sedimentaria en el sinclinal
and the Jackson School of Geosciences, as part of a collaborative research agreement between
de Nuevo Mundo, cuenca Valle Medio del Magdalena: Boletín de Geologia, v. 32, p. 45–77.
ICP and the University of Texas at Austin. Additional support during the data synthesis phase
Caballero, V., Mora, A., Quintero, I., Blanco, V., Parra, M., Rojas, L.E., Lopez, C., Sánchez, N.,
of this research was provided by National Science Foundation grants EAR-1019857 and 1338694.
Horton, B.K., Stockli, D., and Duddy, I., 2013a, Tectonic controls on sedimentation in an inter­
Many Colombian researchers for the ICP-Ecopetrol project “Cronologia de la deformación en las
montane hinterland basin adjacent to inversion structures: The Nuevo Mundo syncline,
Cuencas Subandinas” shared valuable information during this research. We thank Alejandro
Middle Magdalena Valley, Colombia, in Nemčok, M., Mora, A., and Cosgrove, J.W., eds.,
Bande, Todd Housh, Chris Moreno, Julian Naranjo, Juan Carlos Ramírez, Andres Reyes-Harker,
Thick-Skin–Dominated Orogens: From Initial Inversion to Full Accretion: Geological Society
Jorge Rubiano, C. Javier Sanchez, Jair Sierra, Eliseo Tesón, and Vladimir Torres for beneficial
of London Special Publication 377, p. 315–342, doi:​10​.1144​/SP377​.12​.
discussions. Reviews from François Roure and Raymond Russo helped improve the manuscript.
Caballero, V., Parra, M., Mora, A., López, C., Rojas, L.E., and Quintero, I., 2013b, Factors con-
trolling selective abandonment and reactivation in thick-skin orogens: A case study
REFERENCES CITED in the Magdalena Valley, Colombia, in Nemčok, M., Mora, A., and Cosgrove, J.W., eds.,
Thick-skin-dominated Orogens: From Initial Inversion to Full Accretion: Geological Society,
Aleman, A., and Ramos, V.A., 2000, The northern Andes, in Cordani, U.G., Milani, E.J., Thomaz
London, Special Publication, v. 377, p. 343–367, doi:​10​.1144​/SP377​.4​.
Filho, A., and Campos, D.A., eds., Tectonic Evolution of South America, 31st Inter­national
Cáceres, C., Cediel, F., and Etayo, F., 2003, Mapas de Distribución de Facies Sedimentarias y
Geological Congress: Rio de Janeiro, Brazil, p. 453–480.
Armazón Tectónico de Colombia a Través del Proterozoico y del Fanerozoico: Bogotá, Co-
Allen, P.A., 2008, From landscapes into geological history: Nature, v. 451, no. 7176, p. 274–276,
lombia, Ingeominas, 40 p.
doi:​10​.1038​/nature06586​.
Cadena Mendoza, A.F., 2014, Seismic and sequence stratigraphic interpretation of the area of
Amidon, W.H., Burbank, D.W., and Gehrels, G.E., 2005, U-Pb zircon ages as a sediment mixing
influence of the Magdalena submarine fan, offshore northern Colombia: American Associa-
tracer in the Nepal Himalaya: Earth and Planetary Science Letters, v. 235, p. 244–260, doi:​10​
tion of Petroleum Geologists, Search and Discovery article 50971.
.1016​/j​.epsl​.2005​.03​.019​.
Cadena, A.F., and Slatt, R.M., 2013, Seismic and sequence stratigraphic interpretation of the area
Anderson, V., 2015, Uplift and Exhumation of the Eastern Cordillera of Colombia and its Inter­
actions with Climate [Ph.D. dissertation]: Austin, Texas, University of Texas at Austin, 185 p. of influence of the Magdalena submarine fan, offshore northern Colombia: Interpretation,
Aspden, J.A., McCourt, W.J., and Brook, M., 1987, Geometrical control of subduction-related v.  1, no.  1, p.  SA53–SA74, doi:​10​.1190​/INT​-2013​-0028​.1​.
magmatism: The Mesozoic and Cenozoic plutonic history of western Colombia: Journal of Cardona, A., Chew, D., Valencia, V.A., Bayona, G., Mišković, A., and Ibañez-Mejía, M., 2010, Gren-
the Geological Society of London, v. 144, p. 893–905, doi:​10​.1144​/gsjgs​.144​.6​.0893​. villian remnants in the northern Andes: Rodinian and Phanerozoic paleogeographic per-
Ayala, R.C., Bayona, G., Cardona, A., Ojeda, C., Montenegro, O.C., Montes, C., Valencia, V., and spectives: Journal of South American Earth Sciences, v. 29, p. 92–104, doi:​10​.1016​/j​.jsames​
Jaramillo, C., 2012, The paleogene synorogenic succession in the northwestern Mara­caibo .2009​.07​.011​.
block: Tracking intraplate uplifts and changes in sediment delivery systems: Journal of Chapman, A.D., Saleeby, J.B., Wood, D.J., Piasecki, A., Farley, K.A., Kidder, S., and Ducea, M.N.,
South American Earth Sciences, v.  39, p.  93–111, doi:​10​.1016​/j​.jsames​.2012​.04​.005​. 2012, Late Cretaceous gravitational collapse of the southern Sierra Nevada batholith, Cali-
Bande, A., Horton, B.K., Ramirez, J., Mora, A., Parra, M., and Stockli, D.F., 2012, Clastic deposi- fornia: Geosphere, v. 8, p. 314–341, doi:​10​.1130​/GES00740​.1​.
tion, provenance, and sequence of Andean thrusting in the frontal Eastern Cordillera and Chapman, A.D., Ernst, W.G., Gottlieb, E., Powerman, V., and Metzger, E.P., 2015, Detrital zircon
Llanos foreland basin of Colombia: Geological Society of America Bulletin, v. 124, p. 59–76, geochronology of Neoproterozoic–Lower Cambrian passive-margin strata of the White-Inyo
doi:​10​.1130​/B30412​.1​. Range, east-central California: Implications for the Mojave–Snow Lake fault hypothesis:
Bayona, G., Montes, C., Cardona, A., Jaramillo, C., Ojeda, G., and Valencia, V., 2011, Intraplate Geological Society of America Bulletin, v. 127, p. 926–944, doi:​10​.1130​/B31142​.1​.
subsidence and basin filling adjacent to an oceanic arc–continental collision: A case from Clavijo, J., Mantilla, F.L.C., Pinto, J., Bernal, L., and Pérez, A., 2008, Evolución geológica de la
the southern Caribbean–South America plate margin: Basin Research, v. 23, p. 403–422, doi:​ Serranía de San Lucas, norte del Valle Medio del Magdalena y noroeste de la Cordillera
10​.1111​/j​.1365​-2117​.2010​.00495​.x​. Oriental: Boletín de Geología, Universidad Industrial de Santander, v. 30, no. 1, p. 45–62.

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1808
Research Paper

Colletta, B., Hebrard, F., Letouzey, J., Werner, P., and Rudkiweicz, J.L., 1990, Tectonic style and Fildani, A., Cope, T.D., Graham, S.A., and Wooden, J.L., 2003, Initiation of the Magallanes fore-
crustal structure of the Eastern Cordillera, Colombia, from a balanced cross section, in land basin: Timing of the southernmost Patagonian Andes orogeny revised by detrital zircon
­Letouzey, J., ed., Petroleum and Tectonics in Mobile Belts: Paris, Editions Technip, p. 81–100. provenance analysis: Geology, v. 31, p. 1081–1084, doi:​10​.1130​/G20016​.1​.
Coney, P.J., and Evenchick, C.A., 1994, Consolidation of the American Cordilleras: Journal of Flinch, J., 2003, Structural evolution of the Sinu–Lower Magdalena area (northern Colombia), in
South American Earth Sciences, v. 7, p. 241–262, doi:​10​.1016​/0895​-9811​(94)90011​-6​. Bartolini, C., Buffler, R.T., and Blickwede, J., eds., The Circum–Gulf of Mexico and the Carib­
Cooper, M.A., Addison, F.T., Alvarez, R., Coral, M., Graham, R.H., Hayward, A.B., Howe, S., Mar- bean: Hydrocarbon Habitats, Basin Formation, and Plate Tectonics: American Association of
tinez, J., Naar, J., Peñas, R., Pulham, A.J., and Taborda, A., 1995, Basin development and Petroleum Geologists Memoir 79, p. 776–796.
tectonic history of the Llanos Basin, Eastern Cordillera, and Middle Magdalena Valley, Co- Flinch, J., Amaral, J., Doulcet, A., Mouly, B., Osorio, C., and Pince, J.M., 2003a, Structure of the
lombia: American Association of Petroleum Geologists Bulletin, v. 79, p. 1421–1443. offshore Sinu accretionary wedge, northern Colombia, in Memorias VIII Simposio Bolivari-
Cordani, U.G., Cardona, A., Jimenez, D.M., Liu, D., and Nutman, A.P., 2005, Geochronology of ano, Exploracion Petrolera en las Cuencas Subandinas, p. 76–83.
Proterozoic basement inliers in the Colombian Andes: Tectonic history of remnants of a Flinch, J.F., Amaral, J., Doulcet, A., Mouly, B., Osorio, C., and Pince, J.M., 2003b, Onshore-­
fragmented Grenville belt, in Vaughan, A.P.M., Leat, P.T., and Pankhurst, R.J., eds., Terrane offshore structure of the northern Colombia accretionary complex: American Association of
Processes at the Margins of Gondwana: Geological Society of London Special Publication
Petroleum Geologists, Search and Discovery article 90017.
246, p. 329–346.
Forero-Suarez, A., 1990, The basement of the Eastern Cordillera, Colombia: An allochthonous
Cuadros, F.A., Botelho, N.F., Ordóñez-Carmona, O., and Matteini, M., 2014, Mesoproterozoic crust
terrane in northwestern South America: Journal of South American Earth Sciences, v. 3,
in the San Lucas range (Colombia): An insight into the crustal evolution of the northern
p.  141–151, doi:​10​.1016​/0895​-9811​(90)90026​-W​.
­Andes: Precambrian Research, v.  245, p.  186–206, doi:​10​.1016​/j​.precamres​.2014​.02​.010​.
Gehrels, G.E., 2000, Introduction to detrital zircons studies of Paleozoic and Triassic strata in
Davis, S.J., Dickinson, W.R., Gehrels, G.E., Spencer, J.E., Lawton, T.F., and Carroll, A.R., 2010,
western Nevada and northern California, in Soreghan, M.J., and Gehrels, G.E., eds., Paleo-
The Paleogene California River: Evidence of Mojave-Uinta paleodrainage from U-Pb ages of
zoic and Triassic Paleogeography and Tectonics of Western Nevada and Northern California:
detrital zircons: Geology, v. 38, p. 931–934, doi:​10​.1130​/G31250​.1​.
DeCelles, P.G., Carrapa, B., and Gehrels, G.E., 2007, Detrital zircon U-Pb ages provide provenance Geological Society of America Special Paper 347, p. 1–17.
and chronostratigraphic information from Eocene synorogenic deposits in northwestern Gehrels, G., 2014, Detrital zircon U-Pb geochronology applied to tectonics: Annual Review of
Argen­tina: Geology, v. 35, p. 323–326. Earth and Planetary Sciences, v. 42, p. 127–149, doi:​10​.1146​/annurev​-earth​-050212​-124012​.
DeGraaff-Surpless, K., Mahoney, J.B., Wooden, J.L., and McWilliams, M.O., 2003, Lithofacies Gehrels, G.E., Valencia, V.A., and Ruiz, J., 2008, Enhanced precision, accuracy, efficiency,
control in detrital zircon provenance studies: Insights from the Cretaceous Methow Basin, and spatial resolution of U-Pb ages by laser ablation–multicollector–inductively coupled
southern Canadian Cordillera: Geological Society of America Bulletin, v. 115, p. 899–915, plasma–mass spectrometry: Geochemistry Geophysics Geosystems, v. 9, p. QO3017, doi:​
doi:​10​.1130​/B25267​.1​. 10​.1029​/2007GC001805​.
Dengo, C.A., and Covey, M.C., 1993, Structure of the Eastern Cordillera of Colombia: Implication Goldsmith, R., Marvina, R.F., and Mehnert, H.H., 1971, Radiometric Ages in the Santander ­Massif,
for trap styles and regional tectonics: American Association of Petroleum Geologists Bulle- Eastern Cordillera, Colombian Andes: U.S. Geological Survey Professional Paper 750-D,
tin, v. 77, p. 1315–1337. p. D44–D49.
Dickinson, W.R., and Gehrels, G.E., 2008, Sediment delivery to the Cordilleran foreland basin: Gómez, E., Jordan, T.E., Allmendinger, R.W., Hegarty, K., Kelley, S., and Heizler, M., 2003, Con-
Insights from U-Pb ages of detrital zircons in Upper Jurassic and Cretaceous strata of the trols on architecture of the Late Cretaceous to Cenozoic southern Middle Magdalena Valley
Colorado Plateau: American Journal of Science, v. 308, p. 1041–1082. Basin, Colombia: Geological Society of America Bulletin, v. 115, p. 131–147, doi:​10​.1130​/0016​
Dickinson, W.R., and Gehrels, G.E., 2009, Use of U-Pb ages of detrital zircons to infer maximum -7606​(2003)115​<0131:​COAOTL>2​.0​.CO;2​.
depositional ages of strata: A test against a Colorado Plateau Mesozoic database: Earth and Gómez, E., Jordan, T.E., Allmendinger, R.W., and Cardozo, N., 2005a, Development of the Colom-
Planetary Science Letters, v.  288, p.  115–125, doi:​10​.1016​/j​.epsl​.2009​.09​.013​. bian foreland-basin system as a consequence of diachronous exhumation of the northern
Dickinson, W.R., Lawton, T.F., Pecha, M., Davis, S.J., Gehrels, G.E., and Young, R.A., 2012, Andes: Geological Society of America Bulletin, v. 117, p. 1272–1292, doi:​10​.1130​/B25456​.1​.
Provenance of the Paleogene Colton Formation (Uinta Basin) and Cretaceous–Paleogene Gómez, E., Jordan, T.E., Allmendinger, R.W., Hegarty, K., and Kelley, S., 2005b, Syntectonic Cenozoic
provenance evolution in the Utah foreland: Evidence from U-Pb ages of detrital zircons, sedimentation in the northern middle Magdalena Valley Basin of Colombia and implications for
paleocurrent trends, and sandstone petrofacies: Geosphere, v. 8, p. 854–880, doi:​10​.1130​ exhumation of the northern Andes: Geological Society of America Bulletin, v. 117, p. 547–569.
/GES00763​.1​. Gómez Tapias, J., Nivia Guevara, A., Montes Ramírez, N.E., Jiménez Mejía, D.M., Tejada Avella,
Dörr, W., Grösser, J.R., Rodriguez, G.I., and Kramm, U., 1995, Zircon U-Pb age of the Paramo Rico M.L., Sepúlveda Ospina, M.J., Osorio Naranjo, J.A., Gaona Narváez, T., Diederix, H., Uribe
tonalite-granodiorite, Santander Massif (Cordillera Oriental, Colombia) and its geotectonic Peña, H., and Mora Penagos, M., 2007, Mapa Geológico de Colombia (2nd ed.): Bogotá,
significance: Journal of South American Earth Sciences, v. 8, p. 187–194, doi:​10​.1016​/0895​
Ingeominas (Instituto Colombiano de Geología y Minería), 1:1,000,000 scale.
-9811​(95)00004​-Y​.
Guerrero, J., 1993, Magnetostratigraphy of the Upper Part of the Honda Group and Neiva Forma-
Duque-Caro, H., 1978, Major structural elements and evolution of northwestern Colombia, in
tion: Miocene Uplift of the Colombian Andes [Ph.D. dissertation]: Durham, North Carolina,
Watkins, J.S., Montadert, L., and Dickerson, P.W., eds., Geological and Geophysical Inves-
Duke University, 108 p.
tigations of Continental Margins: American Association of Petroleum Geologists Memoir
He, M., Zheng, H., Bookhagen, B., and Clift, P.D., 2014, Controls on erosion intensity in the Y ­ angtze
29, p. 329–351.
River basin tracked by U-Pb detrital zircon dating: Earth-Science Reviews, v. 136, p. 121–140,
Escalona, A., and Mann, P., 2011, Tectonics, basin subsidence mechanisms, and paleogeography
doi:​10​.1016​/j​.earscirev​.2014​.05​.014​.
of the Caribbean–South American plate boundary zone: Marine and Petroleum Geology,
v.  28, p.  8–39, doi:​10​.1016​/j​.marpetgeo​.2010​.01​.016​. Hoorn, C., Guerrero, J., Sarmiento, G.A., and Lorente, M.A., 1995, Andean tectonics as a cause
Etayo-Serna, F., director, 1985, Proyecto Cretácico: Publicaciones Geológicas Especiales del for changing drainage patterns in Miocene northern South America: Geology, v. 23, p. 237–
Ingeo­minas 16, 352 p. 240, doi:​10​.1130​/0091​-7613​(1995)023​<0237:​ATAACF>2​.3​.CO;2​.
Etayo-Serna, F., Barrero, D., Lozano, H., Espinosa, A., González, H., Orrego, A., Ballesteros, I., Hoorn, C., Wesselingh, F.P., ter Steege, H., Bermudez, M.A., Mora, A., Sevink, J., Sanmartin, I.,
Forero, H., Ramírez, C., Zambrano, F., Duque, H., Vargas, R., Núñez, A., Alvarez, J., Ropaín, Sanchez-Meseguer, A., Anderson, C.L., Figueiredo, J.P., Jaramillo, C., Riff, D., Negri, F.R.,
C., Cardozo, E., Galvis, N., Sarmiento, L., Albers, J., Case, J., Singer, D., Bowen, R., Berger, Hooghiemstra, H., Lundberg, J., Stadler, T., Sarkinen, T., and Antonelli, A., 2010, Amazonia
B., Cox, D., and Hodges, C., 1983, Mapa de Terrenos Geológicos de Colombia: Publicaciones through time: Andean uplift, climate change, landscape evolution, and biodiversity: Science,
Geológicas Especiales del Ingeominas 14, 235 p. v.  330, p.  927–931, doi:​10​.1126​/science​.1194585​.
Fabre, A., 1987, Tectonique et géneration d’hydrocarbures: Un modèle de l’evolution de la Cor- Horton, B.K., 2012, Cenozoic evolution of hinterland basins in the Andes and Tibet, in Busby,
dillère Orientale de Colombie et du Bassin de Llanos pendant le Crétacé et le Tertiaire: Ar- C., and Azor, A., eds., Tectonics of Sedimentary Basins: Recent Advances: Oxford, UK,
chives des Sciences Genève, v. 40, p. 145–190. ­Wiley-Blackwell, p. 427–444.

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1809
Research Paper

Horton, B.K., Parra, M., Saylor, J.E., Nie, J., Mora, A., Torres, V., Stockli, D.F., and Strecker, M.R., Piedras-Girardot area, Colombia: Tectonophysics, v. 399, p. 221–250, doi:​10​.1016​/j​.tecto​.2004​
2010a, Resolving uplift of the northern Andes using detrital zircon age signatures: GSA ­Today, .12​.024​.
v. 20, no. 7, p. 4–10, doi:​10​.1130​/GSATG76A​.1​. Mora, A., Gaona, T., Kley, J., Montoya, D., Parra, M., Quiroz, L.I., Reyes, G., and Strecker, M.,
Horton, B.K., Saylor, J.E., Nie, J., Mora, A., Parra, M., Reyes-Harker, A., and Stockli, D.F., 2010b, 2009, The role of inherited extensional fault segmentation and linkage in contractional oro-
Linking sedimentation in the northern Andes to basement configuration, Mesozoic exten- genesis: A reconstruction of Lower Cretaceous inverted rift basin in the Eastern Cordillera of
sion, and Cenozoic shortening: Evidence from detrital zircon U-Pb ages, Eastern Cordillera, Colombia: Basin Research, v.  21, p.  111–137, doi:​10​.1111​/j​.1365​-2117​.2008​.00367​.x​.
Colombia: Geological Society of America Bulletin, v. 122, p. 1423–1442, doi:​10​.1130​/B30118​.1​. Mora, A., Horton, B.K., Mesa, A., Rubiano, J., Ketcham, R.A., Parra, M., Blanco, V., Garcia, D., and
Horton, B.K., Perez, N.D., Fitch, J.D., and Saylor, J.E., 2015, Punctuated shortening and sub­ Stockli, D.F., 2010, Migration of Cenozoic deformation in the Eastern Cordillera of Colombia
sidence in the Altiplano Plateau of southern Peru: Implications for early Andean mountain interpreted from fission track results and structural relationships: Implications for petroleum
building: Lithosphere, v. 7, p. 117–137, doi:​10​.1130​/L397​.1​. systems: American Association of Petroleum Geologists Bulletin, v. 94, p. 1543–1580, doi:​
Ibañez-Mejia, M., Ruiz, J., Valencia, V.A., Cardona, A., Gehrels, G.E., and Mora, A.R., 2011, The 10​.1306​/01051009111​.
Putumayo orogen of Amazonia and its implications for Rodinia reconstructions: New U-Pb Mora, A., Reyes-Harker, A., Rodriguez, G., Tesón, E., Ramirez-Arias, J.C., Parra, M., Caballero, V.,
geochronological insights into the Proterozoic tectonic evolution of northwestern South Mora, J.P., Quintero, I., Valencia, V., Ibañez, M., Horton, B.K., and Stockli, D.F., 2013, Inver-
America: Precambrian Research, v.  191, p.  58–77, doi:​10​.1016​/j​.precamres​.2011​.09​.005​. sion tectonics under increasing rates of shortening and sedimentation: Cenozoic example
Ingersoll, R.V., 1990, Actualistic sandstone petrofacies: Discriminating modern and ancient from the Eastern Cordillera of Colombia, in Nemčok, M., Mora, A., and Cosgrove, J.W., eds.,
source rocks: Geology, v.  18, p.  733–736, doi:​10​.1130​/0091​-7613​(1990)018​<0733:​ASPDMA>2​ Thick-Skin–Dominated Orogens: From Initial Inversion to Full Accretion: Geological Society
.3​.CO;2​. of London Special Publication 377, p. 411–442, doi:​10​.1144​/SP377​.6​.
Ingersoll, R.V., Kretchmer, A.G., and Valles, P.K., 1993, The effect of sampling scale on actual- Morales, L., 1958, General geology and oil occurrences of Middle Magdalena Valley, Colombia,
istic sandstone petrofacies: Sedimentology, v. 40, p. 937–953, doi:​10​.1111​/j​.1365​-3091​.1993​ in Weeks, L.G., ed., Habitat of Oil: American Association of Petroleum Geologists Special
.tb01370​.x​. Publication 18, p. 641–695.
Irving, E.M., 1975, Structural Evolution of the Northernmost Andes, Colombia: U.S. Geological Moreno, C.J., Horton, B.K., Caballero, V., Mora, A., Parra, M., and Sierra, J., 2011, Depositional
Survey Professional Paper 846, 47 p. and provenance record of the Paleogene transition from foreland to hinterland basin evolu-
Jaramillo, C.A., Rueda, M., and Torres, V., 2011, A palynological zonation of the Cenozoic of the tion during Andean orogenesis, northern Middle Magdalena Valley Basin, Colombia: Jour-
Llanos and Llanos foothills of Colombia: Palynology, v. 35, p. 46–84, doi:​10​.1080​/01916122​ nal of South American Earth Sciences, v. 32, p. 246–263, doi:​10​.1016​/j​.jsames​.2011​.03​.018​.
.2010​.515069​. Moreno, N., Silva, A., Mora, A., Tesón, E., Quintero, I., Rojas, L.E., Lopez, C., Blanco, V., Cas­
Kammer, A., and Sánchez, J., 2006, Early Jurassic rift structures associated with the Soapaga tellanos, J., Sanchez, J., Osorio, L., Namson, J., Stockli, D., and Casallas, W., 2013, Inter­
and Boyacá faults of the Eastern Cordillera, Colombia: Sedimentological inferences and re- action between thin- and thick-skinned tectonics in the foothill areas of an inverted graben:
gional implications: Journal of South American Earth Sciences, v. 21, p. 412–422, doi:​10​.1016​ The Middle Magdalena Foothill belt, in Nemčok, M., Mora, A., and Cosgrove, J.W., eds.,
/j​.jsames​.2006​.07​.006​. Thick-Skin–Dominated Orogens: From Initial Inversion to Full Accretion: Geological Society
Latrubesse, E.M., 2015, Large rivers, megafans and other Quaternary avulsive fluvial systems: A of London Special Publication 377, p. 221–255, doi:​10​.1144​/SP377​.18​.
potential “who’s who” in the geological record: Earth-Science Reviews, v. 146, p. 1–30, doi:​ Naranjo-Vesga, J.F., Gómez Gutiérrez, P.D., Gélvez Llanez, J.R., Duque Pardo, N.E., and Moreno
10​.1016​/j​.earscirev​.2015​.03​.004​. Gómez, N.R., 2013, Methodology proposal for correlation studies of fluvial sediments based
Lawrence, R.L., Cox, R., Mapes, R.W., and Coleman, D.S., 2011, Hydrodynamic fractionation of on petrographic and lithogeochemical analysis: Example of its application on Cenozoic rock
zircon age populations: Geological Society of America Bulletin, v. 123, p. 295–305, doi:​10​ from the Lisama Formation (Middle Magdalena Valley, Colombia): Ciencia: Tecnología y
.1130​/B30151​.1​. ­Futuro, v. 5, p. 19–45.
Lawton, T.F., Hunt, G.J., and Gehrels, G.E., 2010, Detrital zircon record of thrust belt unroofing Nie, J., Horton, B.K., Mora, A., Saylor, J.E., Housh, T.B., Rubiano, J., and Naranjo, J., 2010, Track-
in Lower Cretaceous synorogenic conglomerates, central Utah: Geology, v. 38, p. 463–466, ing exhumation of Andean ranges bounding the Middle Magdalena Valley Basin, Colombia:
doi:​10​.1130​/G30684​.1​. Geology, v.  38, p.  451–454, doi:​10​.1130​/G30775​.1​.
Lewis, C.J., and Sircombe, K.N., 2013, Use of U-Pb geochronology to delineate provenance of Nie, J., Horton, B.K., Saylor, J.E., Mora, A., Mange, M., Garzione, C.N., Basu, A., Moreno, C.J.,
North West Shelf sediments, Australia, in Keep, M., and Moss, S.J., eds., The Sedimentary Caballero, V., and Parra, M., 2012, Integrated provenance analysis of a convergent retroarc
Basins of Western Australia IV: Proceedings of the Petroleum Exploration Society of Aus­ foreland system: U-Pb ages, heavy minerals, Nd isotopes, and sandstone compositions of
tralia Symposium 2013: Perth, Australia, Petroleum Exploration Society of Australia, p. 1–27. the Middle Magdalena Valley Basin, northern Andes, Colombia: Earth-Science Reviews,
Link, P.K., Fanning, C.M., and Beranek, L.P., 2005, Reliability and longitudinal change of d ­ etrital-­ v.  110, p.  111–126, doi:​10​.1016​/j​.earscirev​.2011​.11​.002​.
zircon age spectra in the Snake River system, Idaho and Wyoming: An example of repro- Ordóñez-Carmona, O., Restrepo Álvarez, J.J., and Pimentel, M.M., 2006, Geochronological and
ducing the bumpy barcode: Sedimentary Geology, v. 182, p. 101–142, doi:​10​.1016​/j​.sedgeo​ isotopical review of pre-Devonian crustal basement of the Colombian Andes: Journal of
.2005​.07​.012​. South American Earth Sciences, v.  21, p.  372–382, doi:​10​.1016​/j​.jsames​.2006​.07​.005​.
Mackey, G.N., Horton, B.K., and Milliken, K.L., 2012, Provenance of the Paleocene–Eocene Wilcox Parra, M., Mora, A., Jaramillo, C., Strecker, M.R., Sobel, E.R., Quiroz, L.I., Rueda, M., and Torres,
Group, western Gulf of Mexico basin: Evidence for integrated drainage of the southern Lara­ V., 2009a, Orogenic wedge advance in the northern Andes: Evidence from the Oligocene–
mide Rocky Mountains and Cordilleran arc: Geological Society of America Bulletin, v. 124, Miocene sedimentary record of the Medina Basin, Eastern Cordillera, Colombia: Geological
p.  1007–1024, doi:​10​.1130​/B30458​.1​. Society of America Bulletin, v. 121, p. 780–800, doi:​10​.1130​/B26257​.1​.
Mantilla Figueroa, L.C., Bissig, T., Valencia, V., and Hart, C.J.R., 2013, The magmatic history of the Parra, M., Mora, A., Sobel, E.R., Strecker, M.R., and González, R., 2009b, Episodic orogenic-front
Vetas-California mining district, Santander Massif, Eastern Cordillera, Colombia: Journal of migration in the northern Andes: Constraints from low-temperature thermochronology in
South American Earth Sciences, v.  45, p.  235–249, doi:​10​.1016​/j​.jsames​.2013​.03​.006​. the Eastern Cordillera, Colombia: Tectonics, v. 28, p. TC4004, doi:​10​.1029​/2008TC002423​.
McCourt, W.J., Aspden, J.A., and Brook, M., 1984, New geological and geochronological data Parra, M., Mora, A., Lopez, C., Rojas, L.E., and Horton, B.K., 2012, Detecting earliest shortening
from the Colombian Andes: Continental growth by multiple accretion: Journal of the Geo- and deformation advance in thrust belt hinterlands: Example from the Colombian Andes:
logical Society of London, v. 141, p. 831–845, doi:​10​.1144​/gsjgs​.141​.5​.0831​. Geology, v.  40, p.  175–178, doi:​10​.1130​/G32519​.1​.
Métivier, F., and Gaudemer, Y., 1999, Stability of output fluxes of large rivers in South and East Potter, P.E., 1978, Petrology and chemistry of modern big river sands: The Journal of Geology,
Asia during the last 2 million years: Implications on floodplain processes: Basin Research, v.  86, p.  423–449, doi:​10​.1086​/649711​.
v.  11, p.  293–303, doi:​10​.1046​/j​.1365​-2117​.1999​.00101​.x​. Rainbird, R.H., Stern, R.A., Rayner, N., and Jefferson, C.W., 2007, Age, provenance, and regional
Montes, C., Hatcher, R.D., Jr., and Restrepo-Pace, P.A., 2005, Tectonic reconstruction of the north- correlation of the Athabasca Group, Saskatchewan and Alberta, constrained by igneous and
ern Andean blocks: Oblique convergence and rotations derived from the kinematics of the detrital zircon geochronology, in Jefferson, C.W., and Delaney, G., eds., EXTECH IV: Geology

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1810
Research Paper

and Uranium EXploration TECHnology of the Proterozoic Athabasca Basin, Saskatchewan of Colombia, in Nemčok, M., Mora, A., and Cosgrove, J.W., eds., Thick-Skin–Dominated Oro-
and Alberta: Geological Survey of Canada Bulletin 588, p. 193–209. gens: From Initial Inversion to Full Accretion: Geological Society of London Special Publica-
Ramon, J.C., and Rosero, A., 2006, Multiphase structural evolution of the western margin of the tion 377, p.  369–409, doi:​10​.1144​/SP377​.15​.
Girardot subbasin, Upper Magdalena Valley, Colombia: Journal of South American Earth Stacey, J.S., and Kramers, J.D., 1975, Approximation of terrestrial lead isotope evolution by a
Sciences, v.  21, p.  493–509, doi:​10​.1016​/j​.jsames​.2006​.07​.012​. two-stage model: Earth and Planetary Science Letters, v. 26, p. 207–221.
Reyes-Harker, A., Ruiz-Valdivieso, C.F., Mora, A., Ramírez-Arias, J.C., Rodriguez, G., de la Parra, Taboada, A., Rivera, L.A., Fuenzalida, A., Cisternas, A., Philip, H., Bijwaard, H., Olaya, J., and
F., Caballero, V., Parra, M., Moreno, N., Horton, B.K., Saylor, J.E., Silva, A., Valencia, V., Rivera, C., 2000, Geodynamics of the northern Andes: Subductions and intracontinental
Stockli, D., and Blanco, V., 2015, Cenozoic paleogeography of the Andean foreland and retro­ deformation (Colombia): Tectonics, v. 19, p. 787–813, doi:​10​.1029​/2000TC900004​.
arc hinter­land of Colombia: American Association of Petroleum Geologists Bulletin, v. 99, Tesón, E., Mora, A., Silva, A., Namson, J., Teixell, A., Castellanos, J., Casallas, W., Julivert, M.,
p.  1407–1453, doi:​10​.1306​/06181411110​. Taylor, M., Ibañez-Mejia, M., and Valencia, V., 2013, Interrelationships among Mesozoic
Rincón, D., Arenas, J., Cuartas, C., Cárdenas, A., Molinares, C., Caicedo, C., and Jaramillo, C., ­graben distribution, stress, amount of shortening and structural style in the Eastern Cor­di­
2007, Eocene–Pliocene planktonic foraminifera biostratigraphy from the continental margin llera of Colombia, in Nemčok, M., Mora, A., and Cosgrove, J.W., eds., Thick-Skin–Dominated
of the southwest Caribbean: Stratigraphy, v. 4, p. 261–311. Orogens: From Initial Inversion to Full Accretion: Geological Society of London Special Pub-
Roeder, D., and Chamberlain, R.L., 1995, Eastern Cordillera of Colombia: Jurassic–Neogene lication 377, p.  257–283, doi:​10​.1144​/SP377​.10​.
crustal evolution, in Tankard, A.J., Suarez, S.R., and Welsink, H.J., eds., Petroleum Basins Toro, J., Roure, F., Bordas-Le Floch, N., Le Cornec-Lance, S., and Sassi, W., 2004, Thermal and
of South America: American Association of Petroleum Geologists Memoir 62, p. 633–645. kinematic evolution of the Eastern Cordillera fold and thrust belt, Colombia, in Swennen,
Sánchez, J., Horton, B.K., Tesón, E., Mora, A., Ketcham, R.A., and Stockli, D.F., 2012, Kinematic R., Roure, F., and Granath, J., eds., Deformation, Fluid Flow and Reservoir Appraisal in Fore-
evolution of Andean fold-thrust structures along the boundary between the Eastern Cor­di­ land Fold and Thrust Belts: American Association of Petroleum Geologists Hedberg Series
llera and Middle Magdalena Valley Basin, Colombia: Tectonics, v. 31, p. TC3008, doi:​10​.1029​ Memoir 1, p. 79–115.
/2011TC003089​. Toussaint, J.F., and Restrepo, J.J., 1994, The Colombian Andes during Cretaceous times, in
Sarmiento Rojas, L.F., 2001, Mesozoic Rifting and Cenozoic Basin Inversion History of the Eastern ­Salfity, J.A., ed., Cretaceous Tectonics of the Andes: Earth Evolution Sciences: Braunsch-
Cordillera, Colombian Andes: Inferences from Tectonic Models [Ph.D. dissertation]: Amster- weig, Germany, Vieweg, p. 61–100.
dam, Netherlands, Vrije Universiteit, 295 p. Van Houten, F.B., 1976, Late Cenozoic volcaniclastic deposits, Andean foredeep, Colombia: Geo-
Sarmiento-Rojas, L.F., Van Wess, J.D., and Cloetingh, S., 2006, Mesozoic transtensional basin logical Society of America Bulletin, v. 87, p. 481–495, doi:​10​.1130​/0016​-7606​(1976)87​<481:​
history of the Eastern Cordillera, Colombian Andes: Inferences from tectonic models: Jour- LCVDAF>2​.0​.CO;2​.
nal of South American Earth Sciences, v. 21, p. 383–411, doi:​10​.1016​/j​.jsames​.2006​.07​.003​. Van Houten, F.B., and Travis, R.B., 1968, Cenozoic deposits, Upper Magdalena Valley, Colombia:
Saylor, J.E., and Sundell, K.E., II, 2014, Statistics of large detrital geochronology datasets: Wash- American Association of Petroleum Geologists Bulletin, v. 52, p. 675–702.
ington, D.C., American Geophysical Union, Fall Meeting, abstract EP21D–3563. Vermeesch, P., 2013, Multi-sample comparison of detrital age distributions: Chemical Geology,
Saylor, J.E., Horton, B.K., Nie, J., Corredor, J., and Mora, A., 2011, Evaluating foreland basin par- v.  341, p.  140–146, doi:​10​.1016​/j​.chemgeo​.2013​.01​.010​.
titioning in the northern Andes using Cenozoic fill of the Floresta Basin, Eastern Cordillera, Villagómez, D., and Spikings, D., 2013, Thermochronology and tectonics of the Central and West-
Colombia: Basin Research, v.  23, p.  377–402, doi:​10​.1111​/j​.1365​-2117​.2010​.00493​.x​. ern Cordilleras of Colombia: Early Cretaceous–Tertiary evolution of the northern Andes:
Saylor, J.E., Horton, B.K., Stockli, D.F., Mora, A., and Corredor, J., 2012a, Structural and thermo- Lithos, v.  160–161, p.  228–249, doi:​10​.1016​/j​.lithos​.2012​.12​.008​.
chronological evidence for Paleogene basement-involved shortening in the axial Eastern Villagómez, D., Spikings, R., Magna, T., Kammer, A., Winkler, W., and Beltran, A., 2011, Geo-
Cordillera, Colombia: Journal of South American Earth Sciences, v. 39, p. 202–215, doi:​10​ chronology, geochemistry and tectonic evolution of the Western and Central Cordilleras of
.1016​/j​.jsames​.2012​.04​.009​. Colombia: Lithos, v.  125, p.  875–896, doi:​10​.1016​/j​.lithos​.2011​.05​.003​.
Saylor, J.E., Stockli, D.F., Horton, B.K., Nie, J., and Mora, A., 2012b, Discriminating rapid exhuma- Villamil, T., 1999, Campanian–Miocene tectonostratigraphy, depocenter evolution and basin
tion from syndepositional volcanism using detrital zircon double dating: Implications for the development of Colombia and western Venezuela: Palaeogeography, Palaeoclimatology,
tectonic history of the Eastern Cordillera, Colombia: Geological Society of America Bulletin, Palaeo­ecology, v.  153, p.  239–275, doi:​10​.1016​/S0031​-0182​(99)00075​-9​.
v.  124, p.  762–779, doi:​10​.1130​/B30534​.1​. Vinasco, C.J., Cordani, U.G., González, H., Weber, M., and Pelaez, C., 2006, Geochronological,
Saylor, J.E., Knowles, J.N., Horton, B.K., Nie, J., and Mora, A., 2013, Mixing of source populations isotopic, and geochemical data from Permo-Triassic granitic gneisses and granitoids of the
recorded in detrital zircon U-Pb age spectra of modern river sands: The Journal of Geology, Colombian central Andes: Journal of South American Earth Sciences, v. 21, p. 355–371, doi:​
v.  121, p.  17–33, doi:​10​.1086​/668683​. 10​.1016​/j​.jsames​.2006​.07​.007​.
Schamel, S., 1991, Middle and Upper Magdalena Basins, Colombia, in Biddle, K.T., ed., Active Zhang, J.Y., Yin, A., Liu, W.C., Wu, F.Y., Lin, D., and Grove, M., 2012, Coupled U-Pb dating and
Margin Basins: American Association of Petroleum Geologists Memoir 52, p. 283–301. Hf isotopic analysis of detrital zircon of modern river sand from the Yalu River (Yarlung
Silva, A., Mora, A., Caballero, V., Rodríguez, G., Ruiz, C.A., Moreno, N.R., Parra, M., Ramírez- Tsangpo) drainage system in southern Tibet: Constraints on the transport processes and
Arias, J.C., Ibañez, M., and Quintero, I., 2013, Basin compartmentalization and drainage evolution of Himalayan rivers: Geological Society of America Bulletin, v. 124, p. 1449–1473,
evolution during rift inversion: Evidence from multiple techniques in the Eastern Cordillera doi:​10​.1130​/B30592​.1​.

GEOSPHERE  |  Volume 11  |  Number 6 Horton et al.  |  Provenance and paleogeographic reconstruction of the Magdalena Valley, northern Andes 1811
View publication stats

You might also like