You are on page 1of 22

Research Paper

GEOSPHERE Andean topographic growth and basement uplift in southern


Colombia: Implications for the evolution of the Magdalena, Orinoco,
GEOSPHERE; v. 12, no. 4
and Amazon river systems
doi:10.1130/GES01294.1
Veronica J. Anderson1,2, Brian K. Horton1,2, Joel E. Saylor 3, Andrés Mora4, Eliseo Tesón4, Daniel O. Breecker1, and Richard A. Ketcham1
8 figures; 5 tables; 8 supplemental files 1
Department of Geological Sciences, Jackson School of Geosciences, University of Texas at Austin, 2275 Speedway Stop C9000, Austin, Texas 78712, USA
2
Institute for Geophysics, Jackson School of Geosciences, University of Texas at Austin, J.J. Pickle Research Campus, Building 196, 10100 Burnet Road, Austin, Texas 78758, USA
3
Department of Earth and Atmospheric Sciences, University of Houston, Science & Research Building 1, 3507 Cullen Boulevard, Houston, Texas 77204, USA
CORRESPONDENCE:  horton@​jsg​.utexas​.edu 4
Ecopetrol–Instituto Colombiano del Petróleo, Km 7 Autopista Bucaramanga-Piedecuesta, Piedecuesta, Santander, Colombia

CITATION:  Anderson, V.J., Horton, B.K., Saylor, J.E.,


Mora, A., Tesón, E., Breecker, D.O., and ­Ketcham,
R.A., 2016, Andean topographic growth and base- ABSTRACT INTRODUCTION
ment uplift in southern Colombia: Implications for the
evolution of the Magdalena, O ­ rinoco, and ­Amazon
Surface uplift of the Garzón Massif in the northern Andes formed a criti­ The northern Andes form the chief orographic barrier separating the mod-
river systems: Geosphere, v. 12, no. 4, p. 1235–1256,
doi:10.1130/GES01294.1. cal orographic barrier (2500–3000 m elevation) that generated a deep rain ern Orinoco, Amazon, and Magdalena River watersheds (Fig. 1). Although
shadow and strongly influenced the evolution of the largest river systems these river systems have collectively drained the northern half of South Amer-
Received 15 November 2015 draining northern South America. This basement massif and its correspond­ ica, governed significant Andean erosion, and influenced Caribbean and At-
Revision received 14 April 2016 ing foreland basement high define the headwaters and drainage divides of lantic ocean chemistry, their genesis remains highly debated (e.g., Figueiredo
Accepted 6 May 2016
the Amazon, Orinoco, and Magdalena Rivers. Despite its pivotal role, the et al., 2009; Hoorn et al., 2010; Sacek, 2014). During Paleogene time, most of
Published online 16 June 2016
exhumation history of the Garzón Massif and its relationships to the struc­ northern South America drained northward into the Caribbean Sea, forming a
tural evolution of the broader Eastern Cordillera fold-thrust belt remain large delta in the Maracaibo Basin of Venezuela (Díaz de Gamero, 1996; Esca­
unclear. The northern Andes underwent major Cenozoic shortening, with lona and Mann, 2006; Mann et al., 2006). Today, the Orinoco River empties
considerable thin-skinned and thick-skinned deformation and topographic into the equatorial Atlantic >1000 km farther east, and the Magdalena River of
development in the Eastern Cordillera focused during late Miocene time. Colombia is now the largest single contributor of sediment to the Caribbean
On the basis of widespread coarse-grained nonmarine sedimentation, pre­ Sea (Fig. 1). The disruption of the original northward drainage configuration
vious studies have inferred that uplift of the Garzón Massif began during the and the establishment of independent Orinoco, Amazon, and Magdalena sys-
late Miocene, coincident with rapid elevation gain elsewhere in the Eastern tems are critically tied to basement uplift in the Eastern Cordillera fold-thrust
Cordillera. belt and proximal Andean foreland of Colombia.
We take an integrated, multiproxy approach to better reconstruct Andean Paleocurrents, palynological assemblages, and mammal fossils suggest
topographic growth and distinguish between exhumation and surface uplift that Amazon River capture of the former southern Orinoco drainage system
of the Garzón Massif. We present new U-Pb detrital zircon provenance data, and isolation from the inter-Andean Magdalena drainage system (Fig. 1) was
sandstone petrographic data, and paleoprecipitation data from upper Mio­ driven primarily by Neogene uplift of the Eastern Cordillera and Mérida Andes
cene clastic fill of the Neiva Basin within the adjacent Upper Magdalena Valley (Hoorn, 1994; Hoorn et al., 1995; Díaz de Gamero, 1996). In contrast, some re-
of the modern hinterland. In addition, six new apatite fission track (AFT) ages constructions of the northern Andean and southern Caribbean margin propose
from the central segment of the northeast-trending Garzón Massif (Jurassic that an independent Magdalena River was already established by 40–30 Ma
granite and Proterozoic gneiss and schist) directly constrain its Neogene exhu­ (Gómez et al., 2005; Escalona and Mann, 2006). Although uplift of the Mérida
mation history. The results indicate that early exhumation may have initiated Andes in Venezuela (Fig. 1) was long considered the driving mechanism be-
by ca. 12.5 Ma, but a substantial orographic barrier was not fully established hind the delineation of the Orinoco and Magdalena systems, more recent work
until ca. 6–3 Ma, when >1 km/m.y. of material was exhumed. Thermal his­ recognizes the fundamental influence of foreland basement arches, thrust sys-
tory modeling of the AFT data suggests diminished exhumation thereafter tems, and sediment accumulation on drainage evolution (Mora et al., 2010a;
(3–0 Ma), during latest Cenozoic oblique Nazca–South America convergence. Roddaz et al., 2010; Silva et al., 2013; Caballero et al., 2013; Horton et al., 2015a).
This exhumation history is consistent with paleontological data suggesting Understanding the evolving drainage configuration of northern South Amer-
For permission to copy, contact Copyright late Miocene divergence of the three river systems, with associated trans­ ica has fueled a vigorous debate about the formation of the modern Amazon
Permissions, GSA, or editing@geosociety.org. conti­nental drainage of the Amazon River. River, with estimates ranging from a middle Miocene to a Pleistocene onset of
© 2016 Geological Society of America

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1235
Research Paper

Magdalena Gulf of
Delta Venezuela Bonaire Basin
Orinoco

MAGDALENA Delta
10°N Lake
DRAINAGE Maracaibo
Río
SYSTEM C

Río Gu á rico
es

oje
uca
d
An

d
o
inoc

es
a

Ca
VENEZUELA Or

d a lena
R ío id Río
ér
M Río Apur e

í o Mag
ORINOCO
DRAINAGE SYSTEM

e R
y

ato

a
iller
R ío Atr

ra
ll e

all
di

aV
Figure 1. Shaded relief map of north­

l Cord
l e ra

or
western South America showing m ­ ajor

dalen

Río Orin oco


5°N

nC
tectonic provinces, drainage systems,

o rd il
et a a n d d ra i n a ge d i v i d e s a m o n g t h e

t er
t ra
R ío M

M ag
­Orinoco, Amazon, and Magdalena river
COLOMBIA

Eas
Cen
C
systems. Location of Garzón Massif

ern
(Fig. 2) is shown.
Macarena

West
Guainía
high Río

Vau

Río
p es a

Casqui
Figure 2 rch

co
ran
ar e
Río Río V
au

Río B
p és

Aj
Equator jú

a
Río Napo R í o Negro
R ío
Caquetá AMAZON
ECUADOR DRAINAGE SYSTEM BRAZIL
PERU Río Japurá
0 100 200 km
Río Pu t umayo
80°W 75 °W 70°W 65°W

transcontinental drainage (e.g., Potter, 1997; Campbell et al., 2006; Campbell, influx of basement-derived material, and apatite fission track (AFT) thermo-
2010; Latrubesse et al., 2010). New constraints on the timing, mechanics, and chronometry to constrain the timing and pace of exhumation. In addition, we
geomorphology of the transition from mid-Cenozoic to modern drainage con- use two paleosol-based paleoclimatic proxies to place rough bounds on initial
figurations are critical to reconstructions of past rivers and their influence on development of the modern topographic barrier and its corresponding oro-
the geological and biological dynamics of South America (e.g., Hoorn et al., graphic rain shadow.
2010; Ribas et al., 2012; Baker et al., 2014).
In this study we utilize a multiproxy approach to constrain topographic
growth of the Eastern Cordillera in southern Colombia and the resulting estab- GEOLOGIC SETTING
lishment of an isolated Magdalena River system. We focus our study on the
Garzón Massif, the 2500–3000-m-high basement uplift that bounds the Andean The northern Andes contain the Cenozoic structural and sedimentary rec­
headwaters of the Magdalena watershed, and whose foreland counterpart, the ord of advancing fold-thrust systems in the Central and Eastern Cordilleras. In
Macarena high and broader Vaupes arch, forms the Amazon-Orinoco drainage Colombia, the Central Cordillera demarcates the latest Cretaceous–Paleogene
divide (Fig. 1). We constrain the Neogene exhumation age of this massif by thrust front, whereas the Eastern Cordillera represents a younger bivergent
studying clastic deposits of the Neiva hinterland basin in the adjacent Upper contractional belt that reactivated a Mesozoic rift system in a combination of
Magdalena Valley (Fig. 2A) using sediment provenance techniques to track the thin-skinned ramp-flat thrust systems and thick-skinned basement-involved

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1236
Research Paper

A Neiva Basin and B La Venta Area


Garzón Massif 75.3 °W 75.2 °W 75.1 °W 75 °W
La
Victoria
La V 3.4 °N
ent
a
Orinoco

Ca b
lt
Villavieja

dr gda Chusma Fau


drainage

ere
aR
i ve
3°N

ge a

r
La Victoria

na n
ai le
Neiva
3.3 °N

t
a

l
1

au
M

sF
La V
i ent

ra
ec

er
a

M agd alena Riv


Alg
Hobo 2

n
Gigante
3
Villavieja 3.2 °N

Garzón 4

sif
as
M
ón
2°N
5

rz

A rain
Ga

d
m a
0 2.5 5 7.5 10 km

az g
3.1 °N

on e
Florencia
C Gigante Area and AFT Sampling Transect
-
r
oi

rv
se
0 25 50 75 100 km 2.6 °N

Re
2

n ia
GM06

e ta
Hobo GM05
B GM01
76 °W 75 °W GM04
GM02
Sedimentary Upper Cretaceous Metamorphic GM03

Quaternary Alluvium Lower Cretaceous Paleozoic 2.5 °N


Proterozoic Jurassic 2
Pleistocene Alluvium Jurassic granodiorite
Igneous Proterozoic
Neogene Gigante Fm. Upper Triassic
gneiss/schist
Neogene Honda Gp. Lower Triassic Neogene

lt
au
Jurassic 6

sF
Eocene-Oligocene Devonian-Permian

ra
Gigante 0 2.5 5 7.5 10 km

eci
2.4°N
Maastrichtian- Cambrian-Silurian 2

g
Paleocene

Al
drainage divide
75.6 °W 75.5 °W 75.4 °W 75.3 °W 75.2°W

Figure 2. Geologic maps of the Upper Magdalena Valley and flanking Central Cordillera and Eastern Cordillera (after Gómez Tapias et al., 2007; Marquínez and Velandia, 2001), showing location of
apatite fission track (AFT) sample suite from this study (solid ellipse) and previous studies (dashed ellipses). (A) Map of the Garzón Massif and Neiva Basin. (B) Map of the La Venta area, with basin
sample transects (1–5) spanning the Honda Group shown in red. (C) Map of the Gigante area, with basin sample transect (6) and AFT samples (GM01–GM06) shown in red.

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1237
Research Paper

structures (Colletta et al., 1990; Dengo and Covey, 1993; Casero et al., 1997; refine these estimates by integrating new results capable of providing a more
Ramon and Rosero, 2006; Mora et al., 2006; Parra et al., 2009, 2012; Mora et al., complete picture of the Garzón Massif’s topographic history and the influence
2013; Wolaver et al., 2015). The north- to northeast-trending Central and East- of basement block uplifts on large drainage systems.
ern Cordilleras are separated by the Magdalena Valley, which narrows south-
ward toward the Magdalena River headwaters in southern Colombia, where
the two range systems intersect (Fig. 1). BASIN STRATIGRAPHY
The Upper Magdalena Valley is bounded to the west by the Chusma fault
system (Fig. 2A), an Eocene–Oligocene thrust system defining the eastern We studied the Neogene succession at two field localities in the Neiva Ba-
front of the Central Cordillera (Butler and Schamel, 1988; Mojica and Franco, sin within the Upper Magdalena Valley of southern Colombia and obtained
1990; van der Wiel and van den Bergh, 1992; Sarmiento and Rangel, 2004). The samples along six field transects spanning depositional ages of 13.8–6.4 Ma
eastern boundary of the Upper Magdalena Valley is defined by the west- to (Figs. 2 and 3; Supplemental Table 11). The chronostratigraphic framework for
northwest-verging Algeciras fault system (Fig. 2A), a complex series of thrust, studied basin fill was established by isotopic dating of interbedded ash beds
transpressional, and right-lateral strike-slip faults, including the alternatively and previous magnetostratigraphic studies, as described in the next section.
named Garzón, Suaza, Pitalito, and Altamira faults (Chorowicz et al., 1996; This middle to upper Miocene stratigraphic interval is exposed in two main
­Casero et al., 1997; Montes et al., 2005; Velandia et al., 2005), which bound
crystalline basement rocks of the northeast-trending Garzón Massif at the
southern tip of the Eastern Cordillera (Fig. 1). This complex fault system sepa- Sample

Garzón Mbr.
rates Jurassic granitic rocks (Algeciras monzogranite) from late Meso­protero­ Age (Ma) Transect
zoic–early Neoproterozoic high-grade metamorphic rocks (gneiss, schist, and 0
migmatites), marks the southern expression of major active transpression in Quaternary
Colombia, and has been dominated by wrench tectonics throughout the Qua-

Gigante Formation
Los Altares Member
ternary (Chorowicz et al., 1996; Velandia et al., 2005; Bustamante et al., 2010).
However, the pronounced structural relief along the >150-km-long A ­ lgeciras Pliocene
fault zone and low-angle orientation (~15° southeast dip) requires dip-slip
displacement for a significant portion of its history (e.g., Bakioglu, 2014). The 5
temporal transition from shortening to principally strike-slip deformation is 6
considered to have occurred during latest Miocene–Pliocene time, but the pre-

Late
cise timing remains unclear (Velandia et al., 2005; Egbue and Kellogg, 2010;
Egbue et al., 2014). As a result of this varied deformation history, the Garzón

Neiva Mbr.
Figure 3. Chronostratigraphic framework
Massif defines a structural high with an overall north- to northeast-plunging
for the Upper Magdalena Valley (Neiva
configuration in which ­regional-scale exposures of the prevailing Precambrian 10 Basin), showing Honda Group, Gigante
metamorphic basement and flanking Jurassic granitic rocks plunge beneath a Formation, and sample transects (modi­
5 fied after Guerrero, 1997, with updated

Miocene
northern carapace of younger geologic units (Fig. 2).
age estimates based on new U-Pb geo­

Villavieja Formation
Middle
The burial and exhumation histories of the Garzón Massif are not well chronological results from this study).
Supplemental Table 1: Sample information.
constrained. Eocene–Oligocene strata have been mapped in nonconformable
Sample Type Sample
Transect
Sample
ID
Sample
Name
Latitude
(N)
Longitude
(W)
4
contact upon Proterozoic metamorphic rocks on the eastern flank of the
­

Honda Group
Paleocurrent measurement 1 LV1-1 3.28214 75.15238
Paleocurrent measurement
Paleocurrent measurement
Paleocurrent measurement
1
2
2
LV1-2
Ccgl-1
Ccgl-2
3.29847
3.26275
3.22607
75.14776
75.18510
75.13447
Garzón Massif, suggesting that parts of the basement massif were exposed 15
Paleocurrent measurement
Paleocurrent measurement
2
2
TSS-2
TSS-1
3.27179
3.27223
75.11971
75.14490
at the surface prior to large-scale burial beneath this clastic sedimentary suc-
Paleocurrent measurement
Paleocurrent measurement
Paleocurrent measurement
3
3
4
SFSS-1
SFSS-2
PRB-4
3.22265
3.22338
3.21867
75.18974
75.17980
75.21109
cession (Rodrí­guez et al., 2003; Bakioglu, 2014; Wolaver et al., 2015). The full 3
Paleocurrent measurement
Paleocurrent measurement
5
1
PRB-2
LV1-1
3.19745
3.28214
75.20819
75.15238 massif was exhumed at some point after this phase of Eocene–Oligocene
Sandstone petrographic sample 1 S1 LV1-15 3.30757 75.14614
Sandstone petrographic sample
Sandstone petrographic sample
1
1
S2
S3
060811-11
LV1-21
3.29989
3.30082
75.14721
75.14268
sedimentation, although precise constraints on uplift timing remain elusive.
Sandstone petrographic sample 2 S4 TSS-01 3.27168 75.14435
Previously reported AFT ages span from 13.9 ± 2.3 to 9.2 ± 2.0 Ma (van der 2

La Victoria Fm.
Early
Sandstone petrographic sample 2 S5 LV3-09 3.25961 75.16226
Sandstone petrographic sample 3 S6 LVGB-13 3.22781 75.14096

Wiel, 1991), while other estimates based on structural and stratigraphic rela- 20
Sandstone petrographic sample 4 S7 PRB-43 3.20406 75.20350
Sandstone petrographic sample 4 S8 PRB-55 3.19543 75.20638

tionships range from 12.9 to 6.4 Ma (Guerrero, 1993, 1997; Butler and Schamel,
1
Supplemental Table 1. Sample information. Please
1988; van der Wiel et al., 1992; Wolaver et al., 2015). The range of interpreted
visit http://​dx​.doi​.org​/10​.1130​/GES01294​.S1 or the 1

~100 m
full-text article on www​.gsapubs​.org to view Supple- ages for the onset of Neogene basement uplift may reflect true spatial varia-
mental Table 1. tions or the diversity of approaches taken by the previous efforts. We aim to

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1238
Research Paper

areas (Fig. 2A), where the basal contact is commonly buried, or the succes- stones occur as restricted 1–5-m-thick channelized units that persist no more
sion unconformably overlies Jurassic, Cretaceous, or Eocene–Oligocene units. than 10–20 m along strike. The contact between the Villavieja Formation (­upper
The lower Miocene Honda Group is exposed in the northern basin (La Venta Honda Group) and overlying Gigante Formation (Huila Group of Guerrero,
site; Fig. 2B) and the upper Miocene Gigante Formation crops out along the 1997) is marked by a low-angle angular unconformity at La Venta, but is con-
southern basin margin (Gigante site; Fig. 2C). Due to variable exposure and formable to the south near Gigante (van der Wiel et al., 1992; Guerrero, 1997).
accessibility, all Honda and Gigante samples were taken from these localities. Here the Villavieja Formation is thinner (~250 m) and has a greater proportion
Previous efforts have established a stratigraphic framework for Neogene clas- of fluvial channel to overbank facies than the La Venta site (van der Wiel and
tic fill in the Upper Magdalena Valley, resulting in a variety of naming schemes van den Bergh, 1992).
for comparable stratigraphic sections (Van Houten and Travis, 1968; Wellman, The chronology of the middle Miocene Honda Group (La Victoria and Villa-
1970; Van Houten, 1976; Butler and Schamel, 1988; Guerrero, 1997). Here we vieja Formations) was established previously through magnetic polarity stra-
employ the stratigraphic and geochronological framework for the Miocene tigraphy of the basin-fill succession and hornblende and plagioclase 40Ar/ 39Ar
Honda Group detailed in the La Venta paleontological, magnetostratigraphic, geochronology of 8 samples of interbedded tuffs. The Honda Group spans
and 40Ar/ 39Ar geochronological studies of the northern Neiva Basin (e.g., from 13.8 to 11.6 Ma; the La Victoria–Villavieja contact is estimated at 12.5 Ma
­Guerrero, 1993, 1997; Flynn et al., 1997). For the upper Miocene Gigante Forma- (Guerrero, 1993; Flynn et al., 1997) (Fig. 3). On the basis of these age constraints,
tion, exposed primarily in the southern Neiva Basin, we use the nomenclature it appears that the calculated average sediment accumulation rates increased
of van der Wiel et al. (1992), although their Honda divisions differ slightly from from the La Victoria Formation (~400 m/m.y.) to the overlying Villavieja For-
those of Guerrero (1997). mation (~500 m/m.y.). This finding is consistent with a pronounced change in
sedimentary architecture, from the broad, amalgamated channel belts of the
La Victoria Formation to the narrow, isolated channel bodies of the Villavieja
Middle Miocene Fill of the Northern Basin: La Venta Formation (e.g., Leeder, 1977; Mackey and Bridge, 1995; Heller and Paola, 1996).
In addition, paleocurrent orientations were measured in trough cross-bed-
In the north, at the La Venta site (Fig. 2B), the middle Miocene Honda ded sandstones throughout the Honda Group using measurements of at least
Group is divided into the La Victoria and Villavieja Formations (Fig. 3). The 10 right and left limbs at each site (e.g., DeCelles et al., 1983). In lower strati-
units are separated by a 5–10-m-thick clast-supported pebble conglomerate, graphic levels, the La Victoria Formation records predominantly east-directed
the Cerbatana Conglomerate, which is exposed across the northern Neiva Ba- sediment dispersal. The intermediate section, including the Cerbatana Con-
sin (Guerrero, 1997). The older La Victoria Formation is principally composed glomerate, appears to alternate between eastward and westward paleoflow.
of bioturbated mudstones, with decimeter-scale bands of gray, purple, and These deposits are capped by an interval of north-directed paleoflow recorded
green. Many intervals contain carbonate nodules, with well-developed rhizo- in the upper Villavieja Formation (Supplemental Table 22). The paleocurrent
liths near the top of the section. These claystones and siltstones are periodi­ changes and contemporaneous shifts in sedimentation rate and channel archi­
cally interrupted by meter-scale, trough cross-bedded sandstones with ero- tecture are considered indicative of the transition from an unconfined, east-­
sive bases. Although most pinch out laterally over 20–30 m, several thicker directed meandering fluvial system to an axial braided fluvial system confined
sandstone intervals are continuous across the 2–4 km basin width and form by the development of topography in the Garzón Massif and broader Eastern
important marker beds (Guerrero, 1997). The La Victoria section is ~500 m thick Cordillera directly east of the La Venta region (e.g., Guerrero, 1997).
in the type section, thickens southward to ~1000 m near Gigante, and has been
interpreted as a meandering river system with significant soil development in
fine-grained overbank zones (Wellman, 1970; Guerrero, 1997). Upper Miocene Fill of the Southern Basin: Gigante Site
The Villavieja Formation is 580 m thick in its type section and conform-
ably caps the 5–10-m-thick Cerbatana Conglomerate. Whereas the base con- In the southern Neiva Basin (Fig. 2C), near Gigante, the type section of
tains fossiliferous green to gray mudstones and paleosols similar to the La the upper Miocene Gigante Formation (composed of the Neiva, Los Altares,
Supplemental Table 2: Trough-cross stratification paleocurrent data.

Site
Name
Bedding Bedding Left Limb Right Limb Total Trough trend Trough plunge Trough trend Trough plunge
Strike Dip (n) (n) (N) (tilted) (tilted) (tilt-corrected) (tilt-corrected)
Vic­toria Formation, the upper 350 m is characterized by brilliant red paleosols and Garzón Members) (Fig. 3) is exposed at Quebrada Guandinosita (van der
and mudstones referred to as the Polonia red beds. Soil carbonate nodules Wiel, 1991). At the base, the ~150-m-thick Neiva Member conformably caps
PRB-2 082 05 S 10 10 20 039 08 040 12
PRB-4 094 13 S 05 14 19 007 19 014 62
SFSS-1 131 05 SW 15 10 25 299 04 300 04
SFSS-2 094 06 S 10 12 22 125 01 304 04
Ccgl-1
Ccgl-2
TSS-2
105
115
084
08 S
10 S
04 S
11
15
11
10
09
10
21
24
21
069
270
117
11
01
08
070
091
118
16
04
06
are less common and not as well developed as in the red beds, but still occur the upper Honda Group (Villavieja Formation) and represents a considerable
in discrete horizons. Identification and classification of paleosols within these shift in depositional environments, with the lower 50 m predominantly com-
TSS-1 040 10 SE 10 12 22 134 12 314 03
LV1-1 043 15 SE 12 08 20 092 21 095 09
LV1-2 115 08 S 06 16 22 036 07 046 25

mudstones is based on descriptions provided by Guerrero (1993). Trough posed of pebble conglomerates and trough cross-bedded coarse-grained
2
Supplemental Table 2. Trough-cross stratification
cross-bedded sandstones occur frequently in this formation and show an up- sandstones with lenticular channel-fill features. The intermediate 50 m of the
paleocurrent data. Please visit http://​dx​.doi​.org​/10​
.1130​/GES01294​.S2 or the full-text article on www​ section increase in their abundance relative to overbank deposits. In contrast Neiva Member is a poorly exposed, mudstone-dominated interval capped by
.gsapubs​.org to view Supplemental Table 2. to the multistory channel belts of the La Victoria Formation, the Villavieja sand- 50 m of multistory clast-supported cobble-boulder conglomerates interpreted

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1239
Research Paper

to represent a north-flowing longitudinal braided fluvial system resulting from SEDIMENT PROVENANCE
a post–Honda Group increase in sediment supply (van der Wiel, 1991).
Within the Gigante Formation, the upsection stratigraphic transition to the Sandstone Petrography
Los Altares Member is marked by a decrease in grain size, with the lower 70 m
characterized by alternating mudstones and channel sandstone deposits. The To determine the provenance of the Miocene La Victoria, Villavieja, and
most significant change in this member is the appearance of pumice-rich vol- Gigante Formations, modal compositional analyses of petrographic thin
caniclastic debris flow deposits separated by thin sandstone intervals with ero- sections were performed for 15 sandstone samples from the northern (La
sive bases (Van Houten, 1976; van der Wiel, 1991). Planar-laminated vol­cani­ Venta) and southern (Gigante) segments of the Neiva Basin. Thin sections
clastic sandstones dominate upper levels of the Los Altares Member. Clasts were stained for calcium and potassium feldspars to aid in grain identifi-
from the overlying Garzón Member are predominantly of metamorphic origin, cation. For each sample, at least 300 point counts of mineral grains >62.5
suggesting a shift in sediment source from the Central Cordillera magmatic arc µm were recorded following the Gazzi-Dickinson method (Ingersoll et al.,
to basement rocks of the Garzón Massif. 1984). Primary classifications are listed in Supplemental Table 33. The counts
The depositional age of the Gigante Formation is established by previously were gathered into composite metrics of quartz, feldspar, and various lithic
reported biotite and hornblende K-Ar ages of volcaniclastic rock (Van Houten, fractions for plotting on standard ternary diagrams (Dickinson et al., 1983).
1976; van der Wiel, 1991; van der Wiel et al., 1992). These dates roughly delimit Enigmatic microcrystalline grains with iron oxide staining added some un-
the Los Altares Member to 8.0 Ma at the base and 6.4 Ma at the upper transition certainty in distinguishing components of the lithic fraction, particularly
to the Garzón Member (Fig. 3). New constraints on the maximum depositional in terms of chert versus microcrystalline volcanic or metamorphic frag-
ages are provided here (see following) by the youngest U-Pb populations in ments. We focus on Q-F-Lf (quartz–feldspar–Folk [1980] lithic fragments)
two of our detrital zircon samples: 8.6 ± 0.6 Ma from the upper Neiva Member and Qm‑F‑Lt (monocrystalline quartz–feldspar–total lithic fragments) ternary
and 7.7 ± 0.7 Ma from Los Altares volcaniclastic debris flows. These U-Pb age diagrams (Fig. 4) rather than Qt-F-L (total quartz–feldspar–lithic fragments)
constraints are consistent with previous geochronological results, and verify a distributions in order to minimize the influence of chert on the overall quartz
late Miocene age for the Gigante Formation. fraction (e.g., Folk, 1980).

Qm Q
Quartzarenite

Craton
A Interior
B

Transitional Quartzose
Continental Recycled
Subarkose Sublitharenite
Mixed
Gigante Fm.
S15 S15 S2
Villavieja Fm.
S1 S3
Transitional
Recycled
La Victoria Fm.
S4
S2 S6 S11 S10
Dissected Arc S9 S7 S8
S3
Supplemental Table 3: Sandstone petrographic data. S1 S5
Locality Neiva Neiva Neiva Neiva Neiva Neiva Neiva Neiva Neiva Neiva Gigante Gigante Gigante Gigante Gigante
Basement S4
Sample ID
Sample Name
Grain S1 S2 S3 S4 S5 S6 S7 S8 S9 S10 S11 S12 S13 S14 S15
Code LV1-15 060811-11 LV1-21 TSS-01 LV3-09 LVGB-13 PRB-43 PRB-55 PRB-07 PRB-05 070811-03 N01-04 080811-12 UGM11-05 080811-09 S9 S11
Quartz: monocrystalline (Qms)
Quartz: monocrystalline, undulose (Qmu)
Qms
Qmu
74
14
72
24
65
33
54
20
55
12
56
15
62
9
66
7
75
6
65
7
24
47
25
19
35
26
39
21
93
42 Uplift S5 S6 S10
Quartz: polycrystalline, straight (Qps) Qps 16 12 13 8 6 18 8 22 1 21 3 4 2 3 3

S13 S8 S14 S13


S14
Quartz: polycrystalline, undulose (Qpu) Qpu 21 19 18 18 8 19 10 12 16 14 19 4 - 1 1

S7
Feldspar: potassium feldspar (K) K 6 4 7 7 5 9 4 3 17 18 - - 13 17 -
Feldspar: plagioclase (P) P 38 28 27 48 72 48 32 28 33 18 32 66 102 118 75
Feldspar: altered (Falt) Falt 5 24 17 1 29 14 7 - 7 - 11 16 1 4 -
Chert (C)
Siltstone (S)
C
S
22
-
3
-
2
-
3
-
12
-
5
2
44
-
67
2
26
-
72
4
6
-
3
-
1
-
-
-
-
- S12
S12
Lithic: mudstone/shale (Lsh) Lsh 5 5 3 3 6 19 - - - - - - - - -
Lithic: carbonate (Lc) Lc 4 - - - 1 - - 2 - 1 7 - - - -

Transitional Arc Lithic Feldspathic


Lithic: mafic volcanic (Lvm) Lvm 16 16 40 38 19 16 18 19 37 9 47 83 74 46 33
Lithic: felsic volcanic (Lvf) Lvf 28 - 12 32 6 53 22 31 29 34 - 14 8 2 5
Lithic: lathwork volcanic (Lvl)
Lithic: pyroclastic/glassy (Lvp)
Lvl
Lvp
1
14
4
21
1
33
-
37
-
15
-
25
-
52
-
42
-
31
-
34
28
14
-
36
2
7
-
22
-
24
Arkose Lithic Arkose Litharenite
Recylced Litharenite
Lithic: phyllite (Lph) Lph 4 9 12 - 14 4 3 1 12 - 11 - - - -
Lithic: schist (Lsm) Lsm 2 16 8 - 8 - - - 1 - 8 4 1 - 1
Lithic: gneissic (Lg) Lg 2 2 - - 1 - - - - - 2 - - - -
Lithic: metamorphic microcrystalline (Lmm) Lmm 1 13 17 - 15 - 25 - 17 - 19 - - 7 -
Hornblende (Mhbl) Mhbl 9 10 - 1 11 1 - - - - 14 6 23 6 23
Biotite (Mbio) Mbio 10 5 3 15 4 - - - 1 1 - 1 3 - -
Muscovite (Mmus) Mmus 7 9 1 2 5 1 - - - - 13 19 6 16 -
Hematite (Mhem) Mhem 6 1 - 4 - - - 11 - 4 - - - - -
Quartz: Hydrated Oxidized (Qho)
Coal (Co)
Qho
Co
-
1
5
-
-
-
9
6
-
-
7
-
4
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
Undissected Arc
3
Supplemental Table 3. Sandstone petrographic data. F Lt F Lf
Please visit http://​dx​.doi​.org​/10​.1130​/GES01294​.S3
or the full-text article on www​.gsapubs​.org to view Figure 4. Ternary diagrams showing sandstone compositional results for 15 samples from the La Victoria, Villavieja, and Gigante Formations. (A) Qm-F-Lt (monocrys­
Supplemental Table 3. talline quartz–feldspar–total lithic fragments). (B) Q-F-Lf (quartz–feldspar–Folk [1980] lithic fragments).

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1240
Research Paper

Samples throughout all three formations exhibit heterogeneous composi- Sediment Source Areas
tions relatively evenly distributed among quartz, feldspar, and lithic fractions
(Figs. 4 and 5; Table 1). Most samples are categorized as lithic arkose or feld- U-Pb age populations in our sample set (Fig. 5) are compatible with po-
spathic litharenite (Fig. 4B), with the oldest unit, the La Victoria Formation, tential sediment source regions surrounding the Neiva Basin. These sources
slightly more quartzose than the overlying Villavieja and Gigante Formations. include: (1) late Mesoproterozoic–early Neoproterozoic metamorphic base-
An exception is uppermost Gigante sample S15, which exhibits the largest ment of the Garzón Massif; (2) Jurassic granitoids; (3) Cretaceous sedimen-
quartz fraction at 50.2%. Total feldspar is relatively uniform throughout the tary rocks containing recycled Proterozoic cratonic zircons; and (4) Cenozoic
Honda Group (La Victoria and Villavieja Formations), with the exception of volcanic rocks from the Central Cordillera magmatic arc. In the following we
sample S5, which is substantially richer in plagioclase and nonvolcanic lithic describe the age, distribution, and distinguishing characteristics of these dif-
fragments than the rest of the Honda Group. This sample, from immediately ferent sources.
below the Cerbatana Conglomerate (at the La Victoria–Villavieja contact), may
highlight a provenance change linked to the transient shift in depositional con- Garzón Metamorphic Basement
ditions responsible for gravel input. In upper stratigraphic levels, less mature
samples from the Gigante Formation (S11–S15) tend to have much higher The Garzón Massif is composed of several distinct metamorphic units
proportions of volcanic lithic fragments and feldspars than samples from the with both igneous and sedimentary protoliths. Grains derived from prin-
under­lying Honda Group (Figs. 4 and 5; Table 1), likely due to a greater proxim- cipally gneisses and schists of this basement massif display a range of Pre-
ity of magmatic arc rocks in the Central Cordillera. cambrian ages, including principally late Mesoproterozoic–early Neoprotero-
zoic crystallization ages of protolith and metamorphic zircons concentrated
Detrital Zircon U-Pb Geochronology at 1200–900 Ma, with limited older zircon grains and no grains older than
1500 Ma (­Restrepo-Pace et al., 1997; Cordani et al., 2005; Cardona et al., 2010;
Detrital zircon U-Pb geochronological analyses were performed on 11 Ibañez-Mejía et al., 2011). Because comparable metamorphic basement is not
sandstone samples from the Neiva Basin; 7 La Venta and four Gigante sam- exposed farther west, identification of a sizeable 1200–900 Ma population in
ples were crushed and zircon grains separated using standard techniques, in- the detrital record would represent a robust indication of an eastern (Garzón
cluding water, heavy liquid, and magnetic separation. A random selection of Massif) sediment source for the Neiva Basin.
100–130 zircons in each sample was measured via laser ablation–inductively
coupled plasma–mass spectrometry (LA-ICP-MS) at the University of Arizona Jurassic Plutonic Rocks
Supplemental Table 4: U-Pb geochronological data.

1
207Pb/206P 207Pb/235 Corrected 206Pb*/238U* 207Pb*/235U 1 error 206Pb*/207Pb* 1 error Best Age 1 error
Alysis U (ppm) U/Th Error (%) Error (%) 206Pb/238U Error (%) error
b U Error (%) Age Age (Ma) Age (Ma) (Ma) (Ma)
(Ma)

Z1: 010811-01
010811-01-038
010811-01-086
010811-01-003
200
132
233
2.8
2.8
3.5
32.612
1.491
9.944
91.47
822.49
252.88
0.0092
0.1995
0.03155
92.42
822.74
253.03
0.00218
0.00216
0.00228
13.21
20.03
8.78
0.14
0.02
0.03
14.0
14.0
15.0
2.0
3.0
1.0
9.3
184.7
31.5
8.6

78.7
-1092.0

1635.0
1874.0

660.0
14.0
14.0
15.0
2.0
3.0
1.0
LaserChron Center, following standard analytical procedures (Gehrels et al.,
010811-01-058 196 2.2 7.285 153.61 0.04299 155.38 0.00227 23.39 0.15 15.0 3.0 42.7 65.1 2193.0 188.0 15.0 3.0

2008; Gehrels, 2012, 2014). A known standard of Sri Lanka zircon was mea- Considerable volumes of intrusive igneous rocks in the northern Andes
010811-01-057 95 3.3 -4.779 338.35 -0.07601 339.01 0.00263 21.22 0.06 17.0 4.0 -80.3 -290.9 17.0 4.0
010811-01-072 151 1.8 5.636 224.63 0.06528 225.23 0.00267 16.54 0.07 17.0 3.0 64.2 141.0 2629.0 199.0 17.0 3.0
010811-01-037 121 2.3 -4.197 603.51 -0.12404 603.67 0.00378 14.21 0.02 24.0 3.0 -134.5 -1293.4 24.0 3.0
010811-01-014 264 0.7 19.464 19.7 0.03895 20.63 0.0055 6.12 0.3 35.0 2.0 38.8 7.9 258.0 457.0 35.0 2.0
010811-01-033 143 1.3 16.057 78.07 0.04658 78.43 0.00542 7.48 0.1 35.0 3.0 46.2 35.5 684.0 2048.0 35.0 3.0
010811-01-001 164 1.5 29.614 82.29 0.02577 82.47 0.00554 5.40 0.07 36.0 2.0 25.8 21.0 -809.0 2868.0 36.0 2.0

sured between every five unknown grains to correct for transient variations were emplaced during several phases spanning the Jurassic, with an age
010811-01-046 315 1.3 21.184 22.03 0.03682 22.32 0.00566 3.58 0.16 36.0 1.0 36.7 8.0 60.0 530.0 36.0 1.0
010811-01-041 140 0.9 33.236 52.39 0.02423 52.8 0.00584 6.56 0.12 38.0 2.0 24.3 12.7 -1150.0 1704.0 38.0 2.0
010811-01-055 491 0.8 18.906 10 0.04311 11.5 0.00591 5.68 0.49 38.0 2.0 42.9 4.8 324.0 228.0 38.0 2.0
010811-01-118 194 1.1 19.943 39.32 0.04105 39.69 0.00594 5.39 0.14 38.0 2.0 40.9 15.9 202.0 946.0 38.0 2.0
010811-01-090 133 1 23.386 44.61 0.03554 45.31 0.00603 7.89 0.17 39.0 3.0 35.5 15.8 -182.0 1165.0 39.0 3.0

in interelement and intraelement fractionation over the course of the run. The range of 200–150 Ma in southern Colombia at 1°–4°N (Aspden et al., 1987;
010811-01-023 144 0.9 21.445 60.73 0.03984 61.22 0.0062 7.73 0.13 40.0 3.0 39.7 23.8 30.0 1600.0 40.0 3.0
010811-01-035 115 0.9 19.325 45 0.04421 46.04 0.0062 9.74 0.21 40.0 4.0 43.9 19.8 274.0 1082.0 40.0 4.0
010811-01-059 172 0.9 23.397 32.91 0.03638 33.45 0.00617 5.96 0.18 40.0 2.0 36.3 11.9 -183.0 841.0 40.0 2.0
010811-01-044 135 1.1 23.551 115.88 0.03709 116.26 0.00634 9.39 0.08 41.0 4.0 37.0 42.2 -199.0 1355.0 41.0 4.0
010811-01-069 110 1.6 18.759 36.64 0.04645 37.54 0.00632 8.18 0.22 41.0 3.0 46.1 16.9 342.0 856.0 41.0 3.0
010811-01-087 158 1.3 15.358 31.66 0.0567 32.45 0.00632 7.14 0.22 41.0 3.0 56.0 17.7 778.0 682.0 41.0 3.0

Pb/ 207Pb age is reported for all grains with a 206Pb/ 238U age older than 900 Ma; Bustamante et al., 2010; Villagómez et al., 2011). These granitoid plutons
010811-01-114 166 1 28.663 41.37 0.03034 42.1 0.00631 7.80 0.19 41.0 3.0 30.3 12.6 -718.0 1195.0 41.0 3.0
010811-01-018 134 1.1 13.72 91.92 0.06539 92.23 0.00651 7.59 0.08 42.0 3.0 64.3 57.5 1011.0 243.0 42.0 3.0
206

4
Supplemental Table 4. U-Pb geochronological data. due to the relatively low abundance of 207Pb in grains younger than 900 Ma, the are preserved in both the Central and Eastern Cordilleras on the western
Please visit http://​dx​.doi​.org​/10​.1130​/GES01294​.S4 206
Pb/ 238U age is reported for these grains. Individual measurements with er- and eastern flanks of the Neiva Basin (Gómez Tapias et al., 2007). Here the
or the full-text article on www​.gsapubs​.org to view
rors >10% in 206Pb/ 238U and 206Pb/ 207Pb ratios were discarded. For grains where Jurassic granitic plutons are in intrusive and fault contact with Paleozoic
Supplemental Table 4.
the 206Pb/ 207Pb age was reported, measurements exhibiting >30% discordance strata and older metamorphic basement (Figs. 2A, 2C), including princi-
or >5% reverse discordance were also discarded (Gehrels et al., 2008). U-Pb pally Mesoproterozoic rocks of the Eastern Cordillera (Garzón Massif) and
0

results are presented in Supplemental Table 44 and plotted as age histograms Neoproterozoic–lower Paleozoic rocks of the Central Cordillera (Cajamarca
Z9 Z10
5 Z8
and probability functions (Fig. 5) to reveal stratigraphic variations in U-Pb age complex).
Z4 Z5 distributions. For eight samples with populations of young Cenozoic zircons
Age (Ma)

Z3
10 Z1 Z2 n = 23

n=2
n = 15
8.6 ± 0.6 Ma
8.6 ± 3.4 Ma
7.7 ± 0.7 Ma
(samples Z1, Z2, Z3, Z4, Z5, Z8, Z9, and Z10), the calculated mean of tightly Cretaceous Sedimentary Rocks Bearing Recycled Cratonic Zircons
15
n = 34 n=5
n=5
13.2 ± 1.0 Ma
clustered grain ages (Table 2; Fig. 5; Supplemental Fig. 15) provides a robust
constraint on the maximum depositional age of the host strata and may ap- Cretaceous sedimentary rocks in Colombia display substantial variation in
n = 13 13.2 ± 1.3 Ma
n=4
13.75 ± 0.4 Ma 13.75 ± 0.9 Ma
14.4 ± 1.9 Ma
20
proximate the true depositional age in cases with considerable contributions their detrital zircon U-Pb age spectra. Lower Cretaceous stratigraphic units are
5
Supplemental Figure. Plot of the youngest U-Pb age from syndepositional volcanic sources. Seven of the eight samples (excluding often heterogeneous, containing significant zircon populations of Cambrian–
populations for eight samples of the Neiva Basin.
sample Z5) yield mean ages that closely match the depositional ages defined Ordovician age (500–400 Ma), late Mesoproterozoic–earliest Neo­protero­zoic
Please visit http://​dx​.doi​.org​/10​.1130​/GES01294​.S5
or the full-text article on www​.gsapubs​.org to view by previous 40Ar/ 39Ar and magnetostratigraphic results (van der Wiel, 1991; van (Sunsás-Grenville) age (1250–950 Ma), and older Proterozoic grains rang-
the Supplemental Figure. der Wiel et al., 1992; Guerrero, 1993, 1997; Flynn et al., 1997). ing from 2060 to 1300 Ma (Horton et al., 2010, 2015a; Saylor et al., 2013).

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1241
1242
Sandstone

Garzón Mbr.
Composition Detrital Zircon U-Pb Ages
S15
Z11
N = 113 Z11 30
S13
20
Clast-
Supported 10
Conglomerate
Pumice-rich Z10: 7.7 ± 0.7 Ma (n = 23)
Matrix-
S14 N = 107 Z10 30

Los Altares Member


Supported
20

Gigante Formation

Accessory Minerals
Conglomerate
10
Sandstone

Volcanic Lithics
Cross-
Stratified
Z9: 8.6 ± 0.6 Ma (n = 15) N = 105 Z9 30
Sandstone 20
Z10
Siltstone 10

Feldspars
Claystone or
Z8: 8.6 ± 3.4 Ma (n = 2) N = 74
Paleosol Z8 30
20

Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes
Quartz

10

Relative Probability

of Grains
Z9

Number
Neiva Member

S12

N = 112 Z7 30
S11 Z8 20
S10 Z7 10
Paleocurrent
Non-Volcanic Lithics

Direction S9
N = 103 Z6 80
N = 20 Z6 60
40
Villavieja Formation

S8 20
Volcanic Lithics

N = 19 Z5: 13.2 ± 1.0 Ma (n = 5) N = 95


S7 Z5 30
20
10
Z5
Feldspars

N = 25,22 0
Z4: 13.2 ± 1.3 Ma (n = 5) N = 52 Z4
Honda Group

30
20
Quartz

Cerbatana 10
Conglomerate S6
Z3: 13.75 ± 0.9 Ma (n = 13) N = 74
Z4
Z3 30
N = 24,21 S5 20
10
La Victoria Formation

N = 21,22
S4 Z3
Z2: 13.75 ± 0.4 Ma (n = 34) N = 82 Z2 40
30
20
N = 20
Z2 10
100 m

Z1: 14.4 ± 1.9 Ma (n = 4) N = 80


S3
S2 Z1 30
N = 22 20
S1 Z1 10
0 25 50 75 100% 0
Medium sand
Gravel
Coarse sand

Fine sand
Silt
Clay

Relative Abundance 0 250 500 750 1000 1250 1500 1750 2000
Age (Ma)

Figure 5. Plot of sedimentary provenance results relative to a lithologic log (after Guerrero, 1997) for the Honda Group and Gigante Formation of the Neiva Basin. Major columns
GEOSPHERE  |  Volume 12  |  Number 4

show depositional ages from past 40Ar/ 39Ar and magnetostratigraphic results, sandstone petrographic results in the form of relative abundances of major constituent grains,
and detrital zircon U-Pb ages depicted as age histograms and probability density functions.
Research Paper
Research Paper

TABLE 1. SANDSTONE POINT COUNT RESULTS FOR PETROGRAPHIC THIN SECTIONS


Sample Sample Sample Accessory
transect ID name Formation Total Q F Lv Ls + Lm minerals Qm F Lt Q F Lf
1 S1 LV1-15 La Victoria 306 125 49 59 40 33 88 49 136 125 49 98
1 S2 060811-11 La Victoria 302 127 56 41 48 30 96 56 120 127 56 76
1 S3 LV1-21 La Victoria 312 129 51 86 42 4 98 51 159 129 51 111
2 S4 TSS-01 La Victoria 306 100 56 107 6 37 74 56 139 100 56 113
2 S5 LV3-09 La Victoria 304 81 106 40 57 20 67 106 111 81 106 82
3 S6 LVGB-13 Villavieja 312 108 71 94 30 9 71 71 161 108 71 124
4 S7 PRB-43 Villavieja 300 89 43 92 72 4 71 43 182 89 43 139
4 S8 PRB-55 Villavieja 313 107 31 92 72 11 73 31 198 107 31 164
5 S9 PRB-07 Villavieja 309 98 57 97 56 1 81 57 170 98 57 136
6 S10 PRB-05 Villavieja 302 107 36 77 77 5 72 36 189 107 36 154
6 S11 070811-03 Gigante 305 93 43 89 53 27 71 43 164 93 43 123
6 S12 N01-04 Gigante 300 52 82 133 7 26 44 82 148 52 82 140
6 S13 080811-12 Gigante 304 63 116 91 2 32 61 116 95 63 116 93
6 S14 UGM11-05 Gigante 302 64 139 70 7 22 60 139 81 64 139 70
6 S15 080811-09 Gigante 300 139 75 62 1 23 135 75 67 139 75 63
Note: Q—quartz total (Qms + Qmu + Qps + Qpu: m is monocrystalline, p is polycrystalline, s is straight extinction, u is undulose extinction); F—feldspar total (K + P + Falt:
K is potassium, P is plagioclase, Falt is altered feldspar); Lv—lithic volcanic (Lvm + Lvf + Lvl + Lvp: m is microlitic, f is felsic, l is lathwork, p is pyroclastic/glassy); Ls—lithic
sedimentary (C + S + Lsh + Lc: C is chert, S is siltstone, Lsh is mudstone/shale, Lc is carbonate); Lm—lithic metamorphic (Lph + Lsm +Lg + Lmm: Lph is phyllite, Lsm is schist,
Lg is gneiss, Lmm is microcrystalline metamorphic); Qm—quartz monocrystalline (Qms + Qmu: s is straight extinction, u is undulose extinction); Lt—lithic total (Lv + Ls + Lm);
Lf—Folk (1980) lithic total (Lv + Ls + Lm + C). Complete sandstone petrographic data are reported in Supplemental Table 3.

The proportion of younger grains diminishes upsection, such that uppermost U-Pb Results
Cretaceous units are commonly dominated by 2060–1300 Ma ages, likely due
to bulk contributions from Paleoproterozoic–middle Mesoproterozoic cratonic At the base of the Neiva Basin succession, samples Z1–Z4 from the middle
sources and gradual burial of Andean sources (Horton et al., 2010, 2015a). Cre- Miocene La Victoria Formation contain predominantly Cenozoic and subor-
taceous strata were deposited unconformably over Jurassic sedimentary and dinate Jurassic age zircon grains (Fig. 5), suggesting a western source in the
plutonic rocks (Ramon and Rosero, 2006; Mora et al., 2010b), and are wide- ­Central Cordillera during initial filling of the Neiva Basin. The abundance of
spread on both flanks of the Neiva Basin. Miocene zircon grains allows for calculation of the youngest U-Pb age popula-
tions for successive samples Z1–Z4, which yield weighted mean ages of 14.4 ±

Cenozoic Volcanic Rocks TABLE 2. MEAN AGES OF YOUNGEST DETRITAL


ZIRCON U-Pb AGE POPULATIONS
The Central Cordillera has been part of a magmatic arc complex in Weighted mean of
­ olombia since the Late Cretaceous (Aspden et al., 1987; Bayona et al.,
C Sample youngest grains 1σ error ± Number of
2012). Compositional analyses of Neogene basin fill in the Upper Magda- ID (Ma) (Ma) MSWD grains
lena Valley have been tied to coeval magmatic sources in the adjacent Cen- Z1 14.4 1.9 0.046 4
tral Cordillera (Van Houten and Travis, 1968; Van Houten, 1976; Guerrero, Z2 13.75 0.42 0.057 34
1993, 1997). Although minor volcanic provinces exist elsewhere, there are Z3 13.75 0.90 0.088 13
no mapped Cenozoic volcanic units at the southern end of the Eastern Cor­ Z4 13.2 1.3 0.059 5
di­llera (Fig. 2; Gómez Tapias et al., 2007). Therefore, significant populations Z5 13.2 1.0 0.045 5
Z8 8.6 3.4 0.000 2
of Cenozoic-age zircon grains can be considered clear signatures of a Cen-
Z9 8.58 0.58 0.210 15
tral Cordillera source along the western flank of the Neiva Basin. Within the
Z10 7.67 0.66 0.047 23
basin, young zircon grains of middle to late Miocene age may reflect direct
erosion from syndepositional volcanic rocks or direct input from volcanic Note: MSWD—mean square of weighted deviates. Complete U-Pb geochronological
data are reported in Supplemental Table 4.
ash-fall deposition.

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1243
Research Paper

1.9 Ma (Z1, n = 4), 13.75 ± 0.4 Ma (Z2, n = 34), 13.75 ± 0.9 Ma (Z3, n = 13), and zircons (Fig. 5). This age distribution is nearly identical to reported ages from
13.2 ± 1.3 Ma (Z4, n = 5) (Table 2; Fig. 5; Supplemental Fig. 1). These ages agree Garzón crystalline basement to the east (Restrepo-Pace et al., 1997; Cordani
well with existing magnetostratigraphic results constraining accumulation of et al., 2005; Cardona et al., 2010; Ibañez-Mejía et al., 2011), suggesting a shift to
the La Victoria Formation between 13.8 and 12.5 Ma (Guerrero, 1993; Flynn a major sediment source region in the Garzón Massif east of the Neiva Basin.
et al., 1997). Therefore, the youngest zircon populations not only constrain the The provenance variations reflected in samples Z1–Z12 show a middle to
maximum depositional ages, but likely represent accurate estimates of the true late Miocene shift from a western to eastern sediment source, a prolonged
depositional ages of stratigraphic units in close proximity to the Andean mag- transition interpreted to reflect the increasing influence of the Garzón Massif
matic arc (e.g., Perez and Horton, 2014; Horton et al., 2015b). Sample Z4 from the (and broader Eastern Cordillera) relative to the Central Cordillera. Although
Cerbatana Conglomerate at the La Victoria–Villavieja transition contains clasts Garzón-derived sediment may have appeared by ca. 12.5 Ma (sample Z4),
of principally igneous rocks with subordinate quartzite, slate or schist, and chert the first clear signal appeared ca. 11.6 Ma (samples Z6 and Z7), and a fully
(Van Houten and Travis, 1968; Guerrero, 1993, 1997). The chert fragments are dominant signal was not realized until ca. 6.5 Ma (sample Z11) (Fig. 5). This
likely derived from Cretaceous sedimentary rocks, which are preserved on protracted transition included complex depositional alternations involving
both the western and eastern sides of the basin. Therefore, although the U-Pb (1) an integrated contribution of sediments eroded from both the eastern and
age distributions show a sustained influx from the Central Cordillera along the western basin margins in an axial (longitudinal) fluvial system, yielding a his-
western flank of the Neiva Basin, we cannot rule out possible contributions tory of highly diverse age distributions, and/or (2) competing sediment influx
from recycled Cretaceous cover strata from the Garzón Massif to the east. from alluvial-fan systems originating from opposing basin margins, producing
Within intermediate stratigraphic levels, the Villavieja Formation records a record with abrupt appearances and disappearances of particular age pop-
significant variations in U-Pb age distributions reflective of exhumation of the ulations. Given the observed stratigraphic appearance and disappearance of
Garzón Massif within the broader Eastern Cordillera east of the Neiva Basin. Cenozoic, Jurassic, and Precambrian age signatures emblematic of western
Sample Z5 shows a continued dominance of Cenozoic age distributions, in- versus eastern sources (Fig. 5), we favor the second option for most of the
cluding a youngest population of 13.2 ± 1.0 Ma (Z5, n = 5), but an increased middle to late Miocene provenance record. We propose that the Neiva Basin
proportion of Jurassic ages (Table 2; Fig. 5). Upsection, sample Z6 from the was filled from 12.5 to 6.4 Ma by competing distributary systems sourced prin-
upper Villavieja Formation marks the abrupt disappearance of Cenozoic grains cipally from the west (Central Cordillera), but with minor mixing of signals and
(Fig. 5) and wholesale replacement by a unimodal Jurassic signature, with 79 progressively greater contributions from the emerging eastern topographic
of 87 total ages clustered between 200 and 170 Ma. For sample Z7, at the top of barrier (Garzón Massif and broader Eastern Cordillera).
the Villavieja Formation, the Jurassic population abruptly disappears and is re-
placed by a heterogeneous, multimodal mix of Precambrian ages, which span AFT THERMOCHRONOMETRY
principally from 1600 to 1000 Ma (Fig. 5). Although the 1200–900 Ma grains are
emblematic of the Garzón Massif, the older grains indicate input from other AFT thermochronometry was employed to evaluate late Cenozoic exhuma-
sources, most likely recycled from Cretaceous–Paleogene sedimentary units tion of the central segment of the Garzón Massif structural domain. The abun-
(e.g., Horton et al., 2010, 2015a). Overall, the conspicuous absence of Ceno- dance, length, and distribution of tracks resulting from the spontaneous fission
zoic zircon populations in samples Z6 and Z7 suggests no major western sedi- of 238U isotopes provide constraints on the thermal history of a sample below
ment source for the upper Villavieja Formation. This significant departure from the 100–130 °C closure temperature for apatite (e.g., Donelick et al., 2005; Ehlers,
under­lying samples requires an extensive provenance shift ca. 12 Ma, con- 2005). We collected 10 samples of crystalline basement rocks (granite, gneiss, and
sistent with early detrital contributions from the east, likely from the Garzón schist) along a transect across the Garzón Massif, crossing a relatively straight,
Massif and/or Cretaceous–Paleogene cover strata. northeast-striking strand of the Algeciras fault system along the western margin
In the upper levels of the Neiva Basin, further provenance variations are of the massif (Fig. 2). The six lower-elevation samples (Fig. 2C), including three
expressed in detrital zircon U-Pb results for the upper Miocene Gigante For- samples from Jurassic granitic rocks west of the Algeciras fault (samples GM01–
mation. Samples Z8, Z9, and Z10 from the lower to middle Gigante Formation GM03) and three samples from Proterozoic gneisses and schists on the eastern
(Neiva and Los Altares Members) are all dominated by Cenozoic and Jurassic fault block (samples GM04–GM06), were found to have sufficient apatite material
zircons (Fig. 5), similar to lower basin fill (Z1–Z5), possibly implying a return to for AFT analyses, and the results are reported in the following.
a western source. The youngest U-Pb age populations for successive samples
Z8–Z10 yield weighted mean ages of 8.6 ± 3.4 Ma (Z8, n = 2), 8.6 ± 0.6 Ma (Z9, Methods
n = 15), and 7.7 ± 0.7 Ma (Z10, n = 23) (Table 2; Fig. 5; Supplemental Fig. 1). In
contrast to these three samples, sample Z11 from the uppermost Gigante For- Rock samples were crushed and milled, then apatite grains were separated
mation (Garzón Member) contains a large population of zircon grains spanning using a water table, heavy liquids, and magnetic separation. AFT analyses
from 1500 to 950 Ma, a minor population of Jurassic zircons, but no Cenozoic were performed by Apatite to Zircon Inc., following the procedures outlined by

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1244
Research Paper
0.40
20 GM01 AFT Age = 3.59 Ma n = 11
40 (+3.11 / –1.23 Ma) Track Length 0.35
Donelick et al. (2005). Apatite grains were mounted, polished, and etched with

Temperature (°C)
60 μ = 13.36 0.30
5.5N HNO3 to expose fission tracks. Spontaneous tracks were counted to deter- 80
σ = 1.26
0.25
mine the AFT age, and uranium and other elemental concentrations were then 100 0.20
determined via LA-ICP-MS. The samples were irradiated with a 252Cf source, 120
0.15
140
and re-etched to expose more spontaneous fission tracks for measurement of 0.10
160
track lengths. In order to distinguish multiple apatite populations having differ- 180 0.05
ent kinetic parameters, track lengths, AFT ages, and U-Pb ages calculated for 0.00

each grain were plotted against the multiple kinetic parameter rmr0 (Ketcham 20 GM02 AFT Age = 3.91 Ma n=7
40 (+3.84 / –1.47 Ma) Track Length 0.30

Temperature (°C)
et al., 1999; Carlson et al., 1999) and examined for potential dependencies via 60 μ = 14.13
σ = 1.22 0.25
the c2 test. Thermal histories for each sample were inverse-modeled using 80
0.20
HeFTy software (version 1.8.3; June 2014) (Supplemental Files for HeFTy6), with 100
120 0.15
track lengths projected onto the c-axis to account for track angle (Ketcham,
140 0.10
2005). Each modeling run was performed with surface temperature fixed at 160
0.05
20 °C and the condition that the massif reached >140 °C burial temperatures 180
0.00
at some point between 40 and 10 Ma, a loose constraint placed by geologic
20 GM03 AFT Age = 4.60 Ma n = 69
0.35
evidence that the massif was exposed at the surface during the Eocene, and 40 (+2.31 / –1.01 Ma) Track Length 0.30

Temperature (°C)
subsequently buried by mid-Cenozoic strata prior to its final exhumation 60 μ = 14.71
σ = 1.04 0.25
(­Rodríguez et al., 2003; Wolaver et al., 2015). 80
100 0.20
120 0.15
Results 140 0.10
160
180 0.05
AFT results for the six samples show uniform, elevation-correlated values, 0.00
with pooled AFT ages ranging from 5.8 to 3.6 Ma (Fig. 6; Table 3). All samples 20 GM04 AFT Age = 5.76 Ma n = 37 0.40
were completely reset, displaying unimodal track length distributions, consis- 40 (+3.23 / –1.39 Ma) Track Length

Temperature (°C)
μ = 15.08 0.35
tent with cooling-only thermal histories. Whereas the three samples west of 60
σ = 0.78 0.30
80
the Algeciras fault (samples GM01–GM03) yield pooled AFT ages of 4.6–3.6 Ma 100 0.25
and a composite average track length of 13.2 µm, those from the uplifted east- 120 0.20
140 0.15
ern fault block (samples GM04–GM06) exhibit 5.8–5.0 Ma pooled AFT ages and
160 0.10
a 14.5 µm composite average track length (Table 3). Most samples are domi-
180 0.05
nated by a single kinetic population; in all but one case (sample GM05), fewer 0.00
20 GM01 AFT Age = 3.59 Ma than three anomalous grains were removed to obtain a single population for 20 GM05 AFT Age = 5.37 Ma 0.60
n = 5 (pop.1) 0.50
40 (+3.11 / -1.23 Ma)
thermal modeling. 40 pop. 1: (n=5) 12.2 +15.5 / –6.8 Ma 0.40

Temperature (°C)
Track Length
Temperature (°C)

60 60 pop. 2: (n=27) 5.03 + 1.64 / –1.24 Ma μ = 15.47 0.30


80 Our HeFTy modeling reveals that this uniform resetting requires that all 80 σ = 0.87 0.20
100 of these basement samples were originally located below the zone of partial 0.10
100 0.00
120 0.50
140 annealing (Ketcham et al., 1999; Ehlers, 2005), indicating principally post-Mio- 120 n = 27 (pop. 2)
Track Length 0.40
160 140
μ = 15.52 0.30
180 160 σ = 0.8 0.20
180 0.10
20 GM02 AFT Age = 3.91 Ma 0.00
40 (+3.84 / -1.47 Ma) Figure 6. Apatite fission track (AFT) thermal modeling (HeFTy) results for the Garzón Massif GM06 AFT Age = 5.04 Ma 0.60
Temperature (°C)

20 n = 40
60 showing thermal histories (left) and track length distributions (right). Modeled time-tempera­ (+3.72 / –1.53 Ma) 0.55
40 Track Length

Temperature (°C)
80 ture solutions include good fit paths (magenta) and acceptable fit paths (green), where modeled 0.50
60 μ = 15.29
100 0.45
ages and lengths fit the data more closely than statistical expectation and do not fail a 95% σ = 0.81
120 80 0.40
confidence test, respectively (Ketcham, 2005). Reported track lengths use c-axis projection, and 0.35
140 100 0.30
are therefore slightly greater than the raw measured track lengths (Table 3). Observed AFT ages
160 120 0.25
180
are indicated on each plot by a dashed black vertical line; 95% confidence intervals are indicated 0.20
140
by a shaded gray rectangle. Typical ranges of closure temperatures for AFTs are indicated by the 0.15
160
horizontal dashed red lines Sample GM05 fails the χ2 test, and thus could not be modeled as 0.10
180 0.05
6
Supplemental Files for HeFTy. Please visit http://​dx​ a single population. Ages and track lengths were separated into two populations with distinct
0.00
.doi​.org​/10​.1130​/GES01294​.S6 or the full-text article annealing kinetics based on the parameter rmr0 , resulting in two histograms; these two popula­ 20 18 16 14 12 10 8 6 4 2 0 0 4 8 12 16 20
on www​.gsapubs​.org to view the Supplemental Files. tions were modeled jointly. Age (Ma) c-axis Projected
Track Length (µm)

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1245
Research Paper

TABLE 3. APATITE FISSION TRACK RESULTS FOR BASEMENT SAMPLES FROM THE GARZÓN MASSIF
95% Track length
Average Pooled confidence Average standard
Sample Latitude Longitude Elevation Number P (χ2) Dpar Dper U conc. age interval Number Number Average track length deviation
name (N) (W) (m) of grains Ns ρs (%) (µm) (µm) (ppm) (Ma) (Ma) of grains of tracks rmr0 (µm) (µm)
GM01 2.57525 75.37080 676.9 27 27 0.000065 55.8 1.72 0.33 9.08 3.59 (2.36, 5.47) 10 11 0.84 11.87 2.32
GM02 2.54568 75.35829 789.4 21 18 0.000038 77.4 1.61 0.31 6.7 3.91 (2.44, 6.28) 6 7 0.85 12.64 1.71
GM03 2.53648 75.34039 879.5 40 66 0.000120 89.2 1.55 0.31 9.11 4.60 (3.59, 5.90) 44 72 0.85 13.48 1.71
GM04 2.56324 75.27018 1179.3 40 53 0.000076 2.15 1.55 0.34 9.91 5.76 (4.37, 7.60) 11 19 0.83 14.32 0.28
GM05 2.57521 75.25468 1282.2 32 57 0.000088 24.4 1.79 0.38 12.35 5.37 (4.11, 7.01) 21 32 0.84 14.72 1.23
GM05-1 2.33 0.36 17.0 12.20 (5.4, 27.7) 3 5 0.83 14.82 1.23
GM05-2 2.28 0.45 14.7 5.03 (3.79, 6.67) 18 27 0.84 14.70 1.25
GM06 2.59105 75.23712 1405.0 40 31 0.000051 99.9 1.60 0.34 6.25 5.04 (3.51, 7.23) 21 28 0.83 14.41 1.28
Note: Ns—number of spontaneous fission tracks; ρs = density (by area) of spontaneous fission tracks; P(χ2) = chi-squared probability that all single grain ages represent a single age population (<5% indicates
nonhomogeneity); Dpar—mean fission track etch pit diameter parallel to crystallographic c-axis; Dper—mean fission track etch pit diameter perpendicular to crystallographic c-axis; conc.—concentration; rmr0 —reduced
length of the more-resistant apatite at the time-temperature conditions where the reduced length of the less-resistant apatite falls to zero.

cene removal of at least 3–4 km of overburden, assuming typical geothermal deformation boundary along the easternmost Neiva Basin). Although neotec-
gradients (20–25 °C/km). The calculated AFT ages therefore represent mini- tonic data and current geodetic measurements show that the Algeciras fault
mum estimates for the onset of exhumation, suggesting that displacement on system currently accommodates overall right-lateral motion, the significant
the main strand of the Algeciras fault system initiated prior to the AFT cooling structural relief across the fault (with basement rocks juxtaposed against Plio-
ages. The three samples from the western side of the fault (samples GM01– cene–Quaternary basin fill) and its low-angle orientation (~15° east-southeast
GM03) recorded steady, continuous exhumation since ca. 6 Ma, with an aver- dip) are compatible with an earlier history involving major dip-slip displace-
age cooling rate of ~20–25 °C/m.y. (Fig. 6). It is intriguing that the samples from ment (Rodríguez et al., 2003; Bakioglu, 2014).
the eastern side of the Algeciras fault (samples GM04–GM06) consistently The AFT cooling paths (Fig. 6) are considered to record a temporal transi-
show convex-upward cooling paths with rapid exhumation from ca. 6 to 4 Ma tion in exhumation patterns in which the eastern (hanging wall) samples were
followed by a protracted period of slow, near-surface cooling (Fig. 6). These rapidly exhumed prior to 4–3 Ma during thrust-reverse displacement along the
thermal histories reflect the extremely long tracks preserved in these samples, southeast-dipping Algeciras fault, and then slowly approached the surface by
in that the only thermal models producing acceptable results require that the erosion after horizontal transverse motion began to dominate along the fault.
samples were exhumed rapidly and then remained near the surface where This is consistent with plate reconstructions that show that post-Miocene Nazca–
rates of track annealing were diminished (Fig. 6). South America plate convergence became increasingly oblique as the northern
Although our sediment provenance results suggest that exhumation of Andean block began to rigidly escape to the northeast (Freymueller et al., 1993;
parts of the Garzón Massif had commenced by 12.5–11.6 Ma, and previous AFT Kellogg and Vega, 1995; Trenkamp et al., 2002; Boschman et al., 2014). More-
studies from areas farther south suggest rapid cooling of the Garzón Massif over, samples on the western (footwall) side of the fault (samples GM01–GM03)
ca. 14–9 Ma (van der Wiel, 1991), our AFT ages require the greatest amounts of show more continuous cooling paths from 4 to 3 Ma onward, consistent with
exhumation (at least 3–4 km) within the central segment of the Garzón Massif propagation of transpressive deformation and westward encroachment into the
to have occurred after ca. 6 Ma. However, the conflicting thermal histories sug- Neiva Basin, possibly related to strain partitioning as the Algeciras fault began
gest important temporal and spatial variations in exhumation-induced cooling to accommodate more horizontal transverse motion. The further issue of older
on the eastern and western flanks of the fault, and potential variations from (broadly 14–9 Ma) AFT ages farther south in the Garzón Massif (van der Wiel,
north to south within the Garzón Massif. The convex-upward cooling paths 1991) could be related to the regional north to northeast plunge of the massif,
(Fig. 6) displayed by the samples from the eastern fault block (samples GM04– or could reflect an additional north to northeastward along-strike progression of
GM06) reflect a significant reduction in exhumation rate since 4–3 Ma, which deformation, possibly related to a long-term transition to a more oblique trans-
could be indicative of (1) a transition from prevailing dip-slip shortening to pressional configuration (e.g., Acosta et al., 2004). Collectively, the AFT results
strike-slip displacement (with diminished generation of structural relief along suggest that the most rapid vertical exhumation of the Garzón Massif occurred
the straight fault segment with a linear surface trace) and/or (2) reduced dis- in latest Miocene–early Pliocene time (ca. 6–4 Ma) with subsequently slower
placement due to westward propagation of the locus of upper crustal deforma- basement exhumation, greater transverse motion, and focused activity along
tion in this central segment of the Garzón Massif (likely reaching the modern different faults as a result of strain partitioning after ca. 4–3 Ma.

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1246
Research Paper

PALEOSOL GEOCHEMISTRY MAP = 18.64 · CIA-K – 350.4 (1)

and
The preceding results help constrain the timing of exhumation for the
Garzón Massif, but do not provide direct constraints on surface uplift. To detect MAP = 22.69 · CALMAG – 435.8. (2)
climatic changes that could be driven by the development of surface topog­
raphy, we use two paleoclimatic proxies that could potentially help define the The CIA-K calibration has a standard error of ±146 mm/yr (Sheldon et al.,
onset of the semiarid conditions now characterizing the Upper Magdalena 2002), and the CALMAG calibration has a standard error of ±108 mm/yr in the
Valley (Tatacoa Desert) along the western flank of the Garzón Massif. Today, calibration data set (Nordt and Driese, 2010). Although the CIA-K index has
the 2500–3000-m-high massif forms an effective orographic barrier, inter- been used to reconstruct precipitation in a variety of soil types from temper-
cepting moisture derived from easterly trade winds (Rozanski and Araguás, ate climates (e.g., Sheldon et al., 2002; Hamer et al., 2007), there are scant
1995; Poveda and Mesa, 1997). The modern climate of the Tatacoa Desert is data from Vertisols and Inceptisols, which are the predominant soil types in
characterized by an intensely evaporative environment; although it receives the Neiva Basin (Guerrero, 1993). We therefore also apply the Vertisol-specific
1300 mm/yr of precipitation, the potential evapotranspiration is estimated to CALMAG calibration in order to develop a model that is more representa-
be >1600 mm/yr, resulting in some of the most arid conditions in Colombia tive of the Neiva Basin, and to compare with the more thoroughly studied
(Insti­tuto de Hidrología, Meteorología y Estudios Ambientales, 2012). We uti- CIA-K index.
lize two independent paleoclimatic proxies to differentiate between several
possible scenarios for the uplift history of the range.
Typically in paleoelevation studies, oxygen isotope compositions of soil Methods
carbonates are assumed to be primarily controlled by Rayleigh distillation, and
paleosol carbonates reflect the decrease in d18O values of meteoric water at We collected paired soil carbonate and bulk paleosol samples from hori-
higher elevations (e.g., Garzione et al., 2006; Quade et al., 2007; Hoke et al., zons where soil carbonate nodules were present. Pedogenic carbonate nod-
Supplemental Table 5: Paleosol carbonate nodule stable isotopic data.
2009). While this effect is also observed in the orographic rain shadow of a ules were processed then analyzed for d13C and d18O values. Nodules were
Strat
Level
Sample Nodule Amplitude 44
number (mV)
Area 44 13C
(VPDB)
13C
StdDev
18
(VSMOW)
18
(VPDB)
18
StdDev
mountain range, other effects such as aridification resulting from the develop- first sawed in half and inspected for signs of alteration and recrystallized
ment of orographic barriers may cause meteoric waters to be 18O enriched due spar; 3–4 individual nodules showing limited alteration and recrystalliza-
1 LV1-17 1 4955 22.621 -10.37 0.011 21.50 -9.13 0.031
2 5632 25.699 -10.20 0.013 20.99 -9.62 0.021
3 4741 21.598 -10.31 0.024 21.38 -9.25 0.029
4 4152 18.994 -10.38 0.014 21.25 -9.37 0.023
2 060811-01 1
2
5788
6289
26.311
28.592
-9.54
-9.59
0.011
0.013
20.57
20.84
-10.03
-9.76
0.030
0.013
to evaporation. This scenario is precisely that expected for the Neiva B ­ asin: tion were selected from each horizon. The nodules were then drilled with
3 4781 21.652 -9.87 0.011 21.43 -9.19 0.026

3 LV1-03A
4
1*
7656
4070
34.841
20.958
-9.84
-11.22
0.019
0.214
21.22
23.13
-9.40
-7.55
0.022
0.065 although meteoric waters should be depleted after their westward transit over an abrasive microtool, and 150–550 µg of powder from each nodule was
2 1866 9.655 -11.20 0.013 21.70 -8.94 0.031

4 LV1-20B
3
1
2
4365
5459
6390
22.371
24.886
29.216
-11.05
-11.13
-10.95
0.012
0.016
0.026
21.80
22.49
22.55
-8.84
-8.17
-8.11
0.022
0.011
0.012
the Garzón Massif, they should also be increasingly subject to evaporation placed into septum-capped Exetainer vials. The vials were flushed with ultra­
5 LV2-03A
3
4
1
4381
4356
3407
19.998
19.788
15.404
-10.38
-10.50
-10.32
0.012
0.018
0.008
21.94
21.90
21.32
-8.70
-8.74
-9.30
0.024
0.023
0.041
as the effect of the rain shadow intensifies and blocks further moisture from high-purity helium and reacted with 0.1 mL of phosphoric acid at 50 °C for
entering the basin. Thus, to properly interpret the oxygen isotope data in 2 h. Carbon dioxide gas in the headspace of each vial was then measured
2 2530 11.354 -10.20 0.047 21.18 -9.44 0.032
3 2239 10.120 -10.20 0.021 21.11 -9.50 0.033
4 2317 10.437 -10.05 0.029 21.11 -9.51 0.024

light of changes in the amount of precipitation, we use the weathering index using continuous flow isotope ratio mass spectrometry in the stable isotope
7
Supplemental Table 5. Paleosol carbonate nodule of the indi­vidual paleosols hosting the analyzed carbonate nodules to estimate laboratory at the University of Texas at Austin. An internal laboratory stan-
stable isotopic data. Please visit http://​dx​.doi​.org​/10​
.1130​/GES01294​.S7 or the full-text article on www​ the effect of increasing aridity associated with surface uplift. dard of powdered marble was run between every seven unknowns to cor-
.gsapubs​.org to view Supplemental Table 5. The weathering indices use well-known solubility relationships between rect for instrument drift, and NBS-18 and NBS-19 standards were analyzed
major cations to estimate the degree of chemical weathering (Maynard, 1992; at the beginning and end of each session to tie the unknowns to the Vienna
Supplemental Table 6: Bulk paleosol X-ray fluorescence (XRF) geochemical data. Sheldon et al., 2002; Sheldon and Tabor, 2009; Nordt and Driese, 2010). To Peedee belemnite (VPDB) scale. Measured values of the laboratory standard
standardize the measurement of weathering across a variety of soils, the show a standard deviation of 0.05‰ in d18O, and 0.1‰ in d13C. The d13C and
Strat Sample SiO2 TiO2 Al2O3 Fe2O3 MnO MgO CaO Na2O K2O P2O5 Sum LOI
Level (%) (%) (%) (%) (%) (%) (%) (%) (%) (%) (%) (%)
3 LV1-03 54.80 0.83 18.37 9.10 0.12 3.14 2.40 1.58 1.62 0.26 92.22 7.65
4 LV1-20A 62.45 0.74 15.04 6.24 0.04 1.89 1.57 1.75 1.37 0.06 91.15 8.73
5 LV2-3B 61.19 0.74 17.09 7.29 0.04 1.96 1.67 0.98 1.53 0.08 92.57 7.29
6
7
8
9
LV2-5A
LV2-6B
060811-3B
060811-8B
60.18
53.93
55.03
52.17
0.83
0.75
0.87
0.78
17.45
18.42
18.42
16.64
8.08
8.69
8.61
8.20
0.05
0.12
0.11
0.17
1.96
3.48
3.25
3.10
1.53
2.82
1.70
5.91
0.72
0.79
0.77
0.81
1.33
1.55
1.65
1.54
0.05
0.17
0.08
0.16
92.18
90.72
90.49
89.48
7.71
9.09
9.36
10.39
abundance of soluble oxides is calculated in reference to oxides resistant to d18O values of the paleosol carbonate nodules are reported in Supplemental
weathering, such as aluminum and silicon oxides. These are collectively re- Table 57.
10 060811-7B 51.11 0.74 14.74 7.07 0.19 2.89 8.64 1.00 1.33 0.13 87.84 12.03
11 060811-6B 61.63 0.76 15.62 6.30 0.08 2.78 2.18 1.74 1.44 0.15 92.68 7.17
12 LV3-12B 62.70 0.68 15.34 7.07 0.03 2.35 2.26 1.29 1.57 0.06 93.35 6.54
13 LVGB-11 54.57 0.78 18.33 8.95 0.06 3.69 1.58 0.79 1.92 0.19 90.86 9.00
14 LVGB-16A 55.02 0.88 17.86 8.59 0.19 3.37 2.20 0.86 1.66 0.16 90.79 9.06
15
17
18
19
LVGB-17A
LVGB-22
LVGB-24B
ECGB-3
55.42
60.93
51.58
54.93
0.77
0.89
0.80
0.94
18.42
21.66
18.50
18.06
7.53
4.13
7.70
9.18
0.03
0.01
0.15
0.10
2.22
1.06
3.43
3.32
1.15
0.62
4.20
1.70
0.89
0.37
0.79
0.83
1.67
1.65
1.67
1.41
0.10
0.10
0.12
0.17
88.20
91.42
88.94
90.64
11.65
8.40
10.92
9.23
ferred to as weathering indices, and have been shown to correlate well with Bulk paleosol samples were crushed with a hammer and homogenized
mean annual precipitation (MAP) (Sheldon et al., 2002; Nordt and Driese, into a fine powder using a mortar and pestle; all tools were cleaned thor-
20 ECGB-9B 67.40 0.72 15.16 4.93 0.08 0.98 1.37 0.72 1.73 0.12 93.21 6.63
21 ECGB-10B 62.48 0.81 18.16 5.89 0.06 1.28 0.92 0.33 1.66 0.11 91.70 8.15
22 PRB-28B 67.89 0.75 15.18 5.17 0.10 1.17 0.50 0.74 1.88 0.15 93.53 6.33
23 PRB-30A 66.65 0.75 15.95 5.93 0.05 1.06 0.39 0.32 1.64 0.11 92.85 6.98
24 PRB-32 61.08 0.82 18.83 6.76 0.06 1.34 0.88 0.19 1.64 0.17 91.77 8.10
26
27
28
28
PRB-66A
PRB-18B
PRB-3A
PRB-3A
67.84
68.81
62.97
57.84
0.74
0.68
0.86
0.82
13.39
12.27
15.45
19.96
5.16
4.19
7.68
7.75
0.13
0.54
0.07
0.06
1.11
1.07
1.92
1.49
1.57
2.89
0.90
0.81
0.07
0.13
0.73
0.00
1.53
1.54
1.77
1.79
0.16
0.12
0.13
0.12
91.70
92.24
92.48
90.64
8.16
7.53
7.38
9.31
2010). Due to the presence of several soil types within the section (Guerrero, oughly with deionized water between samples. Several grams of powder
Standard RGM-1
RGM-1 GIVEN
% Error
74.23
73.45
1.06
0.27
0.27
0
13.66
13.72
0.44
1.89
1.86
1.61
0.04
0.04
11.11
0.26
0.27
3.70
0.99
1.15
13.91
4.13
4.07
1.47
4.32
4.3
0.47
0.06
0.05
25
99.85
99.17
0.68
1993), we utilized two indices shown to correlate well with MAP in a vari- from each sample were fused into a glassy disc using Li2B4 O7, and were ana-
Standard BHVO-1
BHVO-1 GIVEN
% Error
50.23
49.94
0.58
2.73
2.71
0.74
13.65
13.80
1.09
12.33
12.23
0.82
0.17
0.17
1.19
7.32
7.23
1.24
11.35
11.40
0.44
2.30
2.26
1.77
0.52
0.52
0
0.27
0.27
1.1
100.87
100.53
0.34
ety of soil types. We calculated MAP based on two proxies, the CIA-K index lyzed for their bulk elemental abundance using wavelength dispersive X-ray
(chemical index of alteration, minus potassium), which is calculated as Al2O3 / fluorescence (XRF) spectrometry at the Michigan State University XRF Lab.
8
Supplemental Table 6. Paleosol X-ray fluorescence
(Al2O3 + CaO + Na2O) × 100 and the CALMAG index, defined as Al2O3 /(Al2O3 + Measurement accuracy was determined by comparison with two laboratory
(XRF) bulk geochemical data. Please visit http://​dx​
.doi​.org​/10​.1130​/GES01294​.S8 or the full-text article CaO + MgO) × 100. We used the following empirical calibrations to reconstruct standards, and XRF bulk geochemical results are reported in Supplemental
on www​.gsapubs​.org to view Supplemental Table 6. paleoprecipitation: Table 68.

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1247
Research Paper

Results and Interpretation the Andean foreland basin from the intermontane Magdalena Valley (Van
Houten and Travis, 1968; Gómez et al., 2005; Horton, 2012; Silva et al., 2013;
The oxygen and carbon isotope compositions for 28 sets of carbonate nod- Caballero et al., 2013) and separated the Magdalena River from the Orinoco
ules were measured; bulk elemental abundances of paleosol samples from the and Amazon Rivers (Fig. 8). This topographic development also induced an
same interval were measured for 24 of those samples. Average d18O values for orographic rain shadow, which is expressed in the modern semiarid Tatacoa
the carbonate nodules were remarkably consistent throughout the section, with Desert of the U­ pper Magdalena Valley. The expected signatures of this uplift
an average value of –8.8 ‰ (VPDB) and a standard deviation of 0.7‰. The min- event include (1) rapid exhumation of crystalline basement rocks composing
imum measured d18O value was –9.7‰ near the base of the La Victoria Forma- the Garzón ­Massif and, in the Neiva Basin (Upper Magdalena Valley), (2) shifts
tion, and the maximum value was –7.1‰ (Fig. 7; Table 4). The upper half of the in depositional pathways related to topographic diversion of sediment dis-
Villavieja section displays a systematic increase in d18O values by ~2‰, consis- persal systems, (3) stratigraphic appearance of distinctive erosional products
tent with long-term global cooling during the late Miocene (Zachos et al., 2001). from the Garzón Massif, and (4) postdepositional processes reflective of in-
CIA-K and CALMAG indices were calculated from the measured elemen- creased aridity.
tal abundances in the bulk paleosol samples, and MAP was calculated using AFT data (Fig. 6) help constrain the Neogene exhumation history of the
empirical calibrations developed from reference data sets (Equations 1 and 2; Garzón Massif. Our results suggest that the ca. 6–3 Ma phase of shortening-­
Sheldon et al., 2002; Nordt and Driese, 2010). The CIA-K and CALMAG recon- related basement exhumation and surface uplift (Fig. 8C) was succeeded by re-
structions follow the same trend, although the CALMAG estimate tends to organization and regional partitioning of strain into focused shortening along
show 100–200 mm/yr less precipitation than the CIA-K estimate (Fig. 7; the eastern margin of the Neiva Basin and right-lateral strike-slip motion along
Table­ 5). Both show a pronounced excursion toward drier values in interme- the Algeciras fault (Fig. 8D). Such a shift in deformation style is suggested by
diate levels of the La Victoria Formation, with a gradual upsection increase in a significant decrease in modeled exhumation rates after 4–3 Ma, as observed
precipitation through the Villavieja Formation. for all samples on the uplifted eastern side of the fault. This reconstruction sug-
The relatively small changes in the isotopic composition of pedogenic gests that exhumation and generation of considerable structural relief in the
carbonate nodules combined with the slight increase in precipitation recon- central segment of the Garzón Massif was most rapid during latest Miocene–
structed by the bulk elemental abundances in the host paleosols suggest that early Pliocene shortening, followed by slower exhumation when transpres-
the modern rain shadow was not present during ca. 14–12 Ma accumulation sional motion became well established after 4–3 Ma (Fig. 8). For the western
of the Honda Group. The lack of a large orographic barrier is consistent with side of the Algeciras fault, continuous exhumation since ca. 4.5 Ma, modeled
provenance data from the La Venta area in suggesting that the earliest fault- at a more moderate cooling rate, likely represents exhumation along a trans-
ing in the east did not occur until a period closer to the La Victoria–Villavieja pressional structure that propagated westward into the Neiva Basin.
transition, during the latest middle Miocene. Therefore, initial exhumation and Sediment provenance results from the Upper Magdalena Valley provide in-
probable surface uplift of the Garzón Massif (southernmost Eastern Cordillera) sight into the early exhumational record of the Garzón Massif within the East-
was only just under way during deposition of the uppermost Villavieja Forma- ern Cordillera, prior to the latest Miocene–Quaternary history delineated by
tion, ca. 12–11 Ma (Fig. 8). the AFT results. From the Neogene compositional and paleocurrent data sets
(Fig. 5), the earliest evidence for uplift-induced exhumation east of the Neiva
Basin is expressed in the ca. 12.5 Ma influx of nonvolcanic lithic fragments,
DISCUSSION including chert, likely recycled from Cretaceous–Paleogene sedimentary units
(Figs. 8A, 8B). Although this sediment source is replaced shortly thereafter, it
Basement Uplift and Exhumation may represent early relief generation along faults within the Garzón Massif–­
Neiva Basin transition and associated establishment of a drainage barrier re-
New results bearing on sediment provenance and basement exhumation lated to basement uplift. A more pronounced provenance shift is detected in
in southern Colombia provide an important context for the growth of extensive the uppermost Honda Group (ca. 12.5–11.6 Ma) of the northern Neiva Basin
Andean barriers and the evolution of the Magdalena, Orinoco, and Amazon (La Venta site; Fig. 2B), where possible Cenozoic zircons indicative of a west-
river systems (Figs. 1 and 2). Topographic growth of major basement blocks ern source are replaced by the first appearance of Garzón-derived grains. In
in the northern Andes and adjacent foreland has diverted the Orinoco River the southern Neiva Basin (Gigante site; Fig. 2C), a similar shift is observed in
eastward, revised the northern limit of the Amazon drainage, and funneled the the overlying Gigante Formation, in which a significant fraction of clastic mate-
north-flowing Magdalena River along the narrow hinterland segment between rial appears to be derived principally from the magmatic arc to the west until ca.
the Central Cordillera and Eastern Cordillera of Colombia. Basement uplift of 6.4 Ma, the age of the uppermost sample, which is overwhelmingly dominated
the Garzón Massif, which defines the southernmost Eastern Cordillera in the by detritus shed from the Garzón Massif. This north-south age discrepancy in
Colombian Andes (Fig. 2), forms a critical element of this history, as it divided the first appearance of Garzón-derived material may be related to the over-

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1248
Research Paper

Oxygen Isotopic Composition Mean Annual Precipitation (mm/yr)


(‰ VPDB)

1750

1500

1250

1000

750

500
–15 –10 –5 Neiva Basin Stratigraphic Succession

A B 28 Paleocurrent
Direction

27
N = 20
26
25

Villavieja Formation
N = 19
24
Figure 7. Isotopic and calculated paleo­
23
precipitation values from carbonate nod­
22 ules and paleosol samples, plotted relative
to the stratigraphic section at the La Venta
21 site. Stratigraphic levels of individual sam­
20
N = 25,22 ples (1–28) are indicated by horizontal bars:
19 black bars indicate that both carbonate
nodules and soil material were measured

Honda Group
18
at that stratigraphic level, whereas gray
bars indicate that nodules were measured.
17
16 (A) Measured isotopic compositions of soil
15 carbonate nodules, showing δ18O values
14
13 (black circles) and δ13C values (gray trian­
12 gles). Each point represents an average of
3–4 nodules measured at each horizon. The
11 width of the shaded region represents the
N = 24,21 standard deviation. VPDB—Vienna Peedee
belemnite. (B) Mean annual precipitation
reconstructed from the CIA-K (chemical
index of alteration, minus potassium; gray
diamonds) and CALMAG [Al2 O3 /(Al2 O3
+CaO + MgO) × 100; white squares) indices.

La Victoria Formation
Error envelope represents the calibration
N = 21,22 error for each proxy reported in the global
calibration studies (Nordt and Driese, 2010;
Sheldon et al., 2002).

10
9
8 N = 20
7

6
5

100 m
4
3 N = 22
2
1

δ18O (‰ VPDB) CIA-K Calibration

Gravel
Medium sand
Coarse sand
Fine sand
Silt
Clay
CALMAG Calibration

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1249
Research Paper

TABLE 4. SUMMARY OF PALEOSOL CARBONATE NODULE STABLE ISOTOPIC DATA


Stratigraphic Sample Sample Number of δ13C Standard deviation δ18O Standard deviation
level transect name nodules (‰ VPDB) δ13C (‰ VPDB) δ18O
1 1 LV1-17 4 –10.31 0.08 –9.08 0.21
2 1 060811-01 4 –9.71 0.17 –9.60 0.38
3 1 LV1-03A 2 –11.12 0.11 –8.89 0.07
4 1 LV1-20B 4 –10.74 0.36 –8.43 0.34
5 1 LV2-03A 4 –10.19 0.11 –9.44 0.10
6 1 LV2-5B 4 –9.67 0.79 –9.68 0.08
7 1 LV2-06A 4 –9.31 0.07 –9.15 0.04
8 1 060811-03A 4 –10.03 0.14 –9.52 0.14
9 1 060811-08A 4 –9.57 0.15 –8.81 0.11
10 1 060811-07A 2 –10.90 0.68 –8.85 0.57
11 1 060811-06A 4 –8.79 3.75 –8.91 0.66
12 3 LV3-12A 4 –12.76 1.28 –9.66 0.09
13 3 LVGB-10 4 –10.01 0.24 –8.42 0.15
14 3 LVGB-16B 1 –10.55 N/A –9.36 N/A
15 3 LVGB-17B 4 –10.27 0.08 –9.18 0.28
16 3 LVGB-20 4 –10.83 0.49 –8.50 0.21
17 3 LVGB-21 4 –11.79 0.43 –8.75 0.35
18 3 LVGB-24A 4 –10.56 0.30 –8.88 0.23
19 3 ECGB-05 4 –16.86 1.62 –8.52 0.23
20 3 ECRB-10A 4 –12.14 0.47 –7.15 0.78
21 3 ECRB-09A 4 –13.06 1.93 –8.07 0.40
22 4 PRB-28A 4 –3.27 3.96 –9.37 0.44
23 4 PRB-31A 4 –9.03 1.17 –8.39 0.45
24 4 PRB-33A 4 –4.45 1.27 –9.14 0.45
25 4 PRB-62B 4 –14.40 0.77 –8.31 0.85
26 4 PRB-66B 4 –13.32 1.32 –8.28 0.23
27 5 PRB-18A 2 –11.62 0.77 –7.63 0.76
28 5 PRB-03A 4 –9.57 0.66 –7.09 1.34
Note: Values represent the average of 2–4 individual nodule measurements for each horizon. Complete paleosol carbonate nodule stable isotopic data are reported
in Supplemental Table 5 (see footnote 5). VPDB—Vienna Peedee belemnite; N/A—not applicable.

all north-to-northeast plunge of the basement-involved Garzón Massif struc- face uplift of the Eastern Cordillera. Oxygen isotopic compositions of soil car-
tures, with (1) spatial variations in the magnitude of deformation, (2) a spatial bonates in the Neiva Basin display little change (Fig. 7), and weathering-index–
progression in the timing of uplift-induced exhumation in the massif, and/or based estimates of mean annual precipitation suggest that the basin became
(3) the greater proximity of the Central Cordillera to the southern (­Gigante) site. wetter, rather than drier, as would be expected with the development of an
These options suggest a paleogeographic configuration in which large-scale orographic rain shadow produced by surface uplift in the east. Therefore, al-
exhumation of deeper metamorphic basement within the central segment of though the sedimentary provenance data suggest possible early growth of
the Garzón Massif was unlikely to have commenced until ca. 6.4 Ma. How- eastern sediment sources in the Garzón Massif, the paleoclimatic data suggest
ever, there is evidence for exhumation of part of the Garzón Massif (possibly that it is unlikely that these middle Miocene sources constituted a significant
focused to the south of the study region) from 12.5 Ma onward, as evidenced topographic barrier.
by periodic appearances of Garzón-derived material, a gradual transition from These findings are in agreement with biological data, which suggest that
east-directed to north-directed paleocurrents, an increased sedimentation rate, Magdalena Valley flora and fauna were becoming distinct from the Amazon
and a change in sedimentary architecture (Fig. 5). Basin by ca. 10 Ma (Hoorn et al., 2010; Ochoa et al., 2012; Aguilera et al., 2013).
Paleoclimatic data spanning the 13.8–11.6 Ma deposition of the Honda The possible divergence of the Magdalena River from the proto–Orinoco River
Group suggest minimal evidence for aridification that could be linked to sur- and proto–Amazon River systems at this time lends support to the hypothesis

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1250
Research Paper

TABLE 5. PALEOSOL X-RAY FLUORESCENCE BULK GEOCHEMICAL DATA AND CALCULATED WEATHERING INDICES AND PRECIPITATION ESTIMATES
Elemental abundances Weathering indices Mean annual precipitation

Stratigraphic Sample Sample Al2O3 MgO CaO Na2O CIA-K CALMAG


level transect name (%) (%) (%) (%) CIA-K CALMAG (mm/yr) (mm/yr)

3 1 LV1-03 18.37 3.14 2.40 1.58 82.2 76.8 1181.7 1307.5


4 1 LV1-20A 15.04 1.89 1.57 1.75 81.9 81.3 1176.5 1408.8
5 1 LV2-3B 17.09 1.96 1.67 0.98 86.6 82.5 1263.4 1435.7
6 1 LV2-5A 17.45 1.96 1.53 0.72 88.6 83.3 1300.7 1455.0
7 1 LV2-6B 18.42 3.48 2.82 0.79 83.6 74.5 1208.2 1254.9
8 1 060811-3B 18.42 3.25 1.70 0.77 88.2 78.8 1293.2 1352.6
9 1 060811-8B 16.64 3.10 5.91 0.81 71.2 64.9 977.4 1036.2
10 1 060811-7B 14.74 2.89 8.64 1.00 60.5 56.1 776.6 837.3
11 1 060811-6B 15.62 2.78 2.18 1.74 79.9 75.9 1139.7 1286.3
12 3 LV3-12B 15.34 2.35 2.26 1.29 81.2 76.9 1163.3 1308.9
13 3 LVGB-11 18.33 3.69 1.58 0.79 88.6 77.7 1300.2 1326.5
14 3 LVGB-16A 17.86 3.37 2.20 0.86 85.4 76.2 1240.9 1293.8
15 3 LVGB-17A 18.42 2.22 1.15 0.89 90.0 84.5 1327.7 1482.3
17 3 LVGB-22 21.66 1.06 0.62 0.37 95.6 92.8 1432.1 1669.9
18 3 LVGB-24B 18.50 3.43 4.20 0.79 78.8 70.8 1117.6 1170.6
19 3 ECGB-3 18.06 3.32 1.70 0.83 87.7 78.2 1284.6 1339.7
20 3 ECGB-9B 15.16 0.98 1.37 0.72 87.9 86.6 1287.8 1528.7
21 3 ECGB-10B 18.16 1.28 0.92 0.33 93.6 89.2 1393.6 1588.0
22 4 PRB-28B 15.18 1.17 0.50 0.74 92.4 90.1 1372.8 1608.3
23 4 PRB-30A 15.95 1.06 0.39 0.32 95.7 91.7 1434.2 1644.1
24 4 PRB-32 18.83 1.34 0.88 0.19 94.6 89.5 1413.4 1593.9
26 4 PRB-66A 13.39 1.11 1.57 0.07 89.1 83.3 1310.2 1454.8
27 5 PRB-18B 12.27 1.07 2.89 0.13 80.2 75.6 1145.4 1279.6
28 5 PRB-3A 15.45 1.92 0.90 0.73 90.5 84.6 1335.7 1483.0
28 5 PRB-3A 19.96 1.49 0.81 0.00 96.1 89.7 1440.9 1598.8
Note: Complete paleosol X-ray fluorescence bulk geochemical data are reported in Supplemental Table 6 (see text footnote 6). CIA-K—chemical index of
alteration, minus potassium; CALMAG index—Al2O3 /(Al2O3 +CaO + MgO) × 100.

that the modern South American drainage system developed over a protracted from ca. 12.5 to 6.4 Ma, shortening was accommodated along a principally dip-
period beginning in the late Miocene (Figueiredo et al., 2009; Hoorn et al., 2010; slip segment of the Algeciras fault system east of the Neiva Basin, which inter­
Shephard et al., 2010). In order to reconcile the results of our sediment prove- mittently blocked eastward drainage, but did not form a major topographic
nance, paleoclimatic, and thermochronometric studies, we propose that initial barrier at these latitudes (1°–4°N). The intermediate stage commenced during
segregation of these three continental-scale river systems commenced with deposition of the basal Garzón Member of the Gigante Formation, with the ar-
earliest structural development of the Garzón Massif in southern Colombia rival of coarse gravel bearing metamorphic clasts and zircons with U-Pb ages
ca. 12–10 Ma, but substantial topography in the southernmost Eastern Cor­di­ consistent with a Garzón Massif source. From ca. 6.4 to 4–3 Ma, this stage
llera was not generated until the latest Miocene–Pliocene (ca. 6–3 Ma) phase of involved extremely rapid exhumation as upper crustal shortening was accom-
rapid exhumation of Garzón basement. modated by continued thrust displacement within the Algeciras fault system.
During the final stage, from 4 to 3 Ma onward, exhumation of the Garzón
Massif slowed considerably as significant transpressional deformation was
Tectonic Reconstruction accommodated by strike-slip motion along the main Algeciras fault strand and
dip-slip motion along thrust faults bounding the easternmost Neiva Basin. This
From the detrital zircon U-Pb, sandstone petrographic, and AFT exhuma- history of uplift for the Garzón Massif is similar to the timing of sharp surface
tional results, we interpret a three-stage history for the Garzón Massif in the uplift for the Bogotá plateau along strike to the northeast (Hooghiemstra et al.,
Eastern Cordillera of southern Colombia (Fig. 8). During the earliest stage, 2006; Mora et al., 2008; Anderson et al., 2015), suggesting that the most rapid

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1251
Research Paper

phase of regional uplift occurred between 6 and 3 Ma. The synchroneity of


A >12.5 Ma
B ca. 12.5–6 Ma uplift for these two study localities over a >250 km distance along strike within
La V La V

era
the Eastern Cordillera is suggestive of a shared geodynamic driver.

era
enta enta
Our results suggest that basement uplift of the Garzón Massif may be in-

dill

dill
dicative of a significant geodynamic shift in the northern Andes. Modern con-

Cor

Cor
vergence between the Nazca and South American plates is currently oblique

tral

tral
along the Colombian margin, with regional tectonic escape of the northern

Cen

Cen
Andean block to the northeast. Various mechanisms for the initiation of tec-
tonic escape have been proposed. These include the collision of the Panama
arc (Farris et al., 2011; Montes et al., 2012), the upper plate response to subduc-
tion of the buoyant Carnegie Ridge (Gutscher et al., 1999; Spikings et al., 2001;
Egbue and Kellogg, 2010; Egbue et al., 2014), or a shift in convergence direc-
tion or strain partitioning along the obliquely converging Nazca–South Amer-

e
ican margin (Acosta et al., 2004; Jiménez et al., 2014). Our interpretation of a

nt

nt
ga

ga
Pliocene transition to principally strike-slip motion along the ­Alge­ciras fault
Gi

Gi
25 km 25 km
is most consistent with a shift to more oblique convergence, greater strain
- East-directed drainage systems. - Axial (north-directed) drainage initiates. partitioning, and/or the potential influence of the geodynamic effects of the
- Central Cordillera source. - Central and Eastern Cordillera sources. subducting Carnegie Ridge.
- No eastern topographic barrier. - Low topographic barrier to east. In addition, due to the shared basement configuration and close proximity
of the Garzón Massif to the foreland structures forming the Macarena high
C ca. 6–3 Ma
D ca. 3 Ma–Present and broader Vaupes arch (e.g., Casero et al., 1997; Acosta et al., 2004; Velandia
era

era
La V La V et al., 2005), we speculate that their uplift may share a driving geodynamic
enta enta
dill

dill
mechanism, consistent with previous work suggesting a similar timing based
Cor

Cor
on structural and stratigraphic constraints (Wesselingh et al., 2006; Mora et al.,
tral

tral
2010a). These foreland basement arches form the barrier between the mod-
ern-day Orinoco and Amazon river systems (Figs. 1 and 2), and may play a key
Cen

Cen
role in the initiation of a transcontinental Amazon drainage. This underscores
lt potential links among plate boundary processes, topographic uplift, and re-

au
gional drainage patterns in affecting the biological dynamics of South America.

f
ras
ec i
u lt

a Alg
sf ra
c i ra lle
a

ge
di
er

Al r CONCLUSIONS
ill

Co
rd

rn
Co

e
e

ste
nt
nt

Ea
ga
ga

er

We characterize Neogene topographic development of the Garzón Mas-


st

Gi
Gi

25 km 25 km
Ea

sif (southernmost Eastern Cordillera) of the northern Andes by analyzing the


- Thrusting and exhumation of Garzón - Thrust front propagates westward. sediment provenance of clastic basin fill, AFT thermochronometry of base-
massif along Algeciras Fault. - Strike-slip motion initiates along ment rocks, and paleosol-based indicators of climatic variations in the Upper
- Dominant Eastern Cordillera source. Algeciras Fault. Magdalena Valley (Neiva Basin) of southern Colombia. Our reconstructions of
Cretaceous-Paleogene Jurassic and Cenozoic Garzón metamorphic middle Miocene–Quaternary tectonics, exhumation, and paleodrainage (Fig. 8)
sedimentary units igneous units basement
emphasize the following points.
Figure 8. Schematic reconstruction of the middle Miocene to present history for the Neiva Basin and 1. Sediment sources for the Neiva Basin alternated between principally
Garzón Massif. (A) A pre-uplift configuration with a Central Cordillera sediment source generating magmatic-arc rocks of the Central Cordillera and probable Eastern Cordillera
east-flowing rivers. (B) Initial uplift (ca. 12.5–6 Ma) of the Garzón Massif (Eastern Cordillera), with
sources from the sedimentary cover and basement rocks of the Garzón Mas-
Jurassic–­Cretaceous units exposed by thrusting and a relatively low topographic barrier forcing initial
reorientation into a north-directed axial drainage. (C) Intensified uplift and exhumation (6–3 Ma) of the sif. Most of the middle to upper Miocene basin fill was derived from western,
Eastern Cordillera, with initial exposure of Garzón basement. (D) Final stage (3–0 Ma) involving strain igneous-­dominated sources in the Central Cordillera, with an important shift to
partitioning that resulted in diminished basement exhumation, westward propagation of the deforma­
eastern sources and pronounced contributions from Garzón basement sources
tion front, and strike-slip motion along the Algeciras fault system.
by at least 6.4 Ma. However, earliest exhumation of the Eastern Cordillera at

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1252
Research Paper

Anderson, V.J., Saylor, J.E., Shanahan, T.M., and Horton, B.K., 2015, Paleoelevation records from
these latitudes (1°–4°N) along upper crustal thrust structures defining the east- lipid biomarkers: Application to the tropical Andes: Geological Society of America Bulletin,
ern margin of the Neiva Basin may be recorded by a focused ca. 12.5 Ma influx v.  127, p.  1604–1616; doi:​10​.1130​/B31105​.1​.
of nonvolcanic lithic fragments suggestive of initial stripping of Cretaceous Aspden, J.A., McCourt, W.J., and Brook, M., 1987, Geometrical control of subduction-related
magmatism: The Mesozoic and Cenozoic plutonic history of western Colombia: Journal of
sediment cover rocks capping the Garzón Massif. the Geological Society [London], v. 144, p. 893–905, doi:​10​.1144​/gsjgs​.144​.6​.0893​.
2. Minimal paleoclimatic changes attributable to topographic uplift are Baker, P.A., Fritz, S.C., Dick, C.W., Eckert, A.J., Horton, B.K., Manzoni, S., Ribas, C.C., Garzione,
detected in the 13.8–11.6 Ma interval. Evidence for the generation of an oro- C.N., and Battisti, D.S., 2014, The emerging field of geogenomics: Constraining geological
problems with genetic data: Earth-Science Reviews, v. 135, p. 38–47, doi:​10​.1016​/j​.earscirev​
graphically induced rain shadow is lacking, in that reconstructed precipitation .2014​.04​.001​.
is shown to increase, rather than decrease, over this time interval. In com- Bakioglu, K.B., 2014, Garzón Massif Basement Tectonics: A Geophysical Study, Upper Magda-
bination with the shifting sediment sources inferred from provenance analy- lena Valley, Colombia [M.S. thesis]: Columbia, University of South Carolina, 54 p., http://​
scholarcommons​.sc​.edu​/cgi​/viewcontent​.cgi​?article​=3793​&context​=etd.
ses, these data suggest the Garzón Massif did not form a substantial elevated
Bayona, G., Cardona, A., Jaramillo, C., Mora, A., Montes, C., Valencia, V., Ayala, C., Montenegro,
topographic barrier during the early stages of basement uplift from 12.5 Ma O., and Ibañez-Mejia, M., 2012, Early Paleogene magmatism in the northern Andes: Insights
until 6.4 Ma. on the effects of oceanic plateau–continent convergence: Earth and Planetary Science Let-
ters, v.  331–332, p.  97–111, doi:​10​.1016​/j​.epsl​.2012​.03​.015​.
3. Inverse modeling results of thermal histories defined by new AFT data
Boschman, L.M., van Hinsbergen, D.J.J., Torsvik, T.H., Spakman, W., and Pindell, J.L., 2014,
suggest profoundly different exhumation histories for opposing flanks of the Kinematic reconstruction of the Caribbean region since the Early Jurassic: Earth-Science
composite Algeciras fault system. This pattern may be attributed to a regional Reviews, v.  138, p.  102–136, doi:​10​.1016​/j​.earscirev​.2014​.08​.007​.
Bustamante, C., Cardona, A., Bayona, G., Mora, A.R., Valencia, V., Gehrels, G.E., and Vervoort, J.,
Pliocene reorganization from principally dip-slip thrust motion to strike-slip
2010, U-Pb LA-ICP-MS geochronology and regional correlation of Middle Jurassic intrusive
motion along the Algeciras fault. Based on the minimum exhumational ages rocks from the Garzón Massif, Upper Magdalena Valley and Central Cordillera, southern
provided by the AFT data, as well as the thermal history modeling results, Colombia: Boletin de Geologia, v. 32, p. 93–109.
we suggest that thrust-induced rapid exhumation was focused between ca. 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–
6.4 Ma and 4–3 Ma. Thereafter, the Algeciras fault likely accommodated strike- 120, doi:​10​.1016​/0895​-9811​(88)90019​-3​.
slip displacement, consistent with estimates for right-lateral motion since at Caballero, V., Mora, A., Quintero, I., Blanco, V., Parra, M., Rojas, L.E., Lopez, C., Sanchez, N., Hor-
least 2 Ma. ton, B.K., Stockli, D., and Duddy, I., 2013, Tectonic controls on sedimentation in an intermon-
tane hinterland basin adjacent to inversion structures: The Nuevo Mundo syncline, Middle
4. This sequence of events is consistent with the timing of major uplift and Magdalena Valley, Colombia, in Nemčok, M., et al., eds., Thick-Skin-Dominated Orogens:
exhumation elsewhere in the Eastern Cordillera of Colombia, suggesting a From Initial Inversion to Full Accretion: Geological Society of London Special Publication
shared geodynamic driving mechanism over a large segment of the north- 377, p.  315–342, doi:​10​.1144​/SP377​.12​.
Campbell, K.E., 2010, Late Miocene onset of the Amazon River and the Amazon deep-sea fan:
ernmost Andes. We suggest that basement uplift in the Eastern Cordillera Evidence from the Foz do Amazonas Basin: Comment: Geology, v. 38, e212, doi:​10​.1130​
and northern Andean foreland basin forced the late Miocene isolation of the /G30633C​.1​.
Magdalena, Orinoco, and Amazon drainage systems. By inference, the latest Campbell, K.E., Frailey, C.D., and Romero-Pittman, L., 2006, The Pan-Amazonian Ucayali pene-
plain, late Neogene sedimentation in Amazonia, and the birth of the modern Amazon River
Miocene–Pliocene (6–3 Ma) phase of rapid and widespread exhumation across
system: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 239, p. 166–219, doi:​10​.1016​
the Eastern Cordillera is hypothesized to have provided a huge pulse of clastic /j​.palaeo​.2006​.01​.020​.
material into these continental-scale river systems. Cardona, A., Chew, D., Valencia, V.A., Bayona, G., Mišković, A., and Ibañez-Mejía, M., 2010, Gren-
villian remnants in the Northern Andes: Rodinian and Phanerozoic paleogeographic per-
spectives: Journal of South American Earth Sciences, v. 29, p. 92–104, doi:​10​.1016​/j​.jsames​
ACKNOWLEDGMENTS .2009​.07​.011​.
Carlson, W.D., Donelick, R.A., and Ketcham, R.A., 1999, Variability of apatite fission-track anneal-
Funding was provided by the National Science Foundation (grants EAR-1019857 and EAR-
ing kinetics: I. Experimental results: American Mineralogist, v. 84, no. 1990, p. 1213–1223,
1338694), Ecopetrol–Instituto Colombiano del Petróleo (ICP), and student research grants from the
doi:​10​.2138​/am​-1999​-0901​.
Geological Society of America. We thank Juliana Barrientos-Mesa and Miguel Corcione for field
Casero, P., Salel, J.-F., and Rossato, A., 1997, Multidisciplinary correlative evidence for polyphase
assistance, and James Coogan, Mauricio Ibañez, Alejandro Mora, Mauricio Parra, Alexis Rosero,
geological evolution of the foot-hills of the Cordillera Oriental (Colombia), in VI Simposio
Timothy Shanahan, Lorena Suarez, and Gabriel Veloza for beneficial discussions. We thank Apa-
Bolivariano: Exploracion Petrolera en las Cuencas Subandinas: Cartagena, Colombia, Asoci-
tite to Zircon, Inc., for providing data at reduced academic rates. The manuscript was improved
acion Colombiana de Geologos y Geofisicos del Petroleo, p. 119–128.
through the constructive reviews of Matías Ghiglione and Annia Fayon.
Chorowicz, J., Chotin, P., and Guillande, R., 1996, The Garzón fault: Active southwestern bound-
ary of the Caribbean plate in Colombia: Geologische Rundschau, v. 85, p. 172–179, doi:​10​
.1007​/s005310050066​.
REFERENCES CITED Colletta, B., Hebrard, F., Letouzey, J., Werner, P., and Rudkiewicz, J.L., 1990, Tectonic style and
Acosta, J., Lonergan, L., and Coward, M.P., 2004, Oblique transpression in the western thrust crustal structure of the Eastern Cordillera, Colombia from a balanced cross section, in
front of the Colombian Eastern Cordillera: Journal of South American Earth Sciences, v. 17, ­Letouzey, J., ed., Petroleum and Tectonics in Mobile Belts: Paris, Editions Technip, p. 81–100.
p.  181–194, doi:​10​.1016​/j​.jsames​.2004​.06​.002​. Cordani, U.G., Cardona, A., Jimenez, D.M., Liu, D., and Nutman, P., 2005, Geochronology of
Aguilera, O., Lundberg, J., Birindelli, J., Sabaj Pérez, M., Jaramillo, C., and Sánchez-Villagra, Proterozoic basement inliers in the Colombian Andes: Tectonic history of remnants of a
M.R., 2013, Palaeontological evidence for the last temporal occurrence of the ancient west- fragmented Grenville belt, in Vaughan, A.P.M., Leat, P.T., and Pankhurst, R.J., eds., Terrane
ern Amazonian river outflow into the Caribbean: PLoS One, v. 8, no. 9, e76202, doi:​10​.1371​ Processes at the Margins of Gondwana: Geological Society of London Special Publication
/journal​.pone​.0076202​. 246, p.  329–346, doi:​10​.1144​/GSL​.SP​.2005​.246​.01​.13​.

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1253
Research Paper

DeCelles, P.G., Langford, R.P., and Schwartz, R.K., 1983, Two new methods of paleocurrent deter- Guerrero, J., 1997, Stratigraphy, sedimentary environments, and the Miocene uplift of the Colom-
mination from trough cross-stratification: Journal of Sedimentary Petrology, v. 53, p. 629– bian Andes, in Kay, R.F., et al., eds., Vertebrate paleontology in the neotropics: The Miocene
642, doi:​10​.1306​/212F824C​-2B24​-11D7​-8648000102C1865D​. fauna of La Venta, Colombia: Washington, D.C., Smithsonian Institution Scholarly Press,
Dengo, C.A., and Covey, M.C., 1993, Structure of the Eastern Cordillera of Colombia: Implica- p. 15–43.
tions for trap styles and regional tectonics: American Association of Petroleum Geologists Gutscher, M.A., Malavieille, J., Lallemand, S., and Collot, J.Y., 1999, Tectonic segmentation of the
Bulletin, v. 77, p. 1315–1337. North Andean margin: Impact of the Carnegie Ridge collision: Earth and Planetary Science
Díaz de Gamero, M.L., 1996, The changing course of the Orinoco River during the Neogene: A Letters, v.  168, p.  255–270, doi:​10​.1016​/S0012​-821X​(99)00060​-6​.
review: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 123, p. 385–402, doi:​10​.1016​ Hamer, J.M.M., Sheldon, N.D., Nichols, G.J., and Collinson, M.E., 2007, Late Oligocene–early
/0031​-0182​(96)00115​-0​. Miocene paleosols of distal fluvial systems, Ebro Basin, Spain: Palaeogeography, Palaeo­
Dickinson, W.R., Beard, L.S., Brakenridge, G.R., Erjavec, J.L., Ferguson, R.C., Inman, K.F., Knepp, climatology, Palaeoecology, v.  247, p.  220–235, doi:​10​.1016​/j​.palaeo​.2006​.10​.016​.
R.A., Lindberg, F.A., and Ryberd, P.T., 1983, Provenance of North American Phanerozoic sand- Heller, P.L., and Paola, C., 1996, Downstream changes in alluvial architecture: An exploration of
stones in relation to tectonic setting: Geological Society of America Bulletin, v. 94, p. 222– controls on channel-stacking patterns: Journal of Sedimentary Research, v. 66, p. 297–306,
235, doi:​10​.1130​/0016​-7606​(1983)94​<222:​PONAPS>2​.0​.CO;2​. doi:​10​.1306​/D4268333​-2B26​-11D7​-8648000102C1865D​.
Donelick, R.A., O’Sullivan, P.B., and Ketcham, R.A., 2005, Apatite fission-track analysis: Reviews Hoke, G.D., Garzione, C.N., Araneo, D.C., Latorre, C., Strecker, M.R., and Williams, K.J., 2009, The
in Mineralogy and Geochemistry, v. 58, p. 49–94, doi:​10​.2138​/rmg​.2005​.58​.3​. stable isotope altimeter: Do Quaternary pedogenic carbonates predict modern elevations?:
Egbue, O., and Kellogg, J., 2010, Pleistocene to Present North Andean “escape”: Tectonophysics, Geology, v.  37, p.  1015–1018, doi:​10​.1130​/G30308A​.1​.
v.  489, p.  248–257, doi:​10​.1016​/j​.tecto​.2010​.04​.021​. Hooghiemstra, H., Wijninga, V.M., and Cleef, A.M., 2006, The paleobotanical record of Colom-
Egbue, O., Kellogg, J., Aguirre, H., and Torres, C., 2014, Evolution of the stress and strain fields bia: Implications for biogeography and diversity: Missouri Botanical Garden Annals, v. 93,
in the Eastern Cordillera, Colombia: Journal of Structural Geology, v. 58, p. 8–21, doi:​10​.1016​ p.  297–325, doi:​10​.3417​/0026​-6493​(2006)93​[297:​TPROCI]2​.0​.CO;2​.
/j​.jsg​.2013​.10​.004​. Hoorn, C., 1994, Fluvial palaeoenvironments in the intracratonic Amazonas Basin (early Mio-
Ehlers, T.A., 2005, Crustal thermal processes and the interpretation of thermochronometer data: cene–early Middle Miocene, Colombia): Palaeogeography, Palaeoclimatology, Palaeoecol-
Reviews in Mineralogy and Geochemistry, v. 58, p. 315–350, doi:​10​.2138​/rmg​.2005​.58​.12​. ogy, v.  109, p.  1–54, doi:​10​.1016​/0031​-0182​(94)90117​-1​.
Escalona, A., and Mann, P., 2006, Sequence-stratigraphic analysis of Eocene clastic foreland Hoorn, C., Guerrero, J., Sarmiento, G.A., and Lorente, M.A., 1995, Andean tectonics as a cause
basin deposits in central Lake Maracaibo using high-resolution well correlation and 3-D for changing drainage patterns in Miocene northern South America: Geology, v. 23, p. 237–
seismic data: American Association of Petroleum Geologists Bulletin, v. 90, p. 581–623, doi:​ 240, doi:​10​.1130​/0091​-7613​(1995)023​<0237​.
10​.1306​/10130505037​. Hoorn, C., et al., 2010, Amazonia through time: Andean uplift, climate change, landscape evolu-
Farris, D.W., Jaramillo, C., Bayona, G., Restrepo-Moreno, S.A., Montes, C., Cardona, A., Mora, A., tion, and biodiversity: Science, v. 330, p. 927–931, doi:​10​.1126​/science​.1194585​.
Speakman, R.J., Glascock, M.D., and Valencia, V., 2011, Fracturing of the Panamanian Isthmus Horton, B.K., 2012, Cenozoic evolution of hinterland basins in the Andes and Tibet, in Busby, C.,
during initial collision with: South America: Geology, v. 39, p. 1007–1010, doi:​10​.1130​/G32237​.1​. and Azor, A., eds., Tectonics of Sedimentary Basins: Recent Advances: Chichester, UK, John
Figueiredo, J., Hoorn, C., van der Ven, P., and Soares, E., 2009, Late Miocene onset of the Amazon Wiley & Sons, Ltd., p. 427–444, doi:​10​.1002​/9781444347166​.ch21​.
River and the Amazon deep-sea fan: Evidence from the Foz do Amazonas Basin: Geology, Horton, B.K., Saylor, J.E., Nie, J., Mora, A., Parra, M., Reyes-Harker, A., and Stockli, D.F., 2010,
v.  37, p.  619–622, doi:​10​.1130​/G25567A​.1​. Linking sedimentation in the northern Andes to basement configuration, Mesozoic exten-
Flynn, J.J., Guerrero, J., and Swisher, C.C.I., 1997, Geochronology of the Honda Group, in Kay, sion, and Cenozoic shortening: Evidence from detrital zircon U-Pb ages, Eastern Cordillera,
R.F., et al., eds., Vertebrate paleontology in the neotropics: The Miocene fauna of La Venta, Colombia: Geological Society of America Bulletin, v. 122, p. 1423–1442, doi:​10​.1130​/B30118​.1​.
Colombia: Washington, D.C., Smithsonian Institution Scholarly Press, p. 44–60. Horton, B.K., Anderson, V.J., Saylor, J.E., Parra, M., and Mora, A.R., 2015a, Application of ­detrital
Folk, R.L., 1980, Petrology of Sedimentary Rocks (second edition): Austin, Texas, Hemphill Press, zircon U-Pb geochronology to surface and subsurface correlations of provenance, paleo­
182 p. drainage, and tectonics of the Middle Magdalena Valley Basin of Colombia: Geosphere,
Freymueller, J.T., Kellogg, J.N., and Vega, V., 1993, Plate motions in the north Andean region: v.  11, p.  1790–1811, doi:​10​.1130​/GES01251​.1​.
Journal of Geophysical Research, v. 98, p. 21,853–21,863, doi:​10​.1029​/93JB00520​. Horton, B.K., Perez, N.D., Fitch, J.D., and Saylor, J.E., 2015b, Punctuated shortening and sub-
Garzione, C.N., Molnar, P., Libarkin, J.C., and MacFadden, B.J., 2006, Rapid late Miocene rise sidence in the Altiplano plateau of southern Peru: Implications for early Andean mountain
of the Bolivian Altiplano: Evidence for removal of mantle lithosphere: Earth and Planetary building: Lithosphere, v. 7, p. 117–137, doi:​10​.1130​/L397​.1​.
Science Letters, v.  241, p.  543–556, doi:​10​.1016​/j​.epsl​.2005​.11​.026​. Ibañez-Mejía, M., Ruiz, J., Valencia, V.A., Cardona, A., Gehrels, G.E., and Mora, A.R., 2011, The
Gehrels, G.E., 2012, Detrital zircon U-Pb geochronology: Current methods and new opportu- Putumayo Orogen of Amazonia and its implications for Rodinia reconstructions: New U-Pb
nities, in Busby, C., and Azor, A., eds., Tectonics of sedimentary basins: Recent advances: geochronological insights into the Proterozoic tectonic evolution of northwestern South
Chichester, UK, John Wiley & Sons, Ltd., p. 45–62, doi:​10​.1002​/9781444347166​.ch2​. America: Precambrian Research, v.  191, p.  58–77, doi:​10​.1016​/j​.precamres​.2011​.09​.005​.
Gehrels, G., 2014, Detrital zircon U-Pb geochronology applied to tectonics: Annual Review of Ingersoll, R.V., Bullard, T.F., Ford, R.F., Grimm, J.P., Pickle, J.D., and Sares, S.W., 1984, The effect of
Earth and Planetary Sciences, v. 42, p. 127–149, doi:​10​.1146​/annurev​-earth​-050212​-124012​. grain size on detrital modes: A test of the Gazzi-Dickinson point-counting method: Journal of
Gehrels, G.E., Valencia, V.A., and Ruiz, J., 2008, Enhanced precision, accuracy, efficiency, and Sedimentary Petrology, v.  54, p.  103–116, doi:​10​.1306​/212F83B9​-2B24​-11D7​-8648000102C1865D​.
spatial resolution of U-Pb ages by laser ablation-multicollector-inductively coupled plasma-­ Instituto de Hidrología, Meteorología y Estudios Ambientales, 2012, Promedios climatológicos
mass spectrometry: Geochemistry, Geophysics, Geosystems, v. 9, Q03017, doi:​10​.1029​ actuales 1971–2000: Mapas Clima: Bogotá, Colombia, Instituto de Hidrología, Meteorología,
/2007GC001805​. y Estudios Ambientales.
Gómez, E., Jordan, T.E., Allmendinger, R.W., Hegarty, K., and Kelley, S., 2005, Syntectonic Ceno- Jiménez, G., Speranza, F., Faccenna, C., Bayona, G., and Mora, A., 2014, Paleomagnetism and
zoic sedimentation in the northern middle Magdalena Valley Basin of Colombia and implica- magnetic fabric of the Eastern Cordillera of Colombia: Evidence for oblique convergence
tions for exhumation of the Northern Andes: Geological Society of America Bulletin, v. 117, and nonrotational reactivation of a Mesozoic intracontinental rift: Tectonics, v. 33, p. 2233–
p.  547–569, doi:​10​.1130​/B25454​.1​. 2260, doi:​10​.1002​/2014TC003532​.
Gómez Tapias, J., Nivia Guevara, A., Montes Ramírez, N.E., Jiménez Mejía, D.M., Tejada Avella, Kellogg, J.N., and Vega, V., 1995, Tectonic development of Panama, Costa Rica, and the Co-
M.L., Sepúlveda Ospina, M.J., Osorio Naranjo, J.A., Gaona Narváez, T., Diederix, H., Uribe lombian Andes: Constraints from global positioning system geodetic studies and gravity,
Peña, H., and Mora Penagos, M., 2007, Mapa Geológico de Colombia (second edition): in Mann, P., ed., Geologic and Tectonic Development of the Caribbean Plate Boundary in
­Bogotá, Ingeominas (Instituto Colombiano de Geología y Minería), scale 1:1,000,000. Southern Central America: Geological Society of America Special Paper 295, p. 75–90, doi:​
Guerrero, J., 1993, Magnetostratigraphy of the upper part of the Honda Group and Neiva Forma- 10​.1130​/SPE295​-p75​.
tion, Miocene uplift of the Colombian Andes [Ph.D. thesis]: Durham, North Carolina, Duke Ketcham, R.A., 2005, Forward and inverse modeling of low-temperature thermochronometry
University, 210 p. data: Reviews in Mineralogy and Geochemistry, v. 58, p. 275–314, doi:​10​.2138​/rmg​.2005​.58​.11​.

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1254
Research Paper

Ketcham, R.A., Donelick, R.A., and Carlson, W.D., 1999, Variability of apatite fission-track an- Parra, M., Mora, A., Lopez, C., Ernesto Rojas, L., and Horton, B.K., 2012, Detecting earliest short-
nealing kinetics: III. Extrapolation to geological time scales: American Mineralogist, v. 84, ening and deformation advance in thrust belt hinterlands: Example from the Colombian
p.  1235–1255, doi:​10​.2138​/am​-1999​-0903​. Andes: Geology, v. 40, p. 175–178, doi:​10​.1130​/G32519​.1​.
Latrubesse, E.M., Cozzuol, M., da Silva-Caminha, S.F., Rigsby, C.A., Absy, M.L., and Jaramillo, Perez, N.D., and Horton, B.K., 2014, Oligocene-Miocene deformational and depositional history
C., 2010, The late Miocene paleogeography of the Amazon Basin and the evolution of the of the Andean hinterland basin in the northern Altiplano plateau, southern Peru: Tectonics,
Amazon River system: Earth-Science Reviews, v. 99, p. 99–124, doi:​10​.1016​/j​.earscirev​.2010​ v.  33, p.  1819–1847, doi:​10​.1002​/2014TC003647​.
.02​.005​. Potter, P.E., 1997, The Mesozoic and Cenozoic paleodrainage of South America: A natural his-
Leeder, M.R., 1977, A quantitative stratigraphic model for alluvium, with special reference to tory: Journal of South American Earth Sciences, v. 10, p. 331–344, doi:​10​.1016​/S0895​-9811​
channel deposit density and interconnectedness, in Miall, A.D., ed., Fluvial Sedimentology: (97)00031​-X​.
Canadian Society of Petroleum Geologists Memoir 5, p. 587–596. Poveda, G., and Mesa, O.J., 1997, Feedbacks between hydrological processes in tropical South
Mackey, S.D., and Bridge, J.S., 1995, Three-dimensional model of alluvial stratigraphy: T ­ heory America and large-scale ocean-atmospheric phenomena: Journal of Climate, v. 10, p. 2690–
and application: Journal of Sedimentary Research, v. 65, no. 1B, p. 7–31, doi:​10​.1306​ 2702, doi:​10​.1175​/1520​-0442​(1997)010​<2690:​FBHPIT>2​.0​.CO;2​.
/D42681D5​-2B26​-11D7​-8648000102C1865D​. Quade, J., Garzione, C., and Eiler, J., 2007, Paleoelevation reconstruction using pedogenic car-
Mann, P., Escalona, A., and Castillo, M.V., 2006, Regional geologic and tectonic setting of the bonates: Reviews in Mineralogy and Geochemistry, v. 66, p. 53–87, doi:​10​.2138​/rmg​.2007​.66​.3​.
Maracaibo supergiant basin, western Venezuela: American Association of Petroleum Geolo- Ramon, J.C., and Rosero, A., 2006, Multiphase structural evolution of the western margin of the
gists Bulletin, v. 90, p. 445–477, doi:​10​.1306​/10110505031​. Girardot subbasin, Upper Magdalena Valley, Colombia: Journal of South American Earth
Marquínez, G., and Velandia, F., 2001, Geología del Departmento de Huila: Bogotá, Colombia, Sciences, v.  21, p.  493–509, doi:​10​.1016​/j​.jsames​.2006​.07​.012​.
Ingeominas, scale 1:300,000. Restrepo-Pace, P.A., Ruiz, J., Gehrels, G., and Cosca, M., 1997, Geochronology and Nd isotopic
Maynard, J.B., 1992, Chemistry of modern soils as a guide to interpreting Precambrian paleo­ data of Grenville-age rocks in the Colombian Andes: New constraints for Late Proterozoic–
sols: Journal of Geology, v. 100, p. 279–289, doi:​10​.1086​/629632​. early Paleozoic paleocontinental reconstructions of the Americas: Earth and Planetary Sci-
Mojica, J., and Franco, R., 1990, Estructura y Evolucion Tectonlca del Valle Medio y Superior del ence Letters, v.  150, p.  427–441, doi:​10​.1016​/S0012​-821X​(97)00091​-5​.
Magdalena, Colombia: Geologia Colombiana, v. 17, p. 41–64. Ribas, C.C., Aleixo, A., Nogueira, A.C.R., Miyaki, C.Y., and Cracraft, J., 2012, A palaeobiogeographic
Montes, C., Hatcher, R.D., Jr., and Restrepo-Pace, P.A., 2005, Tectonic reconstruction of the north- model for biotic diversification within Amazonia over the past three million years: Royal So-
ern Andean blocks: Oblique convergence and rotations derived from the kinematics of the ciety of London Proceedings, ser. B, v. 279, no. 1729, p. 681–689, doi:​10​.1098​/rspb​.2011​.1120​.
Piedras-Girardot area, Colombia: Tectonophysics, v. 399, p. 221–250, doi:​10​.1016​/j​.tecto​.2004​ Roddaz, M., Hermoza, W., Mora, A., Baby, P., Parra, M., Christophoul, F., Brusset, S., and Espurt,
.12​.024​. N., 2010, Cenozoic sedimentary evolution of the Amazonian foreland basin system, in Hoorn,
Montes, C., Bayona, G., Cardona, A., Buchs, D.M., Silva, C.A., Morón, S., Hoyos, N., Ramírez, C., and Wesselingh, F.P., eds., Amazonia: Landscape and species evolution: A look into the
D.A., Jaramillo, C.A., and Valencia, V., 2012, Arc-continent collision and orocline formation: past: Oxford, UK, Wiley-Blackwell Publishing Ltd., p. 61–88, doi:​10​.1002​/9781444306408​.ch5​.
Closing of the Central American seaway: Journal of Geophysical Research, v. 117, B04105, Rodríguez, G., Zapata, G., Velásquez, M.E., Cossio, U., and Londoño, A.C., 2003, Memoria Ex-
doi:​10​.1029​/2011JB008959​. plicatica: Geología de las planchas 367 Gigante, 368 San Vicente del Caguán, 389 Timaná,
Mora, A., Parra, M., Strecker, M.R., Kammer, A., Dimaté, C., and Rodríguez, F., 2006, Cenozoic 390 Puerto Rico, 414 El Doncello y parte de la plancha 391 Lusitania: Instituto Nacional de
contractional reactivation of Mesozoic extensional structures in the Eastern Cordillera of Investigaciones Geológico Mineras, Servicio Geológico Colombiano: Bogotá, Colombia,
Colombia: Tectonics, v. 25, TC2010, doi:​10​.1029​/2005TC001854​. ­Ingeominas, scale 1:200,000.
Mora, A., Parra, M., Strecker, M.R., Sobel, E.R., Hooghiemstra, H., Torres, V., and Jaramillo, Rozanski, K., and Araguás, L.A., 1995, Spatial and temporal variability of stable isotope com-
J.V., 2008, Climatic forcing of asymmetric orogenic evolution in the Eastern Cordillera position of precipitation over the South American continent: Bulletin de l’Institut Français
of Colombia: Geological Society of America Bulletin, v. 120, p. 930–949, doi:​ 10​.1130​ d’études Andines, v. 24, p. 379–390.
/B26186​.1​. Sacek, V., 2014, Drainage reversal of the Amazon River due to the coupling of surface and litho-
Mora, A., Baby, P., Roddaz, M., Parra, M., Brusset, S., Hermoza, W., and Espurt, N., 2010a, Tec- spheric processes: Earth and Planetary Science Letters, v. 401, p. 301–312, doi:​10​.1016​/j​.epsl​
tonic history of the Andes and sub-Andean zones: Implications for the development of the .2014​.06​.022​.
Amazon drainage basin, in Hoorn, C., and Wesselingh, F.P., eds., Amazonia: Landscape and Sarmiento, L., and Rangel, A., 2004, Petroleum systems of the Upper Magdalena Valley, Colom-
Species Evolution: A Look into the Past: Oxford, UK, Wiley-Blackwell, p. 38–60, doi:​10​.1002​ bia: Marine and Petroleum Geology, v. 21, p. 373–391, doi:​10​.1016​/j​.marpetgeo​.2003​.11​.019​.
/9781444306408​.ch4​. Saylor, J.E., Knowles, J.N., Horton, B.K., Nie, J., and Mora, A., 2013, Mixing of source populations
Mora, A.R., Horton, B.K., Mesa, A., Rubiano, J., Ketcham, R.A., Parra, M., Blanco, V., Garcia, recorded in detrital zircon U-Pb age spectra of modern river sands: Journal of Geology,
D., and Stockli, D.F., 2010b, Migration of Cenozoic deformation in the Eastern Cordillera of v.  121, p.  17–33, doi:​10​.1086​/668683​.
Colombia interpreted from fission track results and structural relationships: Implications for Sheldon, N.D., and Tabor, N.J., 2009, Quantitative paleoenvironmental and paleoclimatic recon-
petroleum systems: American Association of Petroleum Geologists Bulletin, v. 94, p. 1543– struction using paleosols: Earth-Science Reviews, v. 95, p. 1–52, doi:​10​.1016​/j​.earscirev​.2009​
1580, doi:​10​.1306​/01051009111​. .03​.004​.
Mora, A.R., Reyes-Harker, A., Rodriguez, G., Tesón, E., Ramírez-Arias, J.C., Parra, M., ­Caballero, Sheldon, N.D., Retallack, G.J., and Tanaka, S., 2002, Geochemical climofunctions from North
V., Mora, J.P., Quintero, I., Valencia, V.A., Ibanez-Mejia, M., Horton, B.K., and Stockli, American soils and application to paleosols across the Eocene-Oligocene boundary in Ore-
D.F., 2013, Inversion tectonics under increasing rates of shortening and sedimentation: gon: Journal of Geology, v. 110, p. 687–696, doi:​10​.1086​/342865​.
Ceno­zoic example from the Eastern Cordillera of Colombia, in Nemcok, M., et al., eds., Shephard, G.E., Müller, R.D., Liu, L., and Gurnis, M., 2010, Miocene drainage reversal of the
Thick-Skin-Dominated Orogens: From Initial Inversion to Full Accretion: Geological Society Amazon River driven by plate-mantle interaction: Nature Geoscience, v. 3, p. 870–875, doi:​
of London Special Publication 377, p. 411–442, doi:​10​.1144​/SP377​.6​. 10​.1038​/ngeo1017​.
Nordt, L.C., and Driese, S.D., 2010, New weathering index improves paleorainfall estimates from Silva, A., Mora, A., Caballero, V., Rodriguez, G., Ruiz, C., Moreno, N., Parra, M., Ramirez-Arias,
Vertisols: Geology, v. 38, p. 407–410, doi:​10​.1130​/G30689​.1​. J.C., Ibanez, M., and Quintero, I., 2013, Basin compartmentalization and drainage evolu-
Ochoa, D., Hoorn, C., Jaramillo, C., Bayona, G., Parra, M., and De la Parra, F., 2012, The final tion during rift inversion: Evidence from the Eastern Cordillera of Colombia, in Nemčok, M.,
phase of tropical lowland conditions in the axial zone of the Eastern Cordillera of Colombia: et al., eds., Thick-Skin-Dominated Orogens: From Initial Inversion to Full Accretion: Geologi-
Evidence from three palynological records: Journal of South American Earth Sciences, v. 39, cal Society of London Special Publication 377, p. 369–409, doi:​10​.1144​/SP377​.15​.
p.  157–169, doi:​10​.1016​/j​.jsames​.2012​.04​.010​. Spikings, R.A., Winkler, W., Seward, D., and Handler, R., 2001, Along-strike variations in the
Parra, M., Mora, A., Sobel, E.R., Strecker, M.R., and González, R., 2009, Episodic orogenic front thermal and tectonic response of the continental Ecuadorian Andes to the collision with
migration in the northern Andes: Constraints from low-temperature thermochronology in heterogeneous oceanic crust: Earth and Planetary Science Letters, v. 186, p. 57–73, doi:​10​
the Eastern Cordillera, Colombia: Tectonics, v. 28, TC4004, doi:​10​.1029​/2008TC002423​. .1016​/S0012​-821X​(01)00225​-4​.

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1255
Research Paper

Trenkamp, R., Kellogg, J.N., Freymueller, J.T., and Mora, H.P., 2002, Wide plate margin defor- Velandia, F., Acosta, J., Terraza, R., and Villegas, H., 2005, The current tectonic motion of the
mation, southern Central America and northwestern South America, CASA GPS observa- Northern Andes along the Algeciras fault system in SW Colombia: Tectonophysics, v. 399,
tions: Journal of South American Earth Sciences, v. 15, p. 157–171, doi:​10​.1016​/S0895​-9811​ p.  313–329, doi:​10​.1016​/j​.tecto​.2004​.12​.028​.
(02)00018​-4​. Villagómez, D., Spikings, R., Magna, T., Kammer, A., Winkler, W., and Beltrán, A., 2011, Geo-
van der Wiel, A.M., 1991, Uplift and volcanism of the SE Colombian Andes in relation to sedi- chronology, geochemistry and tectonic evolution of the Western and Central Cordilleras of
mentation in the upper Magdalena Valley [Ph.D. thesis]: Wageningen, Netherlands, Agricul- Colombia: Lithos, v.  125, p.  875–896, doi:​10​.1016​/j​.lithos​.2011​.05​.003​.
tural University of Wagenigen, 208 p. Wellman, S.S., 1970, Stratigraphy and petrology of the nonmarine Honda Group (Miocene),
van der Wiel, A.M., and van den Bergh, G.D., 1992, Uplift, subsidence, and volcanism in the ­Upper Magdalena Valley, Colombia: Geological Society of America Bulletin, v. 81, p. 2353–
southern Neiva Basin, Colombia, Part 1: Influence on fluvial deposition in the Miocene 2374, doi:​10​.1130​/0016​-7606​(1970)81​[2353:​SAPOTN]2​.0​.CO;2​.
Honda Formation: Journal of South American Earth Sciences, v. 5, p. 153–173, doi:​10​.1016​ Wesselingh, F.P., Hoorn, M.C., Guerrero, J., Räsänen, M.E., Romero Pittman, L., and Salo, J.,
/0895​-9811​(92)90036​-X​. 2006, The stratigraphy and regional structure of Miocene deposits in western Amazonia
van der Wiel, A.M., van den Bergh, G.D., and Hebeda, E.H., 1992, Uplift, subsidence, and vol­ (Peru, Colombia and Brazil), with implications for late Neogene landscape evolution: Scripta
canism in the southern Neiva Basin, Colombia, Part 2: Influence on fluvial deposition in the Geologica, v. 133, p. 291–322.
Miocene Gigante Formation: Journal of South American Earth Sciences, v. 5, p. 175–196, doi:​ Wolaver, B.D., Coogan, J.C., Horton, B.K., Suarez Bermudez, L., Sun, A.Y., Wawrzyniec, T.F., Zhang,
10​.1016​/0895​-9811​(92)90037​-Y​. T., Shanahan, T.M., Dunlap, D.B., Costley, R.A., and de la Rocha, L., 2015, Structural and hydro-
Van Houten, F.B., 1976, Late Cenozoic volcaniclastic deposits, Andean foredeep, Colombia: geologic evolution of the Putumayo Basin and adjacent fold-thrust belt, Colombia: American
Geological Society of America Bulletin, v. 87, p. 481–495, doi:​10​.1130​/0016​-7606​(1976)87​ Association of Petroleum Geologists Bulletin, v. 99, p. 1893–1927, doi:​10​.1306​/05121514186​.
<481​. Zachos, J., Pagani, M., Sloan, L.C., Thomas, E., and Billups, K., 2001, Trends, rhythms, and aber­
Van Houten, F.B., and Travis, R.B., 1968, Cenozoic deposits, Upper Magdalena Valley, Colombia: rations in global climate 65 Ma to present: Science, v. 292, p. 686–93, doi:​10​.1126​/science​
American Association of Petroleum Geologists Bulletin, v. 52, p. 675–702. .1059412​.

GEOSPHERE  |  Volume 12  |  Number 4 Anderson et al.  |  Topographic growth and basement uplift in the Colombian Andes 1256

You might also like