You are on page 1of 26

EAGE

Basin Research (2011) 23, 377402, doi: 10.1111/j.1365-2117.2010.00493.x

Evaluating foreland basin partitioning in the northern


Andes using Cenozoic fill of the Floresta basin,
Eastern Cordillera, Colombia
Joel E. Saylor, n Brian K. Horton, n, w Junsheng Nie, n Jaime Corredor,z and Andres Moraz
n
Department of Geological Sciences, Jackson School of Geosciences, University of Texas at Austin,TX, USA
wInstitute for Geophysics, Jackson School of Geosciences, University of Texas at Austin, Austin,TX, USA
zEcopetrol, Instituto Colombiano del Petroleo, Bucaramanga, Colombia

Ab st ra c t
This paper addresses foreland basin fragmentation through integrated detrital zircon U^Pb
geochronology, sandstone petrography, facies analysis and palaeocurrent measurements from a
Mesozoic^Cenozoic clastic succession preserved in the northern Andean retroarc fold-thrust belt.
Situated along the axis of the Eastern Cordillera of Colombia, the Floresta basin rst received
sediment from the eastern craton (Guyana shield) in the Cretaceous^ early Palaeocene and then from
the western magmatic arc (Central Cordillera) starting in the mid-Palaeocene.The upper-crustal
magmatic arc was replaced by a metamorphic basement source in the middle Eocene.This, in turn,
was replaced by an upper-crustal fold-thrust belt source in the late Eocene which persisted until
Oligocene truncation of the Cenozoic section by the eastward advancing thrust front. Sedimentary
facies analysis indicates minimal changes in depositional environments from shallow marine to low-
gradient uvial and estuarine deposits.These same environments are recorded in coeval strata across
the Eastern Cordillera.Throughout the Palaeogene, palaeocurrent and sediment provenance data
point to a uniform western or southwestern sediment source.These data show that the Floresta basin
existed as part of a laterally extensive, unbroken foreland basin connected with the proximal western
(MagdalenaValley) basin from mid-Paleocene to late Eocene time when it was isolated by uplift of the
western ank of the Eastern Cordillera.The Floresta basin was also connected with the distal eastern
(Llanos) basin from the Cretaceous until its late Oligocene truncation by the advancing thrust front.

INTRODUCTION In broken forelands, where crustal- scale reverse faults


partition basin ll, the predictive nature of conventional
Foreland basins are elongate exural troughs produced by foreland basin models may break down (Jordan, 1995). Nu-
orogenic loading of continental lithosphere (Price, 1973;
merous studies have indicated that (a) pre- existing struc-
Dickinson, 1974). Most models for the evolution of fore-
tural anisotropies, (b) inherited stratigraphic frameworks
land basin systems focus on thin- skinned fold-thrust
and (c) subducted slab congurations can inuence the
belts with ramp- at structural geometries and kinematic
style of contractional deformation by incorporating large,
histories involving a systematic cratonward progression of
competent basement blocks into the deforming belt along
deformation. These models envision foreland basins as
high-angle reverse faults in ways not envisioned by critical
stratigraphic wedges propagating with, and incorporated
wedge theory (e.g. Bird, 1984; Boyer, 1995; Marshak et al.,
into, linearly extensive thrust belts which generally behave 2000; Mazzoli et al., 2000; Mora et al., 2008, 2009). These
as critical wedges (Beaumont, 1981; Jordan, 1981; Flemings
variables control the locus of deformation and sedimenta-
& Jordan, 1989; DeCelles & Giles, 1996). The success of
tion by facilitating non- systematic advances of the defor-
these foreland basin models at explaining stratal geome-
mation front (Lacombe & Mouthereau, 2002; Barbeau,
tries, facies distributions, sedimentary stacking patterns
2003; Scisciani & Montefalcone, 2006; Mortimer et al.,
and subsidence histories (Heller et al., 1988; Crampton &
2007; Carrapa et al., 2008).
Allen, 1995; Burbank et al., 1996; Sinclair, 1997; DeCelles &
In terms of basin evolution, a key dierence between
Horton, 2003) has ensured their continued usage.
conventional and broken foreland basins is whether the
foreland is laterally continuous or internally dissected by
local uplifts (Jordan & Allmendinger, 1986; Dickinson et
Correspondence: Joel E. Saylor, Department of Geological
Sciences, Jackson School of Geosciences, University of Texas at al., 1988; Bayona et al., 2008; Smith et al., 2008). Therefore,
Austin, Austin,TX 78712, USA. E-mail: jsaylor@mail.utexas.edu determining whether conventional foreland basin models

r 2010 The Authors


Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 377
J. E. Saylor et al.

are applicable to a specic orogen is important for deter- Eastern Cordillera by the intermontane Magdalena Valley, a
mining the kinematic history of the fold-thrust belt and modern hinterland basin (Horton, in press). East of the East-
the style of deformation in the hinterland.This is particu- ern Cordillera, the Llanos foreland basin overlies Precam-
larly true for orogens with multi-phase histories in which a brian cratonic rocks of the Guyana shield.
younger, non- conventional deformation stage overprints The Eastern Cordillera consists of Phanerozoic strata,
an older conventional stage (e.g. Jordan et al., 2001; Carrapa with local exposures of Proterozoic-lower Palaeozoic base-
et al., 2008). ment (Cardona et al., 2010) (Fig. 1), and has undergone
In few places are the eects of inherited structural and multiple episodes of orogenesis. It is currently expressed
stratigraphic geometries more evident than the Eastern as an asymmetric, bivergent orogen with Mesozoic exten-
Cordillera of Colombia, an isolated 200-km-wide range sional structures overprinted by thin- and thick- skinned
composing the frontal zone of the northern Andean oro - Cenozoic contractional structures aecting basement and
genic belt (Fig. 1). Here, Nazca-South America conver- overlying sedimentary cover (Colletta et al., 1990; Dengo &
gence has driven Cenozoic shortening which overprinted Covey,1993; Cooper etal.,1995; Mora etal., 2006, 2008). It is
Mesozoic extension (Etayo -Serna et al., 1983; Dengo & also the locus of Mesozoic extension-related and Cenozoic
Covey, 1993; Cooper et al., 1995; Kammer & Sanchez, shortening-related sedimentation (Kammer & Sanchez,
2006; Mora et al., 2006, 2008; Sarmiento -Rojas et al., 2006; Bayona et al., 2008; Mora et al., 2009). Estimates for
2006), resulting in fault and basin histories that may de- the onset of shortening in the Eastern Cordillera range
part from typical models of thin- skinned structures and from mid-Cretaceous to Oligocene time (Dengo & Covey,
exural basins (Bayona et al., 2008; Parra et al., 2009a). 1993; Cooper et al., 1995; Wijninga, 1996a, b; Villamil, 1999;
Tectonic reconstructions for the Eastern Cordillera dif- Gregory-Wodzicki, 2000; Cediel et al., 2003; Corredor,
fer on whether the present Magdalena Valley and Llanos 2003; Gomez et al., 2003, 2005a; Bayona et al., 2008; Parra
basins (Fig. 1a) were once contiguous and on the timing of et al., 2009a; Horton et al., 2010a, b; Nie et al., 2010).
potential structural partitioning. Bayona et al. (2008) pro -
pose that the initial Andean foreland basin was disrupted
by multiple localized uplifts throughout the Eastern Cor- SEDIMENTOLOGY AND STRATIGRAPHY
dillera starting in the Maastrichtian^ early Palaeocene. In
this scenario, multiple intermontane basins (Marocco et Methods
al., 1995; Horton, 2005) evolve independently, separated In contrast to conventional foreland basins, facies patterns
by 1^3 km topographic barriers. In contrast, others such in broken foreland basins exhibit (1) rapid lateral changes
as Gomez et al. (2005a) and Parra et al. (2009a) envision dis- (Jordan et al., 2001; Barbeau, 2003) and (2) abrupt intro -
ruption of an initially contiguous Andean foreland in the duction of new, often proximal, sources (Davila & Astini,
middle^late Eocene and complete segmentation of the 2007; Carrapa et al., 2008). Depositional environments
western (Magdalena Valley) and eastern (Llanos) basins in within subbasins of broken forelands are controlled by lo -
the early Oligocene.Yet another model (Villamil, 1999) ad- cal tectonics and climate resulting in lithofacies and litho -
vocates initial disruption of the foreland in the late Oligo - facies stacking patterns that cannot be correlated between
cene and complete segmentation in the middle Miocene. subbasins (Dickinson et al., 1988; Strecker et al., 2007). Fi-
This paper presents integrated detrital zircon U^Pb nally, broken foreland basins tend to have more rapid and
geochronology, sandstone petrography, facies analysis and spatially variable subsidence rates because they lack the
palaeocurrent measurements from the Cretaceous^Oli- long-wavelength exural support provided by unbroken
gocene Floresta basin. These data allow us to evaluate lithosphere and are aected by multiple nonuniform
whether the early Andean basin of Colombia was a broken thrust loads (Dickinson et al., 1988; Jordan, 1995; Jordan
foreland or more conventional contiguous foreland basin et al., 2001). To evaluate the degree of facies variability,
by addressing the sedimentary record of basin partition- enable comparison of depositional environments between
ing in the axial Eastern Cordillera (Fig. 1). possible subbasins and determine the long-term sedimen-
tation rate, we measured four sections totalling 4.5 km
in thickness through the Late Cretaceous^Oligocene suc-
GEOLOGICAL SETTING cession (Fig. 2). Sections were measured at 10- cm- scale
using a Jacob sta and measuring tape. Correlations and
Colombia is marked by four physiographic and tectonic pro- formation identications are based on map- scale litholo -
vinces (Fig.1a).The Western Cordillera is composed primar- gic changes.
ily of mac igneous and pelagic sedimentary rocks and is
separated from the Central Cordillera by the Romeral suture
Results
(McCourt etal.,1984; Aspden etal.,1987).The Central Cordil-
lera is dominated by Mesozoic^Cenozoic plutons intruded Five Cenozoic formations measured and described in the
into Proterozoic-lower Palaeozoic basement and Phanero- Floresta basin (Figs 1b and 2) include, from oldest to
zoic sedimentary rocks (Alvarez, 1983; Etayo-Serna et al., youngest, the Guaduas, Socha Sandstone, Socha Mud-
1983; Aspden & McCourt, 1986; Vinasco et al., 2006; Villago - stone, Picacho and Concentracion formations (Rodriguez
mez, 2010). The Central Cordillera is separated from the & Solano, 2000; Ulloa et al., 2001b; Pardo -Tujillo, 2004).To

r 2010 The Authors


378 Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Cenozoic partitioning of the Floresta basin, Colombia

(a) (b)

Fig. 1. (a) Generalized tectonic map of northwestern South America (modied from Horton et al., 2010b) showing the location of the
Floresta basin (red rectangle) relative to basement exposures in the northern Andes. Exposed basement massifs (light red shading)
contrast with now-buried basement (light yellow shading) of the MagdalenaValley (Gomez etal., 2005b). (b) Geologic map of the Floresta
basin showing the location of measured sections (Fig. 2) and principle structures identied in our eld mapping. CC, Central
Cordillera; EC, Eastern Cordillera; LL, Llanos basin; WC,Western Cordillera.

compare provenance data from pre- orogenic strata to syn- sandstone to granule conglomerate beds (Fig. 3a and c).
orogenic strata, we also measured and described the upper Bedsets include both lenticular and tabular geometries.
Guadalupe Group (Late Cretaceous) and described theTi- Where lenticular, bedsets are tens of m thick, more than
basosa Formation (Early Cretaceous) in the hanging wall 500 m wide and are surrounded by ne-grained deposits
of Soapaga thrust (Fig. 1b). (lithofacies associations F2 or F3).Tabular bedsets are tens
Denitions of13 lithofacies associations and three deposi- of m thick and can be traced laterally for multiple km. Len-
tional environments are based on lithology, texture, grain ticular sandstone beds are 0.2^1m thick 1^5 m wide with
size, bedding geometry and thickness, stacking patterns and erosive bases and tops typically truncated by the overlying
sedimentary structures (Figs 3^5). Lithofacies codes modi- unit. These beds are either ungraded or upward coarsen-
ed from Miall (1978) and DeCelles et al. (1991) are described ing.Tabular sandstone beds are 0.5^1m thick. Conglomer-
in Table 1. Unless otherwise indicated, all deposits are later- ates are clast supported, typically 0.05^0.1m thick, o10 m
ally continuous for hundreds of m to several km. wide and upward ning. Lateral accretion beds 0.5^1m
thick are present, though rare. Sedimentary structures in-
Fluvial facies associations clude trough and planar cross- stratication.

Description
F2: Sr, St, Sh, Socr, Ml. Association F2 consists of upward
F1: Sr, St, Sh, Sm, Gcm, Gcmi. Lithofacies association F1 coarsening and thickening sequences composed internally
consists of bedsets of amalgamated medium-grained of upward ning beds or bedsets (Fig. 3b). Beds are tabular

r 2010 The Authors


Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 379
J. E. Saylor et al.

Fig. 2. Measured stratigraphic sections of Late Cretaceous^Oligocene basin ll depicting sedimentary lithofacies, palaeocurrent data,
sample stratigraphic levels for sandstone point counts and detrital zircon U^Pb analyses and interpreted depositional environments.
Grain sizes: C, clay; S, silt; SS, sandstone (very ne to very coarse); Cgm, conglomerate (granule to boulder).

r 2010 The Authors


380 Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Cenozoic partitioning of the Floresta basin, Colombia

(a)
F1

(b)
F2 (d)
F3

(c)
F1
~2 m

Fig. 3. Photographs of uvial lithofacies associations F1^F3. (a) Amalgamated channelized sandstones and conglomerates (F1) of the
Picacho Formation. Resistant subhorizontal strata dene four stratigraphic packages composing the main cli face ( 80 m high). (b)
Upward coarsening and thickening sets of upward ning beds (F2) in steeply dipping strata of the lower Picacho Formation. Arrow
shows the direction of stratigraphic younging. Long inverted triangle shows upward coarsening bedsets and short upright triangles
indicate upward ning beds. (c) Large scale trough cross- stratied granule conglomerate (F1) in the upper Picacho Formation. Person
for scale. d) Massive, mottled siltstone and claystone (F3) in steeply dipping strata of the uppermost Socha Group.

with sharp or erosive bases. Grain size varies from ne mottled, with colours ranging from green^grey to mar-
sand to silt. Sedimentary lithofacies within each upward oon, and a common vertical fabric. Blocks with1^5 cm pla-
ning bedset (0.2^1m thick) are, from base to top, climb- nar faces displaying slickenlines and aligned clay minerals
ing ripple-, horizontal-, planar- and symmetrical ripple are present in some proles. Bioturbation and mottling
cross- stratied sandstones and laminated mudstones. often extend into underlying beds.
Stacked deposits are up to 7 m thick. Thinner sandstone
beds (o0.1m) featuring planar cross- stratication are ar-
Interpretation
ranged in intervals to 4 m thick.
Lithofacies associations F1, F2 and F3 are identied as
having a uvial origin based on the evidence of subaerial
F3: SMx
deposition (e.g. soil development), multistory channel
This lithofacies association consists of laterally extensive forms lled with cross-bedded sandstone and a lack of
massive sandstone and mudstone (Fig. 3d). Deposits are marine trace fossils.

r 2010 The Authors


Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 381
J. E. Saylor et al.

Lenticular sandstone beds of F1 represent deposits of large width/height ratios (440) are also present. Sedi-
migrating or avulsing channels. Tabular sandstone beds mentary structures include symmetrical ripple-, planar-
or thin, laterally discontinuous conglomerate lenses are and bidirectional (herringbone) cross- stratication and
interpreted as products of migrating bedload sheets or aser bedding. Plant fragments are common on bedding
bars (Lunt et al., 2004). Where large- scale lenticular bed- planes.Where present, mudstones are typically laminated.
sets are encased in ner grained deposits (F2, F3), we in- Bioturbation is rare, and laterally continuous oyster beds
terpret F1 to be channel lls (Mohrig et al., 2000; Jones & are present, although rare. Soft sediment deformation, in-
Hajek, 2007) within an anastomosing river system (e.g. cluding load casts and slump folding, is present in the
Smith & Smith, 1980; Smith, 1986). Although we are un- middle to upper portions of these sequences.
able to demonstrate multiple concurrently active channels
(Makaske, 2001), an anastomosing uvial setting is
D2: Mm, Mh, Ml, Socr, Sr. Lithofacies association D2 is
suggested by a low ratio of channel- ll to overbank
composed of tabular, mm to cm thick beds of laminated
deposits, abundant crevasse splay and levee deposits and
mudstone stacked in 25^100 m sequences (Fig. 4c). Subor-
an absence of lateral accretion surfaces in channel- ll
dinate massive mudstone and thin sandstone lenses and
deposits (Kraus & Gwinn, 1997). Within this setting,
layers are present. Sequences are either upward coarsen-
pedogenesis of overbank deposits (F2 and F3, see below)
ing or show no grain size trends. Sedimentary structures
attests to long periods of channel stability. In contrast,
include symmetrical ripple- and planar cross- stratica-
where F1 occurs as laterally extensive tabular bedsets of
tion. Plant fragments on bedding planes are common.
amalgamated channel deposits, we interpret it to repre-
Bioturbation is rare and limited to horizontal burrows of
sent a low-gradient braided river system (Cant & Walker,
the Cruziana ichnofacies.
1978; Bristow & Best, 1993). High lateral channel mobility
is indicated by lateral accretion surfaces and the lack of
overbank deposits. D3: Ml, Mh. This lithofacies association is marked by 25^
Association F2 is attributed to proximal overbank de- 50 m stacks of mm to cm thick tabular laminated mudstone
position (e.g. Johnson & Pierce, 1990). Upward coarsen- beds. Unlike lithofacies association D2, sandstone layers,
ing and thickening sequences composed of medium- sedimentary structures and ichnofossils are rare.
thick (40.2 m), upward ning beds are interpreted as
proximal crevasse splay deposits (Collinson, 1996; Kraus D4: Sm, Sh, St, Ml, Mh. D4 is distinguished by 10^50 m
& Wells, 1999). The consistent stacking pattern of sedi- thick, upward thickening and coarsening sequences
mentary structures within upward ning beds repre- composed of 0.05^2 m tabular beds (Fig. 4c). Beds coarsen
sents the result of waning ow with initial rapid and thicken upward from laminated mudstone to symme-
suspension fallout followed by upper- and then lower- trical ripple cross- stratied and nally to horizontally
ow regime conditions and nally standing water condi- laminated sandstones. The upper levels of the sequences
tions. Thinner (o0.1 m) beds with planar cross- strati - have extensive Cruziana ichnofacies bioturbation and shell
cation are interpreted as either distal crevasse splay or fragments.
levee deposits (Collinson, 1996; Kraus & Wells, 1999;
Bridge, 2003).
Interpretation
F3 is interpreted as oodplain palaeosol (gleyed verti-
sol or oxisol of Mack et al., 1993). Clay-lined blocks are in- Marine trace fossils and bidirectional cross- stratication
terpreted as argillaceous peds. Vertical mottling is indicate marine conditions for the deltaic lithofacies asso -
interpreted as anomalous redox conditions surrounding ciations. The 50^150 m upward thickening and coarsening
roots or burrows. Its association with F1 and F2 supports sequences featuring traction- current sedimentary struc-
the interpretation of an anastomosing river setting. tures within tabular beds suggests a deltaic environment.
Channels were stable enough to allow palaeosol develop - Lithofacies association D1represents deposits of prograd-
ment in adjacent oodplains before avulsion (Kraus, ing distal to medial delta mouthbars and channels. Her-
1999; Makaske, 2001). ringbone cross- stratication and aser bedding coupled
with planar, climbing-ripple and symmetrical ripple
cross- stratication indicates deposition above fair-weath-
Deltaic facies associations er wave base on a mixed uvial- and tide-dominated delta
front (Coleman & Wright, 1975; Bhattacharya & Walker,
Description
1992; Boyd et al., 1992; Orton & Reading, 1993; Reading &
D1: Socr, Sr, Shb, Si, Scr, Ml. Lithofacies association D1 Collinson, 1996).
consists of 0.01^0.1m beds stacked into 50^150 m thick, Association D2 represents the prodelta deposits (Cole-
upward coarsening sequences (Fig. 4a and b). Individual man & Wright, 1975; Bhattacharya & Walker, 1992; Orton &
beds are laterally continuous on an outcrop scale and up- Reading, 1993; Reading & Collinson, 1996; Mutti et al.,
ward coarsening sequences can be traced laterally for sev- 2003). The presence of rare and thin symmetrical- and
eral km. Beds are largely tabular and composed of asymmetrical ripple cross- stratied sandstones indicates
sandstone with secondary mudstone. Rare channels with weak currents and a location below fairweather wave base.

r 2010 The Authors


382 Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Cenozoic partitioning of the Floresta basin, Colombia

(a)
D1

(b) (c)

D1 D4
Oolitic Ironstone

D2

Fig. 4. Photographs of deltaic lithofacies associations D1. (a) Tabular interbedded sandstone and mudstone beds (D1) of the middle
Guaduas Formation. (b) Herringbone and climbing ripple cross- stratication (D1) in the middle Guaduas Formation. (c) Stratigraphic
transition from D2 to D4 lithofacies association in the lower Concentracion Formation. Cli face is 30 m high and capping oolitic
ironstone is 3 m thick.

The lack of hummocky cross- stratication and presence Coastal facies associations
of Cruziana ichnofacies are consistent with a low- energy
Description
environment.
D3 is closely associated with D2 but the lack of a sandy C1: Gmm, Sh, Sr, Socr, St, Ml. Lithofacies association
component or traction- current structures indicates low- C1 is composed of 0.2^1m sandstone beds interbedded
energy, low-accumulation rate processes below fair- with 0.05^0.1m thick beds (Fig. 5a). Beds are laterally
weather wave base. We therefore interpret D3 as oshore, continuous for 100 s of metres. This lithofacies associa-
deepwater shelf deposits. tion lacks any clear grain- size trends although sandstones
The thick, massive, tabular, sandstones of lithofacies become thicker and mudstones thinner upsection.
association D4 indicate unconned deposition. D4 may Sedimentary structures include trough, planar and
represent either lower^middle shoreface or proximal symmetrical ripple cross- stratication, horizontal
mouthbar deposits on the basis of their tabular bedding laminations and vertical burrows of the Skolithos ichnofa-
geometries, upward coarsening grain size trends and pre- cies. Floating granules are common in the sandstone
sence of horizontal burrows (Bhattacharya & Walker,1992). beds.
We favour interpretation of D4 as proximal mouthbar de-
posits due to the lack of storm deposits characteristic of C2: Gct, St, Sh, Sr. C2 channels 4^5 m thick have width/
non-deltaic coasts (Bhattacharya & Walker, 1992). The ex- length ratios of 20^40 and are found at the top of upward
tensive bioturbation at the top of the 10^50 m upward coarsening sequences, surrounded by clay/siltstone de-
thickening sequences indicates delta mouthbar abandon- posits. Basal surfaces are erosional. Rare internal struc-
ment due to delta lobe avulsion or marine transgression tures include lateral accretion surfaces dened by
(Reading & Collinson, 1996). interbedded sandstone and mudstone.

r 2010 The Authors


Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 383
J. E. Saylor et al.

(a)
C1 (b)
C3

(c)
C4 (d)
C5

(e)
C4
C5
C6

Fig. 5. Photographs of coastal lithofacies associations C1^C6. (a) Meter-scale horizontally bedded sandstones interbedded with 10^
30 cm thick laminated siltstone (C1) which mark the transition between the Picacho and Concentracion formations. (b) Large- scale
foresets in gently dipping, planar cross- stratied sandstone (C3) of the upper Concentracion Formation. (c) Laminated mudstones (C4)
showing the lateral continuity of even, thinly laminated beds of the middle Concentracion Formation. Quaternary alluvium caps the
5 m cli face. (d) Laterally continuous sandstone beds in the upper Concentracion Formation featuring horizontally laminated and
ripple cross- stratied beds (C5) interpreted as sand at or tidal bar deposits. (e) Stratigraphic transition from C6 to overlying C4 and C5
lithofacies association of the upper Concentracion Formation.

C3: Sh, Sf. Association C3 consists of 1^20 m stacks of planes in both the Ml and Mh (Fig. 5c and e). Local accumu-
horizontally bedded ne-grained sandstones interbedded lations of oolitic ironstone are up to 3 m thick (Fig. 4c).
with 1^2 m thick sandstones featuring large- scale planar
cross- stratication (Fig. 5b). C3 occasionally shows an up- C5: Sh, Sr, Sm. C5 includes interbedded1^3 m tabular sand-
ward coarsening trend. stones and secondary (o50 cm) laminated tabular mudstones
and caps lithofacies association C4 (Fig. 5d and e). Basal sur-
C4: Ml, Mh. C4 is characterized by 5^10 m laminated mud- faces are either gradational or abrupt and top surfaces are
stone beds typically capped by C6 beds. Iron nodules (1^10 cm marked by an abrupt decrease in grain size and increase in or-
diameter) and plant fragments are common along bedding ganic content. Sedimentary structures include horizontal la-

r 2010 The Authors


384 Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Cenozoic partitioning of the Floresta basin, Colombia

Table1. Lithofacies codes, descriptions, and interpretations

Lithofacies Description Interpretation

Ml Laminated mudstone Suspension settling in shallow standing water. Plant material


commonly denes laminations
Mh Horizontally bedded mudstone Suspension settling in relatively deep water
Mm Massive mudstone Mud deposit disrupted by post-depositional bioturbation
SMx Highly weathered massive sandstone or Sand or mud deposit aected by pedogenic alteration (palaeosol)
mudstone
Si Interbedded sand ripples and mud laminae Deposition by intermittent ripple migration and suspension settling
Sf Sandstone with large planar foreset Deposition on the foresets of migrating subaerial dunes
stratication
Sr Asymmetrical 2D cross- stratied sandstone Deposition of 2D dunes by unidirectional ow
Socr Symmetrical 2D cross- stratied sandstone Deposition of 2D dunes by oscillatory ow
Shb Herringbone cross- stratied sandstone Deposition of ripples by alternating ow
St Trough cross- stratied sandstone Deposition by migrating subaqueous 3D dunes
Scr Climbing ripple cross- stratied sandstone Rapid deposition of sand ripples
Sm Massive sandstone Rapid deposition of sand without stable bedform development.
Alternatively, post-depositional bioturbation of sand deposit
Sh Horizontally laminated sandstone Deposition during upper ow regime plane-bed conditions
Gmm Massive, matrix- supported conglomerate Deposition within a buoyancy-modied, cohesive, sand-matrix
sheet
Gct Trough cross- stratied, clast supported Deposition by traction currents in migrating subaqueous 3D dunes
conglomerate
Gcm Massive, clast- supported conglomerate Deposition by traction currents in a buoyancy- or dispersive-
pressure-modied sheet
Gcmi Massive, clast- supported, imbricated Deposition by traction currents in migrating subaqueous bar
conglomerate

minations and rare planar cross-stratication. Bioturbated absence of closely associated subaerial deposits. C3 sand-
and massive sandstone beds are common. stones featuring horizontal bedding and large- scale planar
cross-bedding are interpreted as swash zone and eolian
C6: Sr, St, Ml, Mm. This lithofacies association consists of back-beach dune deposits, respectively (Reading & Collin-
upward coarsening 1^3 m mudstone bodies capped by son, 1996). C4 laminated clay/siltstones are coupled with
upward coarsening and thickening 1^3 m sandstone plant fragments and rare oolitic ironstones, suggestive of
bodies (Fig. 5e). Mudstone bodies are tabular. Sandstone quietwater setting with occasional wave agitation in close
bodies are tabular to broadly lenticular. Plant fragments proximity to nonmarine plant material (Kimberley, 1980;
are common on bedding planes of both. Sandstone Carr et al., 2003). Close association of C4 with other coastal
bodies often contain oating granule grains and lenses. lithofacies suggests deposition in a lagoon or partially
Sedimentary structures include planar, climbing-ripple closed estuarine setting (Dalrymple et al., 1992; Reinson,
and trough cross- stratication. Rare weakly developed 1992; Reading & Collinson, 1996; Dalrymple & Choi,
palaeosols are present in massive, siltstone intervals up to 2007).Within this context, C5 sandstone bodies are inter-
3 m thick. preted as deposits of upper ow regime sand ats or mi-
grating tidal bars; in contrast, C5 mudstone bodies
represent deposition within the low- energy centre of the
Interpretation
estuary, removed from high- energy uvial or shoreface
Lithofacies association C1exhibits sedimentary structures environments (Dalrymple et al., 1992; Ehlers & Chan,
consistent with a high- energy wave-dominated regime. 1999; Dalrymple & Choi, 2007). The coarser grain size
We infer upper shoreface deposition (Reineck et al., 1970; and abundance of traction- current sedimentary struc-
Howard & Reineck, 1981; Reading & Collinson, 1996). The tures of C6 are indicative of relatively strong unidirectional
broadly channelized geometry, close association with currents. Coupled with strong evidence of nonmarine
coastal facies, lateral accretion surfaces and repetitive oc- processes, the upward coarsening trend is attributed to
currence of mud^ sand couplets in C2 indicate deposition uvially inuenced deposition within a bayhead delta
either within a tidal inlet channel or distributary channel (Carr et al., 2003; Ascho, 2008). Within this context,
(Hoyt & Henry, 1967; Dalrymple et al., 1992; Reinson, 1992; the upward coarsening trends are interpreted as prog-
Ehlers & Chan, 1999; Carr et al., 2003).We favour a tidal in- rading channel mouthbars over the bay or lagoon oor
let channel due to (1) the presence of sandstone and mud- and broadly lenticular sand bodies are interpreted as
stone couplets in lateral accretion surfaces and (2) the distributary channels. The presence of weakly developed

r 2010 The Authors


Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 385
J. E. Saylor et al.

palaeosols indicates intermittent subaerial exposure and dm scale, which, combined with the poor outcrop expres-
pedogenesis. sion due to high mud content, precludes a denitive thick-
ness estimate.
The Socha Sandstone is 100 m thick and composed
Lithostratigraphy
of amalgamated lenticular and tabular uvial deposits of
To assess depositional variability and correlate deposi- medium-grained sandstone to granule conglomerate (F1,
tional systems in the Floresta basin with those in other Fig. 2). The base of the formation is extremely lithic rich
subbasins, we summarize the observations above into (up to 50% of total framework sand grains are unstable
chronological, formation- scale interpretations of deposi- lithic grains); however, upsection it quickly changes to a
tional environments. Where described in its type locality largely quartzose composition.The overlying Socha Mud-
near the town of Tibasosa (5.751N, 73.01W) the Lower Cre- stone ( 150 m thick) is a oodplain succession including
taceous (Hauterivian^Aptian) Tibasosa Formation is re- palaeosols (F3), levee and crevasse splay deposits (F2).
ported to be 350^574 m thick (Ayala-Calvo & Delgado - There are also 10^20 m thick, broadly lenticular bodies of
Rozo, 2004). Near the town of Beteitiva (5.91N, 72.81W) it coarse, amalgamated uvial deposits (F1). The formation
is 600 m thick. However, it is in thrust contact at both is capped by a well-developed palaeosol horizon at least
its base and top, precluding a denitive thickness assess- 30^50 m thick. The introduction of signicant levee and
ment. Contrary to previous reports and mapping, we con- crevasse splay deposits in the Socha Mudstone suggests a
clude that the moderately dipping beds (35^401) of the change upsection from a braided to anastamosing uvial
Tibasosa Formation are upright based on observed cross- deposition from the Socha Sandstone to the Socha Mud-
stratication and scour marks. The formation coarsens stone. Extensive and pervasive pedogenesis of overbank
and thickens upward from ssile grey slate with pencil deposits also attests to extended channel stability followed
cleavage and nodular iron-rich horizons (D2 and D3) to by rapid avulsion in the Socha Mudstone. The increase in
granule conglomerate (C1) in beds up to 8 m thick. It is mud content and isolation of channel deposits between
capped by a thick fossiliferous wackestone (sensu Dunham, the Socha Sandstone and Socha Mudstone may be indica-
1962) featuring largeTrigonia shells.The Upper Cretaceous tive of increased accommodation (Bristow & Best, 1993;
Guadalupe Group includes the Labor, Pinos and Tierna Miall, 1996; Currie, 1997). However, any trend of increas-
Sandstone formations (Ulloa et al., 2001a). These forma- ing accommodation is abruptly reversed at the top of the
tions total 100 m thick and compose an upward coarsen- section where the thick palaeosol horizon attests to low ac-
ing and thickening sequence of D3, D2 and D4. We cumulation rates and extended exposure and weathering.
conclude that both the Tibasosa Formation and upper Based on palynology, the Socha Sandstone ranges in age
Guadalupe Group represent progradational shallow-mar- from early to late Palaeocene (Pardo -Tujillo, 2004),
ine sequences (Fig. 2). although an entirely late Palaeocene age for the base of
The Guaduas Formation is 400 m thick and straddles the formation has also been proposed (Guerrero & Sar-
the Maastrichtian- early Palaeocene boundary (Pardo -Tu- miento, 1996). The age of the Socha Mudstone is late Pa-
jillo, 2004). It is composed of mm to cm scale beds stacked laeocene and may extend into the early Eocene (Ypresian)
into 75^200 m upward coarsening sequences (Fig. 2). The (Guerrero & Sarmiento, 1996; Pardo -Tujillo, 2004).
formation is composed primarily of organic-rich ssile The contact between the Socha Mudstone and the over-
shales and coal with secondary sandstone. Plant fragments lying middle Eocene Picacho Formation (Pardo -Tujillo,
are common on bedding planes. Bioturbation is rare at the 2004; Pulido et al., 2006) is erosional, although the time re-
base of the upward coarsening sequences, but becomes presented by the unconformity is unclear (Fig. 2). The Pi-
more prevalent in upper levels. Beds are largely tabular, cacho Formation is 200 m thick. It is composed almost
although rare channels with large width/height ratios entirely of amalgamated F1 uvial sandstones and granule
(440) are also present. Sedimentary structures include conglomerates, with subordinate ner grained oodplain
symmetrical ripple-, planar- and herringbone cross- stra- mudstone and crevasse splay deposits. Like the Socha
tication and aser bedding. Evidence of soft sediment de- Sandstone, the depositional environment is inferred to be
formation, including load casts and slump folding, is a large braided river system.
present in the mid to upper portions of upward coarsening The contact between the Picacho and overlying Con-
sequences. The uppermost Guaduas Formation includes centracion Formation is conformable and transitional
organic-rich, laminated mudstone, palaeosols and chan- (Fig. 2). Toward the top of the Picacho Formation, the u-
nelized uvial sandstones. These observations are consis- vial sandstones become less amalgamated and are sepa-
tent with lithofacies associations D1, D2, D3, F1, F2 and rated by intercalated mudstones. We place the contact
F3 and we interpret the Guaduas Formation as the result between the formations at the top of the highest thick u-
of progradational sedimentation within a mixed uvial- vial sandstone body in this transitional zone.The Concen-
and tide-dominated delta (Coleman & Wright, 1975; Bhat- tracion Formation is 1km thick and spans the middle
tacharya & Walker, 1992; Orton & Reading, 1993; Reading & Eocene (Bartonian) to late Oligocene (early Chattian) (Par-
Collinson, 1996). The uppermost Guaduas Formation was do -Tujillo, 2004; Gomez et al., 2005a; Pulido et al., 2006).
deposited in a delta top environment. The Guaduas For- The top of the Concentracion Formation is in fault contact
mation is intensely deformed by internal folding on cm to with the Cretaceous Tibasosa or Jurassic Giron forma-

r 2010 The Authors


386 Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Cenozoic partitioning of the Floresta basin, Colombia

tions. No younger units are exposed in the Floresta basin. ne-grained nature of the sediments. Palaeocurrent data
The Concentracion Formation is composed primarily of consist of measurements of 241limbs of trough cross- stra-
upward coarsening C4 and C5 sequences. Sequence bases ta (method I of DeCelles et al., 1983) at 14 sites and mea-
are composed of laminated silt/claystones which coarsen surements of the rake of the axis of 48 trough cross- strata
upward into massive or cross- stratied sandstones. The at four sites.
tops of some sequences locally have granule conglomer-
ates. There is a slight upward coarsening trend in the for- Results and interpretation
mation. The Concentracion Formation is inferred to have
been deposited in a lagoonal or partially closed estuarine Palaeocurrents were measured in the Socha Sandstone,
environment (Kimberley, 1980; Cazier et al., 1997; Villamil, Socha Mudstone, Picacho and Concentracion formations.
1999; Santos et al., 2008). Socha Group palaeocurrents are variable with both north-
ward and eastward components (Fig. 6a). However, there is
no systematic stratigraphic trend (Fig. 2) and we attribute
Summary
these results to a single uvial system with an average
The sedimentological descriptions highlight several rele- northeastward trend (Fig. 6a). This dispersal pattern
vant observations. (1) The Palaeocene^upper Oligocene changes abruptly in the Picacho Formation to a highly uni-
strata exposed in the Floresta basin are typically composed form northward orientation (Fig. 6b). Finally, in the Con-
of siltstone to ne-grained sandstone. Granule conglom- centracion Formation there is a return to a generally
erates are rare and pebble-boulder conglomerates are eastward palaeoow (Fig. 6c). However, we note that the
absent. (2) The uvial systems that deposited the Socha Concentracion is a marginal marine unit subject to nonu-
Group and Picacho Formation are inferred to be low-gra- niform ow conditions and a high degree of variability, and
dient rivers close to (local) base level based on the sedi- that measured palaeocurrent indicators may not be repre-
mentological evidence of anastomosing uvial deposition sentative of the larger depositional system.
and absence of coarse-grained material (Makaske, 2001).
Hence the basin shows a minimal shift in depositional en-
vironments from marginal marine to low-gradient uvial
settings. (3) The Palaeocene (middle Guaduas Formation)
SANDSTONE PETROGRAPHY
to late Oligocene (Concentracion Formation) succession is Methods
2 km thick. Given the lack of major unconformities, this
succession represents a remarkably slow, 50 m Myr  1 Modal sandstone composition data were collected from 47
long-term average rate of undecompacted sediment accu- standard petrographic thin sections. Thin sections were
stained for potassium and calcium feldspar and point
mulation. The results of the lithofacies analysis and in-
counted (450 counts per slide of grains larger than silt)
ferred depositional processes indicate no rapid or
using the Gazzi^Dickinson method (Ingersoll et al., 1984).
dramatic changes in depositional environment.Therefore,
Petrographic counting parameters are listed inTable 2, and
the sedimentological evidence does not support an intro -
recalculated modal data in Table S1.
duction of proximal source areas during basin evolution.

Results
PROVENANCE Quartz is present as monocrystalline, polycrystalline and
foliated polycrystalline grains, with monocrystalline
Broken forelands are typied by multiple sediment grains also present within grains of sandstone or quartzite.
sources and complex dispersal patterns (Dickinson et al.,
Lithic grains are primarily chert with minor mudstone.
1988; Davila & Astini, 2007; Carrapa et al., 2008; Smith et
Volcanic grains are present primarily as vitric or lath-work
al., 2008). In contrast, unbroken foreland basins display
grains containing plagioclase crystals.Volcaniclastic grains
simple, transverse or longitudinal sediment dispersal pat-
featuring angular rock or mineral fragments within a vitric
terns and provenance primarily from the orogenic hinter-
or lath-work matrix are also present. Accessory minerals
land. To identify source areas for the Floresta basin, we
include muscovite, zircon and hornblende. Clay is the
measured palaeocurrents throughout the basin and con-
most abundant cement and often includes recognizable
ducted petrographic modal composition analysis and U^ sub- silt sized quartz grains.
Pb geochronology of detrital zircons.
Cretaceous (Tibasosa Formation and Guadalupe
Group) sandstones consist of variably sorted, tightly
packed quartz arenites (495% quartz; Folk, 1980) (Figs
PALAEOCURRENTS 7c and S1c). Monocrystalline quartz is the dominant con-
stituent with lesser polycrystalline quartz. Metamorphic
Methods
grains are limited to polycrystalline quartz. No unstable
Sedimentary structures and outcrops suitable for measur- metamorphic grains, volcanic or sedimentary grains (in-
ing palaeocurrents are rare in the Palaeogene section of the cluding chert) are present. Unstable lithic fragments and
Floresta basin primarily due to poor exposure and the feldspars compose o3% of samples from these units.

r 2010 The Authors


Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 387
J. E. Saylor et al.

In contrast to underlying Cretaceous units, sandstones count. Polycrystalline quartz constitutes up to 21%. Other
from the Guaduas Formation consist of both quartz are- major components that appear for the rst time in the So -
nites and sublitharenites (Figs 7c and S1c). Guaduas sand- cha Group include foliated polycrystalline quartz (foliated
stones are tightly packed although they generally exhibit polycrystalline quartz), felsic volcanic lithic fragments (in-
better sorting than other Cretaceous samples. Monocrys- cluding vitric and volcaniclastic grains), chert fragments
talline quartz is the dominant component with subordi- (up to 16%) and metamorphic lithic fragments (including
nate polycrystalline quartz. schist and phyllite fragments).
Sandstones from the Socha Sandstone and Socha Mud- The Picacho Formation is composed of poorly to
stone continue the trend toward increasing lithic frag- moderately sorted, subrounded-rounded sublitharenites
ments (Figs 7c and S1c). These samples straddle the and arenites (Figs 7c and S1c). Within the formation,
sublitharenite and litharenite elds of Folk (1980). Sand- the composition shifts quickly upsection from monocrys-
stones are moderately or poorly sorted and composed of talline quartz of 64% to 95% synchronous with increased
subrounded to angular grains. Monocrystalline quartz is polycrystalline quartz (Figs 7a, b and S1a, S1b). Only
the major component, although present in proportions as the lowest sample contains other components in signi -
low as 47% (Figs 7b and S1b) of the total detrital grain cant quantity. These components include foliated poly-
crystalline quartz, felsic volcanic fragments and chert
fragments.
Sandstones of the Concentracion Formation likewise
(b) are of sublitharenite to arenite composition (Figs 7c and
S1c). Sorting and rounding decrease upsection from well
sorted, sub-rounded to rounded at the base to poorly
Concentracion Fm. sorted, angular grains at the top of the section. Polycrystal-
Mean = 124
Angular Deviation=37
line quartz and volcanic lithic grains are present through-
n= 40 out the section but show no upsection trend. However, the

Table 2. Modal petrographic point- count parameters

Symbol Description

Qm Monocrystalline quartz
Qp Polycrystalline quartz
Qpt Foliated polycrystalline quartz
(b) Qms Monocrystalline quartz grain within sandstone or
quartzite lithic grain
C Chert
Picacho Fm.
Mean = 340 S Siltstone
Angular Deviation=19 Qt Total quartz (Qm1Qp1Qpt1Qms1C1S)
n= 158 Q Total quartz (sensu Folk (1980):
Qm1Qp1Qpt1Qms)
K Potassium feldspar
P Plagioclase feldspar
F Total feldspar (P1K)
Lvm Mac volcanic grains
Lvf Felsic volcanic grains
Lvv Vitric volcanic grains
Lvl Lathwork volcanic grains
(a) Lv Total volcanic grains (Lvm1Lvf1Lvv1Lvl)
Lsh Mudstone/shale
Lc Carbonate
Socha Grp. Ls Total sedimentary lithic grains
Mean = 44 (Lsh1Lc1C1S1Qms)
Angular Deviation=98 Lph Phyllite
n= 98
Lsm Schist
Lg Gneissic/mylonitic grains
Lm Total metamorphic grains (Qpt1Lph1Lsm1Lg)
Lt Total lithic grains (Lv1Ls1Lm1Qp)
L Total non-quartzose lithic grains
(Lv1Lsh1Lc1Lph1Lsm1Lg)
Fig. 6. (a) Rose diagrams showing measured palaeocurrent Lf Total lithic grains (sensu Folk (1980): L1C1S)
orientations from (a) Socha Group, (b) Picacho Formation and (c)
Concentracion Formation. Accessory minerals include zircon, muscovite, biotite and hornblende.

r 2010 The Authors


388 Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Cenozoic partitioning of the Floresta basin, Colombia

proportion of chert fragments increase upsection (Figs 7a Lm^Lv^Ls ternary diagram, Guaduas sandstones plot
and S1a). within the recycled orogen eld (Fig. 7a) possibly suggest-
ing that orogenesis was ongoing with sediment derived
from a recycled orogenic hinterland. Previous thermochro-
Interpretation nology suggests that exhumation of the Central Cordillera
Owing to extremely high weathering rates associated with was ongoing by the latest Cretaceous (Gomez et al., 2003;
Colombias tropical climate (Johnsson et al., 1991), unstable Villagomez et al., 2008). The Guaduas Formation sand-
lithic fragments and minerals have been highly degraded, stones thus present an enigma because if the Central Cor-
leaving sandstones enriched in stable grains, particularly dillera were a sediment source in the latest Cretaceous^
quartz.The result is that all samples are rather quartz-rich early Palaeocene, the sandstone composition should incor-
(focused in upper apex of their respective ternary plots) porate more volcanic components (i.e. trend toward the Lv
and tectonically driven variation is minimal. apex in the Lm^Lv^Ls diagram), which is not observed un-
Extremely monocrystalline quartz-rich lithologies of til the upper Palaeocene Socha Sandstone (Fig. 7a).
the Lower CretaceousTibasosa Formation and Upper Cre- Compositions of Socha Group and Picacho Formation
taceous Guadalupe Group are consistent with derivation sandstones display a reversal in trend that requires expla-
from a cratonic source (Fig. 7) (Dickinson & Suczek, 1979; nation. Socha sandstones have increasing amounts of vol-
Dickinson et al., 1983; Dickinson, 1985). Sandstones from canic lithic grains (trend toward the Lt and Lv apexes in
the Maastrichtian- early Palaeocene Guaduas Formation Fig. 7a and b), a potential result of exhumation of the Cen-
similarly plot within the cratonic interior eld of the Qm^ tral Cordillera and partial dissection of its volcanic cara-
F^Lt and Qt^F^L ternary plots (Fig. 7b). However, in the pace. However, sandstones from the overlying Picacho

(a) (b)

(a) (d)

Fig. 7. Modal sandstone petrographic data plotted in ternary diagrams: (a) Lm^Lv^Ls, (b) Qm^F^Lt, (c) Q^F^Lf and (d) Qt^F^L. Black
arrows indicate upsection compositional trends. Large solid symbols depict formation averages. Shaded polygons represent 1s
uncertainty envelopes on formation averages. Fields for A, B and D from Dickinson et al., (1983) and Dickinson (1985). Fields for C from
Folk (1980).

r 2010 The Authors


Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 389
J. E. Saylor et al.

Formation contradict this trend, incorporating higher tensity of the 207Pb signal. For most analyses, the cross-
percentages of monocrystalline quartz and metamorphic over in precision of 206Pb/238U and 206Pb/207Pb ages occurs
lithic grains (closer to the Qm and Lm apexes in Fig. 7a at 0.8^1.0 Ga. Hence, the cuto between 206Pb/238U-based
and b). Continued and deeper dissection of the Central ages and 206Pb/207Pb-based ages is between 0.8 and 1 Ga
Cordilleran magmatic arc should result in compositions and was chosen to prevent articially separating zircon
increased in feldspar and volcanic lithic grains, which is populations. Analyses with 420% uncertainty, 430%
not observed. The observed compositional trend is inter- discordance (by comparison of 206Pb/238U and 206Pb/207Pb
preted as the result of eastward migration of the deforma- ages) or 45% reverse discordance are omitted from
tion front coupled with deep exhumation of metamorphic further consideration. Interpretations are based on age
basement. Finally, Concentracion Formation sandstones peaks dened by 3 or more analyses (Dickinson & Gehrels,
have greater proportions of sedimentary lithic grains 2008). This criterion minimizes the likelihood of a false
(trend toward the Lt and Ls apexes in Fig. 7a and b), indi- positive based on ages aected by Pb loss in young grains,
cating renewed introduction of upper crustal recycled se- inheritance, common Pb, or inaccurate ages for single
dimentary material likely as a result of a further eastward grains.
advance of deformation and incorporation of foreland ba-
sin strata into the thrust belt. Results
A total of 1692 new zircon ages from 17 samples are re-
ported here (Table S2). The preferred ages are shown on
DETRITAL ZIRCON U ^PB age histograms and relative age-probability diagrams
(from Ludwig, 2003). The age-probability diagrams show
GEOCHRONOLOGY
each age and its uncertainty (for measurement error only)
Methods as a normal distribution, and sum all ages from a sample
into a single curve (Fig. 8).
U^Pb geochronology was conducted on zircons separated
from 17 sandstone samples from the Floresta basin using
Interpretation
standard procedures described by Gehrels (2000) and
Gehrels et al. (2008). Analyses were conducted using laser- Interpretations of detrital zircon U^Pb age spectra
ablation, multicollector, inductively coupled-plasma mass are based on previously reported isotopic ages for the
spectrometry (LA-MC-ICPMS) at the University of Ari- Eastern, Central and Western Cordilleras and the South
zona LaserChron Center. Approximately 100 individual American craton. Sources are based both on K^Ar
zircon grains were analysed from each sample.These were and Rb^Sr methods (Goldsmith et al., 1971; Aspden
selected randomly, although metamict grains or those with et al., 1987) as well as 40Ar/39Ar and zircon U^Pb dating
obvious cracks or inclusions were avoided. In-run analysis (Dorr et al., 1995; Restrepo -Pace et al., 1997; Molina et al.,
of fragments of a large Sri Lanka zircon crystal (generally 2006; Vinasco et al., 2006; Chew et al., 2008; Horton et al.,
every fth measurement) with known age of 564  4 Ma 2010a, b; Nie et al., 2010). Based on these studies, we
(2s) was used to correct for inter- and intra- element frac- identify several age groups unique to each of the
tionation (Gehrels et al., 2008). The uncertainty resulting Cordilleras and craton. Igneous activity in the Western
from the calibration correction is generally 1^2% (2s) for Cordillera was most intense in the late Cenozoic (25^
both 206Pb/207Pb and 206Pb/238U ages. Common Pb was 10 Ma) (McCourt et al., 1984; Aspden et al., 1987). Middle
corrected using the measured 204Pb and assuming an initi- Jurrasic^Palaeogene igneous ages ( 170 to 40 Ma) are ty-
al Pb composition from Stacey & Kramers (1975) (with un- pical of the Central Cordillera (Aspden et al., 1987; Restre-
certainties of 1.0 for 206Pb/204Pb and 0.3 for 207Pb/204Pb). po -Pace, 1992; Villagomez, 2010). Only the Central
Analytical data are reported in Table S2. Details of operat- Cordillera and Santander Massif include prominent Per-
ing conditions and analytical procedures are available in mo -Triassic ages (Vinasco et al., 2006; Horton et al., 2010b;
Gehrels et al. (2008). For each analysis, errors in determin- Nie etal., 2010).The Eastern Cordillera includes Grenville-
ing 206Pb/238U and 206Pb/204Pb result in a measurement er- aged basement with secondary Cambro -Ordovician in-
ror of 1^2% (2s) in the 206Pb/238U age. The errors in trusive activity (Dorr et al., 1995; Restrepo -Pace et al., 1997;
measurement of 206Pb/207Pb and 206Pb/204Pb also result in Cordani et al., 2005; Cardona et al., 2010; Horton et al.,
1^2% (2s) uncertainty in age for grains 41.0 Ga, but 2010bl; Nie et al., 2010). Neoproterozoic zircons are pre-
are substantially larger for younger grains due to low in- sent in the Eastern Cordillera in Palaeozoic strata, possibly

Fig. 8. U^Pb ages for detrital zircons from the Floresta basin, shown as age-probability density diagrams (black lines) and age
histograms (grey shading). Samples are plotted in stratigraphic order from bottom to top. Key changes in U^Pb age distributions include
the introduction of Central Cordilleran zircons in the Socha Sandstone (D) and Eastern Cordilleran zircons in the middle Picacho
Formation (I).The number scale on the right applies for age-histograms only (scale for B^Q is as A). Note that the age histogram for the
youngest peak for G is dened by 16 grains, exceeding the maximum vertical scale. Signicant age peaks are those dened by a
minimum of three grains. CC, Central Cordillera; EC, Eastern Cordillera.

r 2010 The Authors


390 Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Cenozoic partitioning of the Floresta basin, Colombia

derived from a Neoproterozoic magmatic arc (Chew et al., craton is marked by several northwest-trending provinces
2008; Horton etal., 2010b), and Lower Cretaceous strata re- with crystallization ages ranging from 42300 to 1000 Ma
cycled from Palaeozoic units (Horton et al., 2010b). The (Teixeira et al., 1989; Tassinari & Macambira, 1999; Chew et

r 2010 The Authors


Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 391
J. E. Saylor et al.

al., 2008). Zircons of Grenville age have also been docu- of the Santander massif and western margin of the Eastern
mented in basement exposures in the Eastern Cordillera Cordillera (Dorr et al., 1995; Horton et al., 2010b). Although
(Cordani et al., 2005), implying that Grenville-aged zircons the 200^300 Ma zircons could also be derived from the
could come from either the craton to the southeast or from Central Cordillera, we consider this unlikely for two rea-
Eastern Cordilleran basement. This scenario is compli- sons. First, the 200^300 Ma zircons did not rst appear
cated by zircon recycling (e.g. Dickinson & Gehrels, with Cretaceous-age zircons which were certainly derived
2008). Most notably, both the Central and Eastern Cordil- from the Central Cordilleran magmatic arc. Second, the
lera were at least partially covered during the Mesozoic, 200^300 Ma zircons did appear along with 500^650 Ma
with craton-derived clastic sediment. zircons which must derive ultimately from Eastern Cordil-
The Lower CretaceousTibasosa Formation (Fig. 8a) has leran basement.
major peaks at 1000, 1200, 1500 and 1800 Ma. We inter- The remainder of the Picacho Formation and most of
pret this spectrum as indicating a distal cratonic source, the Concentracion Formation have similar zircon age dis-
along with, possibly, local Eastern Cordilleran basement tributions. These include o175, 200^300, 500^650 Ma
exposed during a Mesozoic rifting event (Cooper et al., and Proterozoic peaks (Fig. 8i^ o).There is a divergence be-
1995; Sarmiento -Rojas et al., 2006). Because the Gren- tween stratigraphic age and the age of the youngest signif-
ville-age zircons have multiple possible sources, we can- icant peaks at stratigraphically younger levels of the
not draw denitive conclusions based on their presence. Concentracion Formation, despite ongoing magmatism
In the Upper Cretaceous Tierna Sandstone, the Grenville in the Central Cordillera. As the strata become progres-
peak is suppressed relative to the underlyingTibasosa For- sively younger, the youngest zircon peaks remain Late Cre-
mation, but the 1300^1600 Ma peaks and the 1800 Ma peak taceous to Palaeocene in age, but are systematically older
remain (Fig. 8b).We interpret this pattern as indicating al- than the youngest peaks observed in the Socha Mudstone
most complete burial of the previously exposed Eastern and lower Picacho formations [for example, compare the
Cordilleran basement or westernmost cratonic province mean age of youngest peaks from the Palaeocene Socha
by sediments derived from the craton (Horton et al., Group (61 Ma) and upper Eocene^Oligocene Concentra-
2010b). cion Formation (173 Ma)].We attribute this contrast to the
Whereas the modal petrography and detrital zircon cessation of rst- cycle zircons derived from the Central
spectra for the Cretaceous samples point to a cratonic Cordillera and the recycling of zircons from Palaeocene
source, the petrographic data for the uppermost Cretac- and Cretaceous units in the Eastern Cordillera.
eous^lower Palaeocene Guaduas Formation samples are
less clear. However, the Guaduas Formation detrital zircon
data are nearly identical to the underlying Cretaceous BASIN SYNTHESIS
samples and reect an eastern, cratonic source. Like un-
derlying formations, the Guaduas Formation (Fig. 8c) lacks Combined sedimentary petrology, palaeocurrent and geo -
signicant peaks younger than 900 Ma. Hence, if exhuma- chronological provenance analysis of the Cretaceous^Oli-
tion of the Central Cordillera was ongoing by the latest gocene ll conrm that the Floresta basin was part of a
Cretaceous (Gomez et al., 2003; Villagomez et al., 2008), broad foreland basin by mid-Palaeocene time. For the
there is no sedimentary record of it in the Eastern Cordil- Maastrichtian-lower Palaeocene Guaduas Formation, Pa-
lera, the region presumed to represent the earliest foreland laeo - and Mesoproterozoic detrital zircon U^Pb ages sug-
basin. gest derivation from the Guyana shield, similar to the
The rst unambiguous indication of Central Cordiller- underlying Lower^Upper Cretaceous strata (Fig. 8a^ c).
an input is a 79 Ma peak for the Palaeocene Socha Sand- The rst arrival of west-derived detritus occurs in the
stone (Fig. 8d). However, in addition to this Late mid-Palaeocene when Late Cretaceous zircons (Fig. 8d),
Cretaceous peak, the Socha Sandstone also has 1000, abundant volcanic lithic fragments (Fig. 7a), and north-
1300, 1500 and 1800 Ma peaks. Because palaeocurrent di- east-directed palaeocurrents (Fig. 6a) appear in the basal
rections in the Socha Group are east directed we interpret Socha Group. Palaeocurrent and petrographic data indi-
the Proterozoic ages as erosion of Central Cordilleran se- cate a western source area dominated by volcanic rocks.
dimentary cover or metamorphic units.This same pattern The youngest zircon U^Pb ages from the Socha Group
continues upsection throughout the Socha Group to the are within error of its late Palaeocene depositional age
contact with the basal Picacho Formation (Fig. 8d^h). (Fig. 8e), suggesting direct derivation from a coeval volca-
Zircons from the middle Picacho Formation (Fig. 8i) re- nic highland with minimal sediment storage. These data
cord the introduction of a new source with igneous ages of indicate a reversal in sediment dispersal polarity (e.g. Con-
200^300 and 500^650 Ma. These suggest initial erosion of ey & Evenchick, 1994) between the early and late Palaeo -
sources with Eastern Cordilleran anities. Co - occurrence cene and point to the emergence of the western magmatic
of Late Cretaceous and Proterozoic peaks indicates con- arc as a signicant sediment source by the late Palaeocene
tinued Central Cordilleran input (either rst cycle or re- (Nie et al., 2010). All of these data are consistent with a
cycled) and recycling of pre-Cenozoic sequences. Likely mid-Palaeocene onset of large- scale exhumation of the
sources for the 200^300 and 500^650 Ma zircons are Juras- Central Cordillera and foreland basin sedimentation in
sic-Lower Cretaceous sediments or basement exposures the Floresta basin.

r 2010 The Authors


392 Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Cenozoic partitioning of the Floresta basin, Colombia

Zircon U^Pb ages and sandstone petrographic analysis DISCUSSION


indicate that by the middle levels of the Eocene Picacho
Formation, a new source area had been introduced in the Laterally continuous foreland basin
hinterland of the Floresta basin. This new source area is
characterized by zircons with 200^300 and 500^600 Ma Data presented here allow us to address whether the Co -
crystallization ages and a high percentage of metamorphic lombian foreland developed as a single, integrated, later-
lithic fragments (Figs 7a and 8i^q). One possible candi- ally contiguous basin (Cooper et al., 1995; Villamil, 1999;
date, the Santander massif north of the study area, appears Gomez et al., 2005a) or as multiple, isolated basins (Sar-
to be excluded as a potential sediment source area by miento -Rojas, 2001; Sarmiento -Rojas et al., 2006; Bayona
north-directed palaeocurrents (Fig. 6b) and Mio -Pliocene et al., 2008). The tectonic history described above, com-
apatite ssion track ages from the Santander massif (Sha- bined with additional published data (e.g. Horton et al.,
gam et al., 1984). Exposed basement rocks of the Quetame 2010b; Nie et al., 2010), suggests the Palaeocene^late Eo -
and Garzon massifs to the south (Fig. 1a) are excluded by cene Colombian foreland basin was a broad feature span-
zircon and apatite ssion track ages suggesting initial ning from the Magdalena basin in the west to the Llanos
basement exhumation in the early and late Miocene, re- basin in the east (Fig. 9).
spectively (van der Wiel, 1991; Parra et al., 2009b). A possi- We nd no direct evidence for sediment derived from
ble source for this deeply exhumed metamorphic material Late Cretaceous uplift in the Central Cordillera, despite
may be suggested by the regional middle Eocene uncon- widespread conclusions to the contrary. Cratonic zircons
formity which truncates Mesozoic units and crystalline occur in lower Palaeocene strata in both the Nuevo Mundo
basement in the Magdalena Valley (Buitrago, 1994; Gomez syncline (eastern Middle Magdalena Valley; Fig. 1a) and
et al., 2003). Although this hypothesis remains speculative, Floresta basin. This coeval deposition of eastern, craton-
it is favoured by the apparent lack of viable alternatives and derived detritus indicates that the basins topographic axis,
the evidence that Magdalena Valley basement was exposed dened as the boundary between craton- and orogen-de-
at the time. rived depositional systems, was approximately in the Mag-
Increased sedimentary lithic fragments in the upper dalena Valley region during the early Palaeocene (Fig. 9a).
Eocene^Oligocene Concentracion Formation suggest in- Lower Palaeocene strata from the Nuevo Mundo syncline
troduction of a new, upper- crustal source to the Floresta are composed of ne-grained distal facies. If there was
basin. However, continued eastward palaeocurrents show Late Cretaceous^ early Palaeocene uplift of the Central
that this new source remained to the west of the basin. In Cordillera (Gomez et al., 2003; Villagomez et al., 2008),
addition, zircon U^Pb ages display a general increase in there is no evidence of it in either the preserved sedimen-
the age of the youngest peak from the Picacho to Concen- tary facies or available provenance data (note widespread
tracion Formation (Fig. 8i^q) even though volcanism in distal depositional environments in Figs 9 and 10 of Coop-
the Central Cordillera continued during this time (Aspden er et al., 1995; Fig. 6 of Villamil, 1999; Fig. 2.16 of Sarmien-
et al., 1987; Restrepo -Pace, 1992). These two points suggest to -Rojas, 2001; also Nie et al., 2010). Based on zircon U^Pb
the source for Concentracion detritus west of the Floresta ages on igneous rocks from the Central Cordillera, there
basin did not include a signicant rst- cycle volcanic was certainly a major Late Cretaceous magmatic event at
component. Nor, based on the lack of metamorphic lithic 100 to 71 Ma (Aspden & McCourt, 1986; Villagomez et
sandstone fragments, did it include a signicant basement al., 2008; Restrepo -Moreno et al., 2009; Villagomez,
component.These observations rule out both of the afore- 2010). Therefore, reported Late Cretaceous low-tempera-
mentioned western sources^the Central Cordilleran mag- ture thermochronological ages from the Central Cordil-
matic arc and Magdalena Valley basement. Therefore, the lera may be complicated by post-magmatic cooling and,
new upper- crustal source most likely represents an east- particularly for igneous rocks, may record magmatic
ward step in the deformation front and recycling of fore- rather than exhumational cooling. However, a lack Central
land basin strata in the westernmost Eastern Cordillera. Cordilleran uplift does not rule out eastward sediment
The implication is that the Concentracion Formation was progradation from the Mesozoic volcanic edice to proxi-
sourced principally from recycled Mesozoic and Cenozoic mal locations (e.g. the Magdalena Valley basin, Gomez et
strata during initial exhumation of the western ank of the al., 2003).
Eastern Cordillera. A synchronous, mid-Palaeocene change from craton-
Following the mid-Palaeocene arrival of orogenic detri- to Central Cordillera- sourced sediment is observed in
tus in the Floresta basin, there is clear eastward progression both the Nuevo Mundo syncline in the Middle Magdalena
in exhumation of successive source regions. Sequentially, Valley (Nie et al., 2010) and Floresta basin in the axial East-
the Central Cordilleran arc, the MadgalenaValley basement ern Cordillera (this study). In the MagdalenaValley, the ob-
and nally the western ank of the Eastern Cordilleran served change occurs between the lower and upper
fold-thrust belt were exhumed as a result of systematic cra- Palaeocene Lisama Formation and involves a shift from
tonward advance of deformation. The evidence for defor- 4500 Ma to Cretaceous-age zircons (Nie et al., 2010).
mation advance from the magmatic arc toward the craton The combined data suggest a single integrated basin in
is consistent with conventional foreland basin models but the Palaeocene encompassing the Floresta and Magdalena
contradicts broken foreland basin models. basins. Sediment shed westward from the craton in the

r 2010 The Authors


Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 393
J. E. Saylor et al.

early Palaeocene reached both the Floresta and Magdalena Cordillera (Cooper et al., 1995) indicates that the basin axis
basins. Similarly, sediment shed eastward from the Central had shifted 4200 km eastward from the Magdalena Valley
Cordillera in the late Palaeocene reached both basins. Ba- to the eastern foothills (Fig. 9b). Thick, well developed
sin subsidence histories for the Llanos basin and Magdale- palaeosols and an erosional unconformity at the top of
na Valley indicate an increase in subsidence rate in the the Socha Mudstone are tentatively interpreted as devel-
Palaeocene (Gomez et al., 2005a; Bayona et al., 2008). In opment and passage of a exural forebulge in the Floresta
light of the combined sediment accumulation, provenance basin area (DeCelles & Currie, 1996; DeCelles & Horton,
and palaeocurrent data, we interpret the coeval arrival of 2003). If accurate, this would imply that the foreland basin
Central Cordilleran detritus in the Magdalena and Flores- was overlled (i.e. sediment derived from the orogenic
ta basins to be the result of the mid-Palaeocene onset of hinterland reached the backbulge depozone) in the Palaeo -
uplift in the Central Cordillera (Fig. 9b) (Nie et al., 2010). cene (Fig. 9b).
Synchronous late Palaeocene deposition of Central Cor- The middle Eocene involved eastward migration of the
dilleran detritus in the Floresta and Magdalena basins and deformation front into the Middle Magdalena Valley, ex-
cratonic detritus in the eastern foothills of the Eastern pansion of the region of uplift and deepening of the fore-

(a) (b)

(c) (d)

Fig. 9. Schematic palaeogeographic reconstructions using a modern (non-restored) map base showing present 500 m topographic
contour [modied from Cooper et al. (1995),Villamil, (1999), Sarmiento -Rojas, 2001, Gomez et al., (2005a), and Bayona et al., (2008)]. Key
observations are listed as notes 1^14.

r 2010 The Authors


394 Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Cenozoic partitioning of the Floresta basin, Colombia

land basin (Gomez et al., 2003, 2005b) (Fig. 9c). Several source areas during basin evolution. In addition, the de-
lines of evidence point to an underlled foreland basin at positional systems observed in the Floresta basin can be
this time. Average palaeocurrent orientations in the Flor- correlated westward to the Magdalena Valley, southward
esta basin shift from eastward (perpendicular to basin axis) along the Eastern Cordillera axis and eastward to the east-
to northward (parallel to basin axis). Likewise palaeocur- ern foothills and Llanos basin. Specically, both the Gua-
rent orientations in the eastern foothills change upsection duas (Floresta basin) and Lisama (Magdalena Valley)
from westward to north or northeastward (Cooper et al., formations record the Maastrichtian- early Palaeocene
1995; Parra et al., 2009a). Synchronous deposition of west- transition from marine to uvial deposition (Gomez et al.,
derived detritus in the Floresta basin and axial Eastern 2005b; Moreno, 2010).The transition from sand- to mud-
Cordillera and craton-derived detritus in the easternmost dominated uvial deposits in the Socha Group has lateral
Eastern Cordillera indicates that sediment derived from equivalents to the south (Cacho and Bogota Formations)
the orogenic hinterland was not reaching the eastern foot- (Gomez et al., 2005a) and the east (Barco and Los Cuervos
hills at this time. Formations) (Cazier et al., 1995; Bayona et al., 2008). The
Eastward migration of the forebulge is consistent with a Eocene Picacho Formation (Floresta basin) has been cor-
decrease in sedimentation rate in the eastern foothills in related to the Mirador Formation (eastern foothills) (San-
the middle Eocene- early Oligocene (Bayona et al., 2008; tos et al., 2008; Bande, 2010) and the Regadera Formation
Parra et al., 2009a, 2010). A decrease in sedimentation rate (south of Floresta basin) (Gomez et al., 2005a). Finally, both
associated with a forebulge is also suggested by the pre- the Concentracion (Floresta basin) and Carbonera (eastern
sence of a sequence-bounding erosional unconformity foothills) formations record the late Eocene^Oligocene
within the Eocene Mirador Formation (Villamil, 1999; marine incursion (Cazier et al., 1995; Cooper et al., 1995;
Warren & Pulham, 2002; Bayona et al., 2007, 2008) and Bayona et al., 2008; Santos et al., 2008). Signicantly, only
change from a quartz-rich, cratonic source to lithic-rich, after the late Eocene^Oligocene uplift of the westernmost
orogenic source across the unconformity (Bayona et al., Eastern Cordillera (Nie et al., 2010; this study) are devia-
2008; Bande, 2010). tions detected in depositional environments and palaeo -
Introduction of sediment recycled from the western currents between the time- equivalent Concentracion
ank of the Eastern Cordillera in the Concentracion For- (Floresta basin) and Esmeraldas (MagdalenaValley) forma-
mation does not necessitate a Laramide- style broken fore- tions (Gomez et al., 2003; Moreno, 2010).
land as this could be the result of eastward propagation of (2) The 2 km Cenozoic ll of the Floresta basin was
the deformation front and transition of the MagdalenaVal- deposited over 40 Myr suggesting a remarkably slow un-
ley from a foredeep or wedge-top depozone (DeCelles & decompacted sedimentation rate of 50 m Myr  1, implying
Giles, 1996) to a hinterland basin (e.g. McQuarrie et al., an even lower tectonic subsidence rate (Van Hinte, 1978;
2005; Horton, in press). Nor do palaeocurrent data pub- Dickinson et al., 1987; Allen & Allen, 2005). Low rates of
lished herein point to a complex palaeoow system as is ty- long-term tectonic subsidence across the Eastern Cordil-
pical of broken foreland basins (Jordan, 1995; Gomez et al., lera of 12.5^25 m Myr  1 (Fabre, 1983; Bayona et al., 2008)
2005a; Smith et al., 2008). It does, however, indicate that, are similar to low rates for the eastern foothill and Llanos
by the late Eocene, the Madgalena and Floresta basins basin (o12.5^23 m Myr  1) (Bayona et al., 2008; Parra et al.,
were evolving separately. 2009a). Palaeogene subsidence rates are uniformly low and
do not vary signicantly between the axial Eastern Cordil-
leran and eastern foothills or Llanos basin until the mid-
late Oligocene (Fabre, 1983; Bayona et al., 2008; Parra et al.,
BROKEN FORELAND BASIN 2009a).
Several studies suggest the Palaeocene^ early Eocene East- (3) No signicant occurrences of Palaeocene coarse-
ern Cordillera developed as a broken foreland (Sarmiento - grained proximal deposits are reported for the Eastern
Rojas, 2001) with subbasins separated by 1^3 km topo - Cordillera. Preserved Palaeocene facies represent distal en-
graphic barriers (Fig. 12a^d of Bayona et al., 2008). In the vironments and stratigraphic units can often be correlated
following paragraphs we compare the predictions of this regionally across the Eastern Cordillera (e.g. Cooper et al.,
hypothesis to observations concerning (1) variability and 1995; Villamil, 1999; Sarmiento-Rojas, 2001; Gomez et al.,
uniqueness of depositional environments in individual 2005a; Bayona et al., 2008). Although not impossible, it is
subbasins, (2) rate and uniqueness of subsidence patterns considered highly unlikely that all coarse-grained detritus
in potential subbasins, (3) coarse-grained, proximal sedi- was selectively cannibalized during subsequent exhuma-
ments, (4) sediment composition, (5) sediment dispersal tion, preferentially preserving only ne-grained deposits.
patterns, (6) growth strata and (7) thermochronology. Therefore, if localized uplifts were present, they must have
(1) The Floresta basin recorded neither rapid nor dra- been too small to produce coarse-grained detritus.
matic shifts in depositional environment. Lithofacies ana- (4) Proposed Palaeocene uplift of 1^3 km topographic
lysis suggests marine to marginal marine (including low- barriers (Bayona et al., 2008) in the Eastern Cordillera
gradient uvial) as opposed to piedmont (alluvial fan or would be associated with erosion of local, Mesozoic cover
high-gradient uvial) depositional conditions. The sedi- and local basement. However, there is no compelling
mentology does not support introduction of proximal provenance data requiring that the Santander massif or

r 2010 The Authors


Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 395
J. E. Saylor et al.

Eastern Cordilleran basement was exhumed at this time et al., 1984) raise the possibility of early Andean exhuma-
(see also evidence presented by Parra et al., 2009b). tion. However, the Santander massif was a palaeohigh dur-
Although Maastrichtian-lower Palaeocene strata are mar- ing Mesozoic rifting (Cooper et al., 1995; Kammer &
ginally more lithic rich than underlying Cretaceous strata Sanchez, 2006) and it is unlikely that it was covered with
(Bayona et al., 2008; this study), detrital zircon U^Pb geo - the 8 km of sediment needed to fully anneal ZFTs (240^
chronology indicates an eastern cratonic sediment source 250 1C at cooling rates of 0.5^1km Myr  1 (Brandon et al.,
in the early Palaeocene and a western, magmatic arc source 1998; Garver et al., 1999; Bernet et al., 2001; Tagami, 2005)).
in the late Palaeocene. Detrital zircon U^Pb age spectra In summary, although we cannot entirely rule out the
from the Cretaceous^lower Eocene sequence lack signi - occurrence of small, localized uplifts in the Palaeogene,
cant peaks between 900 and 200 Ma (Fig. 8d^h) which the lateral continuity of depositional environments, slow
would be expected if the Santander massif or Eastern Cor- subsidence rates, absence of clear provenance data and
dilleran basement were exposed (Horton et al., 2010b; Nie lack of coarse-grained proximal deposits, growth struc-
et al., 2010). tures, and thermochronological data to support their exis-
(5) Broken foreland basins are typied by complex, often tence is compelling. These observations and (a) the
reversing, sediment dispersal patterns (Davila & Astini, similarity in provenance shifts in the MagdalenaValley ba-
2007; Carrapa et al., 2008; Smith et al., 2008). Our palaeo - sin and Floresta basin, (b) depositional and stratigraphic
current data show a relatively uniform sediment dispersal continuity temporally within the Floresta succession and
pattern, in contrast to a recent compilation which shows laterally among Palaeogene exposures, (c) low sediment ac-
signicant scatter (Bayona etal., 2008 and references there- cumulation rates, (d) simple palaeocurrent patterns and (e)
in).We favour a simple sediment dispersal pattern because systematic cratonward progression of exhumation are in-
it provides the most straightforward explanation for com- consistent with a broken foreland setting (Jordan, 1995;
bined palaeocurrent and U^Pb provenance data. For ex- Jordan et al., 2001; Strecker et al., 2007) but consistent with
ample, given the clear U^Pb signal of Central Cordilleran a conventional foreland basin setting, particularly a distal
provenance (this study), previous suggestions of west-di- foreland basin setting (Sinclair et al., 1991; Crampton & Al-
rected palaeocurrents in the Floresta basin area in the late len, 1995; DeCelles & Currie, 1996; Horton et al., 2001; De-
Palaeocene (Fig. 8b of Bayona et al., 2008) would only be Celles & Horton, 2003). If the Palaeogene evolution of the
tenable if Central Cordilleran detritus was exhumed and Colombian foreland basin, and, by implication, orogenic
transported eastward, then re- eroded and transported wedge, followed such a traditional model, then later Neo -
back to the west to a nal location in the Floresta basin. gene reactivation of Mesozoic extensional structures
Further, because zircon crystallization ages are indistin- (Kammer & Sanchez, 2006; Mora et al., 2006, 2009; Parra
guishable from depositional age, this complex scenario et al., 2009b) suggests a major change in deformation style
would have to take place in a narrow time interval. In sum- during the mid-Cenozoic.
mary, following the arrival of orogenic sediment in the
mid-Palaeocene, sediment sources for the Floresta basin
were located entirely to the west throughout the Palaeo -
CONCLUSIONS
gene^ early Oligocene (this study), suggesting a lack of sig-
nicant topography east of the Floresta basin until the late Upper Cretaceous^Oligocene strata in the Floresta basin
Oligocene. At no point in the Palaeocene^Oligocene record a change from oshore marine to deltaic, then u-
development of the Floresta basin do we see evidence of vial and, nally, estuarine deposition. The 2 km thick
signicant (1^3 km) topographic barriers or multiple sedi- Cenozoic section accumulated over 40 Myr at an aver-
ment sources as expected for a broken foreland setting (e.g. age long-term, undecompacted rate of 0.05 mm yr  1. No -
Mortimer et al., 2007; Carrapa et al., 2008). where in the Palaeogene section are coarse-grained
(6) A broken foreland with localized uplift would gener- sediments observed. Detrital zircon age spectra and sand-
ate growth strata in coarse-grained, proximal deposits stone petrographic results indicate that sediment sources
within the Eastern Cordillera (e.g. Zapata & Allmendinger, for the Floresta basin changed during the Palaeocene (So -
1996; Seager et al., 1997; Casas-Sainz et al., 2002; Lopez- cha Group) from an exclusively Palaeo - and Mesoprotero -
Blanco, 2002; Strecker et al., 2007). However, examination zoic, cratonic source to a lithic-rich source with Jurassic^
of key examples of possible growth strata (Julivert, 1963, Cenozoic zircons. Because palaeocurrent orientations in
1970) leads to the conclusion that the Palaeocene^lower the Palaeocene are east-directed, this change is considered
Eocene units in question (Guaduas, Cacho and Bogota to reect the rst unambiguous inux of Central Cordiller-
Formations) are concordant over a stratigraphic thickness an detritus, most likely in response to the onset of shorten-
of 1000 m, with no clear evidence of internal unconfor- ing-related uplift.
mities before middle-late Eocene deposition (Usme For- A subsequent, middle Eocene provenance shift (Pica-
mation) (see Fig. 3b of Gomez et al., 2005a). cho Formation) records the introduction of Triassic to
(7) Available thermochronological data do not support a Neoproterozoic zircons and inux of more metamorphic
series of Cretaceous^Palaeocene uplifts within the Eastern lithic grains. Northward palaeocurrent orientations for
Cordillera (Parra et al., 2009b). Albian-Eocene zircon s- these strata and previously published basement low-tem-
sion track (ZFT) ages from the Santander massif (Shagam perature thermochronology appear to rule out all possible

r 2010 The Authors


396 Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Cenozoic partitioning of the Floresta basin, Colombia

sources of metamorphic basement except the Magdalena and d) Qt-F-L. Black arrows indicate upsection composi-
Valley basement which was exposed during Eocene hinter- tional trends. Colored small symbols are individual sam-
land shortening. As middle Eocene sediments are inferred ples. Fields for a, b, and d from Dickinson (1983; 1985).
to be sourced from the Magdalena Valley basement in the Fields for c from Folk (1980).
west, we conclude that the western (Magdalena) basin was Table S1. Recalculated point counting data.
united with basins to the east (Floresta and Llanos) until Table S2. U-Pb data.
the late Eocene. Please note: Wiley-Blackwell is not responsible for the
Introduction of a new upper- crustal sediment source in content or functionality of any supporting materials sup-
the late Eocene (Concentracion Formation) is linked to in- plied by the authors. Any queries (other than missing ma-
itial uplift of the western ank of the Eastern Cordillera terial) should be directed to the corresponding author for
and recycling of older foreland basin sediments. the article.
Simple eastward or northeastward palaeocurrent orien-
tations western sediment sources and similarities in
depositional environments and subsidence rates suggest REFERENCES
that the Floresta basin was connected with the Llanos ba-
sin from the Late Cretaceous through late Oligocene. Allen, P. & Allen, J. (2005) Basin Analysis: Principles and
In considering available thermochronology, sedimen- Applications, 2nd edn. Blackwell Scientic Publications,
tology and structural results, we nd no direct evidence Oxford.
Alvarez, J. (1983) Geolog| a de la Cordillera Central y el Occi-
to support uplift of the Central Cordillera until the mid-
dente Colombiano y petroqu| mica de los intrusivos grani-
Palaeocene when a synchronous arrival of Central Cordil-
toides Mesocenozoicos. Bolet| n de Geolog| a, Colombia
ler-derived sediments occurs in the easternmost Magda- Instituto Nacional de Investigaciones Geologico -Mineras.
lena Valley (Nuevo Mundo syncline) and Eastern 26, 175.
Cordillera (Floresta basin). Likewise, the lack of growth Aschoff, J.L. (2008) Controls on clastic wedge and growth stra-
structures, coarse-grained deposits, clear provenance ta development in foreland basins: examples from Cretaceous
data, or thermochronological data to support numerous Cordilleran foreland basin strata, USA. Ph.D. dissertation,
localized uplifts in the Eastern Cordillera is consistent The University of Texas at Austin, Austin.
with a single integrated foreland basin during the Palaeo - Aspden, J.A. & McCourt,W.J. (1986) Mesozoic oceanic terrane
gene. Moreover, the progressive introduction of Central in the central Andes of Colombia. Geology, 14, 415^418.
Cordillera, Magdalena Valley and Eastern Cordillera sedi- Aspden, J.A., McCourt,W.J. & Brook, M. (1987) Geometrical
control of subduction-related magmatism ^ the Mesozoic and
ment sources along with the lateral continuity of deposi-
Cenozoic plutonic history of western Colombia. J. Geol. Soc.,
tional environments are consistent with a systematic 144, 893^905.
eastward propagation of the deformation front into an un- Ayala-Calvo, R. & Delgado-Rozo, J. (2004) Analisis favial,
broken foreland basin throughout the Palaeogene. petrograa y eventos diageneticos de los intervalos calcareos de la for-
macion Tibasosa en los sectores de Tibasosa y Corrales ^ Beteitiva
(Boyaca, Colombia). Universidad Nacional de Colombia,
Bogota.
ACKNOWLEDGEMENTS Bande, A. (2010) Foreland basin evolution and exhumation
We thank W. Cavazza, B. Carrapa and F. Roure for reviews along the deformation front of the Eastern Cordillera, north-
which helped to strengthen this paper. This research also ern Andes, Colombia. M.S. thesis, The University of Texas at
benetted from discussions with C. Olariu, C. J. Moreno, Austin, Austin.
Barbeau, D.L. (2003) A exural model for the Paradox basin: im-
A. Bande, G. Bayona and E. Gomez. Laboratory assistance
plications for the tectonics of the Ancestral Rocky Mountains.
was provided by J. Knowles. Funding was provided by
Basin Res., 15, 97^115.
Ecopetrol-Instituto Colombiano del Petroleo (ICP-Project Bayona, G., Cortes, M., Jaramillo, C., Ojeda, G., Aristiza-
Cronologia de la deformacion en las Cuencas Subandi- bal, J.J. & Reyes-Harker, A. (2008) An integrated analysis of
nas), and the Jackson School of Geosciences as part of a an orogen- sedimentary basin pair: latest Cretaceous^Ceno -
collaborative research agreement between ICP and the zoic evolution of the linked Eastern Cordillera orogen and the
University of Texas at Austin. Many Colombian research- Llanos foreland basin of Colombia. Geol. Soc. Am. Bull., 120,
ers from Ecopetrol shared valuable information and logis- 1171^1197.
tical support during this research. Bayona, G., Jaramillo, C., Rueda, M., Reyes-Harker, A. &
Torres,V. (2007) Paleocene-middle Miocene exural-margin
migration of the nonmarine Llanos foreland basin of Colom-
bia. Ciencia,Tecnol. Futuro, 3, 141^160.
SUPPORTING INFORMATION Beaumont, C. (1981) Foreland basins. Geophys. J. Roy. Astronom.
Soc., 65, 291^329.
Additional Supporting Information may be found in the Bernet, M., Zattin, M., Garver, J., Brandon, M. & Vance, J.
online version of this article: (2001) Steady- state exhumation of the European Alps. Geology,
29, 35^38.
Figure S1. Modal sandstone petrographic data plotted Bhattacharya, J.P. & Walker, R.G. (1992) Deltas. In: Facies
in ternary diagrams: a) Lm-Lv-Ls, b) Qm-F-Lt, c) Q-F-L, Models: Response to Sea Level Change (Ed. by R. Walker & N.

r 2010 The Authors


Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 397
J. E. Saylor et al.

James), pp. 157^178. Geological Association of Canada, St. zircon ngerprint of the Proto -Andes: Evidence for a Neopro -
Johns, Canada. terozoic active margin? Precambrian Res., 167, 186^200.
Bird, P. (1984) Laramide crustal thickening event in the Rocky- Coleman, J.M. & Wright, L.D. (1975) Modern river deltas:
mountain foreland and Great Plains.Tectonics, 3, 741^758. variability of processes and sand bodies. In: Deltas: Models for
Boyd, R., Dalrymple, R. & Zaitlin, B.A. (1992) Classication Exploration (Ed. by M.L. Broussard), pp.99^149. Houston Geo -
of clastic coastal depositional environments. Sediment. Geol., logical Society, Houston.
80, 139^150. Colletta, B., Hebrard, F., Letouzey, J.,Werner, P. & Rudki-
Boyer, S.E. (1995) Sedimentary basin taper as a factor control- weicz, J. (1990) Tectonic style and crustal structure of the
ling the geometry and advance of thrust belts. Am. J. Sci., 295, Eastern Cordillera, Colombia, from a balanced cross section.
1220^1254. In: Petroleum and Tectonics in Mobile Belts (Ed. by J. Letouzey),
Brandon, M.T., Roden-Tice, M.K. & Garver, J.I. (1998) Late pp. 81^100. EditionsTechnip, Paris.
Cenozoic exhumation of the Cascadia accretionary wedge in Collinson, J. (1996) Alluvial sediments. In: Sedimentary Envir-
the Olympic Mountains, northwest Washington state. Geol. onments: Processes, Facies and Stratigraphy (Ed. by H.G. Reading),
Soc. Am. Bull., 110, 985^1009. pp. 37^82. Blackwell Science, Oxford.
Bridge, J. (2003) Rivers and Floodplains,1st edn. Blackwell Science Coney, P.J. & Evenchick, C.A. (1994) Consolidation of the
Inc., Oxford. American Cordilleras. J. Soc. Am. Earth Sci., 7, 241^262.
Bristow, C. & Best, J. (1993) Braided rivers: perspectives and Cooper, M.A., Addison, F.T., Alvarez, R., Coral, M., Gra-
problems. In: Braided Rivers (Ed. by J.L. Best & C.S. Bristow), ham, R.H., Hayward, A.B., Howe, S., Martinez, J., Naar,
Geol. Soc. Spec. Publ. Lond., 75, 1^11. J., Penas, R., Pulham, A.J. & Taborda, A. (1995) Basin devel-
Buitrago, J. (1994) Petroleum systems of the Neiva area, Upper opment and tectonic history of the Llanos basin, Eastern Cor-
Magdalena Valley, Colombia. In: The Petroleum System ^ from dillera, and Middle Magdelena Valley, Colombia. Am. Assoc.
Source to Trap (Ed. by L. Magoon & W. Dow), Am. Assoc. Petrol. Petrol. Geol. Bull., 79, 1421^1443.
Geol. Mem.,Tulsa, 60, 483^497. Cordani, U.G., Cardona, A., Jiminez, D.M., Liu, D. & Nut-
Burbank, D., Beck, R. & Mulder, T. (1996) The Himalayan man, A.P. (2005) Geochronology of Proterozoic basement in-
foreland basin. In: The Tectonic Evolution of Asia (Ed. by A. Yin liers in the Colombian Andes: tectonic history of remnants of a
& M. Harrison), pp. 148^188. Cambridge University Press, fragmented Grenville belt. In: Terrane Processes at the Margins of
Cambridge. Gondwana (Ed. by A.P.M.Vaughan, P.T. Leat & R.J. Pankhurst),
Cant, D. & Walker, R. (1978) Fluvial processes and facies se- Geol. Soc. London Spec. Publ., Lond.,, 246, 329^346.
quences in sandy braided South Saskatchewan River, Canada. Corredor, F. (2003) Eastward extent of the late Eocene^ early
Sedimentology, 25, 625^648. Oligocene onset of deformation across the northern Andes:
Cardona, A., Chew, D., Valencia, V.A., Bayona, G., Misko- constraints from the northern portion of the Eastern Cordil-
vic, A. & Ibanez-Mejia, M. (2010) Grenvillian remnants in lera fold belt, Colombia. J. Soc. Am. Earth Sci., 16, 445^457.
the northern Andes: Rodinian and Phanerozoic paleogeo - Crampton, S.L. & Allen, P.A. (1995) Recognition of forebulge
graphic perspectives. J. S. Am. Earth Sci., 29, 92^104. unconformities associated with early- stage foreland basin de-
Carr, I.D., Gawthorpe, R.L., Jackson, C.A.L., Sharp, I.R. & velopment ^ example from the north Alpine foreland basin.
Sadek, A. (2003) Sedimentology and sequence stratigraphy of Am. Assoc. Petrol. Geol. Bull., 79, 1495^1514.
early syn-rift tidal sediments: the Nukhul Formation, Suez Currie, B.S. (1997) Sequence stratigraphy of nonmarine Juras-
rift, Egypt. J. Sediment. Res., 73, 407^420. sic-Cretaceous rocks, central Cordilleran foreland-basin sys-
Carrapa, B., Hauer, J., Schoenbohm, L., Strecker, M.R., tem. Geol. Soc. Am. Bull., 109, 1206^1222.
Schmitt, A.K., Villanueva, A. & Gomez, J.S. (2008) Dalrymple, R.W. & Choi, K. (2007) Morphologic and facies
Dynamics of deformation and sedimentation in the northern trends through the uvial-marine transition in tide-domi-
Sierras Pampeanas: An integrated study of the Neogene nated depositional systems: A schematic framework for envir-
Fiambala basin, NW Argentina. Geol. Soc. Am. Bull., 120, 1518^ onmental and sequence- stratigraphic interpretation. Earth
1543. Sci. Rev., 81, 135^174.
Casas-Sainz, A.M., Cortes, A.L. & Maestro, A. (2002) Se- Dalrymple, R.W., Zaitlin, B.A. & Boyd, R. (1992) Estuarine
quential limb rotation and kink-band migration recorded by facies models ^ conceptual basis and stratigraphic implica-
growth strata, Almazan basin, north Spain. Sediment. Geol., tions. J. Sediment. Petrol., 62, 1130^1146.
146, 25^45. Davila, F.M. & Astini, R.A. (2007) Cenozoic provenance his-
Cazier, E.C., Cooper, M.A., Eaton, S.G. & Pulham, A.J. tory of synorogenic conglomerates in western Argentina (Fa-
(1997) Basin development and tectonic history of the Llanos matina belt): Implications for Central Andean foreland
basin, Eastern Cordillera, and Middle Magdalena Valley, Co - development. Geol. Soc. Am. Bull., 119, 609^622.
lombia: reply. Am. Assoc. Petrol. Geol. Bull., 81, 1332^1335. DeCelles, P.G., Gray, M., Ridgway, K., Cole, R., Pivnik, D.,
Cazier, E.C., Hayward, A.B., Espinosa, G., Velandia, J., Pequera, N. & Srivastava, P. (1991) Controls on synorogenic
Mugniot, J.F. & Leel,W.H. (1995) Petroleum geology of the alluvial-fan architecture, Beartooth conglomerate (Paleocene),
Cusiana eld, Llanos basin foothills, Colombia. Am. Assoc. Pet- Wyoming and Montana. Sedimentology, 38, 567^590.
rol. Geol. Bull., 79, 1444^1463. DeCelles, P.G., Langford, R. & Schwartz, R. (1983) Two
Cediel, F., Shaw, R. & Caceres, C. (2003) Tectonic assembly of new methods of paleocurrent determination from trough
the northern Andean block. In: The Circum-Gulf of Mexico and cross- stratication. J. Sediment. Petrol., 53, 629^642.
the Caribbean: Hydrocarbon Habitats, Basin Formation, and Plate DeCelles, P.G. & Currie, B.S. (1996) Long-term sediment ac-
Tectonics (Ed. by C. Bartolini, R.T. Buer & J. Blickwede), Am. cumulation in the middle Jurassic- early Eocene Cordilleran
Assoc. Petrol. Geol. Mem.,Tulsa, 79, 815^848. retroarc foreland-basin system. Geology, 24, 591^594.
Chew, D.M., Magna,T., Kirkland, C.L., Miskovic, A., Car- DeCelles, P.G. & Giles, K.A. (1996) Foreland basin systems.
dona, A., Spikings, R. & Schaltegger, U. (2008) Detrital Basin Res., 8, 105^123.

r 2010 The Authors


398 Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Cenozoic partitioning of the Floresta basin, Colombia

DeCelles, P.G. & Horton, B.K. (2003) Early to middleTertiary by detrital fssion-track thermochronology. In: Exhumation
foreland basin development and the history of Andean crustal Processes: Normal Faulting, Ductile Flow and Erosion (Ed. by U.
shortening in Bolivia. Geol. Soc. Am. Bull., 115, 58^77. Ring, M.T. Brandon, G.S. Lister & S.D. Willett), Geol. Soc.
Dengo, C.A. & Covey, M.C. (1993) Structure of the Eastern Lond., 154, 283^304.
Cordillera of Colombia - Implications for trap styles and re- Gehrels, G. (2000) Introduction to detrital zircons studies
gional tectonics. Am. Assoc. Petrol. Geol. Bull., 77, 1315^1337. of Paleozoic and Triassic strata in western Nevada and north-
Dickinson, W.R. (1985) Provenance relations from detrital ern California. In: Paleozoic and Triassic Paleogeography and
modes of sandstones. In: Provenace of Arenites (Ed. by G.G. Zuf- Tectonics of Western Nevada and Northern California (Ed. by M.J.
fa), pp. 333^361. D. Reidel Publishing Co., Dordrecht. Soreghan & G.E. Gehrels), Geol. Soc. Am. Spec. Pap., Boulder,
Dickinson, W.R., Armin, R., Beckvar, N., Goodlin, T., Ja- 347, 1^17.
necke, S., Mark, R., Norris, R., Radel, G. & Wortman, Gehrels, G.E.,Valencia,V.A. & Ruiz, J. (2008) Enhanced pre-
A. (1987) Geohistory analysis of rates of sediment accumula- cision, accuracy, eciency, and spatial resolution of U^Pb ages
tion and subsidence for selected California basins. In: Cenozoic by laser ablation-multicollector-inductively coupled plasma-
Basin Development of Coastal California (Ed. by R.V. Ingersoll & mass spectrometry. Geochem. Geophys. Geosyst., 9, 1^13.
W.G. Ernst), pp. 1^23. Prentice Hall, Englewood Clis. Goldsmith, R., Marvina, R.F. & Mehnert, H.H. (1971)
Dickinson, W.R. (1974) Plate tectonics and sedimentation. Radiometric ages in the Santander massif, Eastern Cordillera,
In: Tectonics and Sedimentation (Ed. by W.R. Dickinson), pp. 1^ Colombian Andes. U.S. Geol. Sur., Professional Pap., 750 -D,
27. Society of Economic Paleontologists and Mineralogists, D44^D49.
Tulsa. Gomez, E., Jordan,T.E., Allmendinger, R.W. & Cardozo, N.
Dickinson, W.R., Beard, L.S., Brakenridge, G.R., Erjavec, (2005a) Development of the Colombian foreland-basin system
J.L., Ferguson, R.C., Inman, K.F., Knepp, R.A., Lind- as a consequence of diachronous exhumation of the northern
berg, F.A. & Ryberg, P.T. (1983) Provenance of North Ameri- Andes. Geol. Soc. Am. Bull., 117, 1272^1292.
can Phanerozoic sandstones in relation to tectonic setting. Gomez, E., Jordan, T.E., Allmendinger, R.W., Hegarty, K.
Geol. Soc. Am. Bull., 94, 222^235. & Kelley, S. (2005b) Syntectonic Cenozoic sedimentation in
Dickinson,W.R. & Gehrels, G.E. (2008) Sediment delivery to the northern Middle Magdalena Valley basin of Colombia and
the cordilleran foreland basin: Insights from U^Pb ages of det- implications for exhumation of the northern Andes. Geol. Soc.
rital zircons in Upper Jurassic and Cretaceous strata of the Am. Bull., 117, 547^569.
Colorado plateau. Am. J. Sci., 308, 1041^1082. Gomez, E., Jordan, T.E., Allmendinger, R.W., Hegarty, K.,
Dickinson, W.R., Klute, M.A., Hayes, M.J., Janecke, S.U., Kelley, S. & Heizler, M. (2003) Controls on architecture of
Lundin, E.R., McKittrick, M.A. & Olivares, M.D. (1988) the Late Cretaceous to Cenozoic southern Middle Magdalena
Paleogeographic and paleotectonic setting of Laramide sedi- Valley basin, Colombia. Geol. Soc. Am. Bull., 115, 131^147.
mentary basins in the central Rocky Mountain region. Geol. Gregory-Wodzicki, K.M. (2000) Uplift history of the Central
Soc. Am. Bull., 100, 1023^1039. and Northern Andes: A review. Geol. Soc. Am. Bull., 112, 1091^
Dickinson, W.R. & Suczek, C.A. (1979) Plate tectonics and 1105.
sandstone compositions. Am. Assoc. Petrol. Geol. Bull., 63, Guerrero, J. & Sarmiento, G.A. (1996) Estratigraf| a f| sica, pa-
2164^2182. linologica, sedimentologica y secuencial del Creta cico super-
Dorr,W., Grosser, J.R., Rodriguez, G.I. & Kramm, U. (1995) ior y Paleoceno del piedemonte Llanero. Implicaciones en
Zircon U^Pb age of the Paramo -Rico tonalite-granodiorite, exploracion petrolera. Geol. Colomb., 20, 3^66.
Santander massif (Cordillera Oriental, Colombia) and its geo - Heller, P.L., Angevine, C.L., Winslow, N.S. & Paola, C.
tectonic signicance. J. S. Am. Earth Sci., 8, 187^194. (1988) Two-phase stratigraphic model of foreland-basin se-
Dunham, R. (1962) Classication of carbonate rocks according quences. Geology, 16, 501^504.
to depositional texture. Am. Assoc. Petrol. Geol. Mem., 1, 108^121. Horton, B.K. (2005) Revised deformation history of the Central
Ehlers,T.A. & Chan, M.A. (1999) Tidal sedimentology and es- Andes: Inferences from Cenozoic foredeep and intermontane
tuarine deposition of the Proterozoic Big Cottonwood Forma- basins of the Eastern Cordillera, Bolivia. Tectonics, 24, TC3011,
tion, Utah. J. Sediment. Res., 69, 1169^1180. doi: 10.1029/2003TC001619.
Etayo-Serna, F., Barrero, D., Lozano, H., Espinosa, A., Horton, B.K. (in press) Cenozoic evolution of hinterland basins
Gonzalez, H., Orrego, A., Ballesteros, I., Forero, H., in the Andes and Tibet. In: Recent Advances in Tectonics of Sedi-
Ramrez, C., Zambrano, F., Duque, H., Vargas, R., Nu- mentary Basins (Ed. by C.J. Busby & A. Azor) Wiley-Blackwell,
nez, A., Alvarez, J., Ropan, C., Cardozo, E., Galvis, N., Oxford.
Sarmiento, L., Albers, J., Case, J., Singer, D., Bowen, R., Horton, B.K., Hampton, B.A. & Waanders, G.L. (2001) Pa-
Berger, B., Cox, D. & Hodges, C. (1983) Mapa de terrenos leogene synorogenic sedimentation in the Altiplano plateau
geologicos de Colombia, INGEOMINAS. Bogota, Map with and implications for initial mountain building in the central
report (235pp.). Andes. Geol. Soc. Am. Bull., 113, 1387^1400.
Fabre, A. (1983) La subsidencia de la cuenca del Cocuy (Cordil- Horton, B.K., Parra, M., Saylor, J.E., Nie, J., Mora, A.,
lera Oriental de Colombia) durante el Creta ceo y el Terciario Torres, V., Stockli, D.F. & Strecker, M.R. (2010a) Resol-
primera parte estudio cuantitativo de la subsidencia. Geol. ving uplift of the northern Andes using detrital zircon age sig-
Norand., 8, 49^61. natures. GSA Today, 20, 4^9, doi: 10.1130/GSATG76A.1.
Flemings, P.B. & Jordan, T.E. (1989) A synthetic stratigraphic Horton, B.K., Saylor, J.E., Nie, J., Mora, A., Parra, M.,
model of foreland basin development. J. Geophys. Res. ^ Solid Reyes-Harker, A. & Stockli, D.F. (2010b) Linking sedi-
Earth, 94, 3851^3866. mentation in the northern Andes to basement conguration,
Folk, R.F. (1980) Petrology of Sedimentary Rocks. Hamphill, Austin. Mesozoic extension, and Cenozoic shortening: Evidence from
Garver, J.I., Brandon, M.T., Roden-Tice, M. & Kamp, P.J.J. detrital zircon U^Pb ages, Eastern Cordillera, Colombia. Geol.
(1999) Exhumation history of orogenic highlands determined Soc. Am. Bull., 122, 1423^1442.

r 2010 The Authors


Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 399
J. E. Saylor et al.

Howard, J.D. & Reineck, H.E. (1981) Depositional facies Lopez-Blanco, M. (2002) Sedimentary response to thrusting
of high- energy beach-to - oshore sequence ^ comparison and fold growing on the SE margin of the Ebro basin (Paleo -
with low- energy sequence. Am. Assoc. Petrol. Geol. Bull., 65, gene, NE spain). Sediment. Geol., 146, 133^154.
807^830. Ludwig, K. (2003) Isoplot 3.0, Berkeley Geochronology Center
Hoyt, J.H. & Henry,V.J. (1967) Inuence of island migration on Special Publication. Berkeley. 4, 70.
barrier-island sedimentation. Geol. Soc. Am. Bull., 78, 77^86. Lunt, I.A., Bridge, J.S. & Tye, R.S. (2004) A quantitative,
Ingersoll, R., Bullard, T., Ford, R., Grimm, J., Pickle, J. & three-dimensional depositional model of gravelly braided riv-
Sares, S. (1984) The eect of grain- size on detrital modes: a ers. Sedimentology, 51, 377^414.
test of the Gazzi-Dickinson point- counting method. J. Sedi- Mack, G.H., James,W.C. & Monger, H.C. (1993) Classication
ment. Petrol., 54, 103^116. of paleosols. Geol. Soc. Am. Bull., 105, 129^136.
Johnson, E.A. & Pierce, F.W. (1990) Variations in uvial Makaske, B. (2001) Anastomosing rivers: A review of their clas-
deposition on an alluvial plain ^ an example from the Tongue sication, origin and sedimentary products. Earth-Sci. Rev.,
River Member of the Fort Union Formation (Paleocene), 53, 149^196.
southeastern Powder River basin, Wyoming, USA. Sediment. Marocco, R., Lavenu, A. & Baudino, R. (1995) Intermontane
Geol., 69, 21^36. late Paleogene^Neogene basins of the Andes of Ecuador and
Johnsson, M.J., Stallard, R.F. & Lundberg, N. (1991) Con- Peru-Sedimentologic and tectonic characteristics. In: Petro-
trols on the composition of uvial sands from a tropical leum Basins of South America (Ed. by A.J. Tankard, R. Suarez
weathering environment: Sands of the Orinoco river Soruco & H.J. Welsink), Am. Assoc. Petrol. Geol. Mem., Tulsa, 62,
drainage-basin, Venezuela and Colombia. Geol. Soc. Am. Bull., 597^613.
103, 1622^1647. Marshak, S., Karlstrom, K. & Timmons, J.M. (2000) Inver-
Jones, H.L. & Hajek, E.A. (2007) Characterizing avulsion stra- sion of Proterozoic extensional faults: An explanation for the
tigraphy in ancient alluvial deposits. Sediment. Geol., 202, 124^ pattern of Laramide and Ancestral Rockies intracratonic de-
137. formation, United States. Geology, 28, 735^738.
Jordan, T.E. (1981) Thrust loads and foreland basin evolution, Mazzoli, S., Corrado, S., De Donatis, M., Scrocca, D.,
Cretaceous, western United-States. Am. Assoc. Petrol. Geol. Butler, R., Bucci, D., Naso, G., Nicolai, C. & Zucconi,V.
Bull., 65, 2506^2520. (2000) Time and space variability of thin- skinned and
Jordan, T.E. (1995) Retroarc foreland and related basins. In:Tec- thick- skinned thrust tectonics in the Apennines (Italy).
tonics of Sedimentary Basins (Ed. by C.J. Busby & R.V. Ingersoll), Rend. Lincei, 11, 5^39.
pp. 331^362. Blackwell Science, Cambridge. McCourt,W.J., Aspden, J.A. & Brook, M. (1984) New geologi-
Jordan, T.E. & Allmendinger, R.W. (1986) The Sierras Pam- cal and geochronological data from the Colombian Andes:
peanas of Argentina: A modern analog of Rocky Mountain continental growth by multiple accretion. J. Geol. Soc., 141,
foreland deformation. Am. J. Sci., 286, 737^764. 831^845.
Jordan, T.E., Schlunegger, F. & Cardozo, N. (2001) Unstea- McQuarrie, N., Horton, B.K., Zandt, G., Beck, S. &
dy and spatially variable evolution of the Neogene Andean DeCelles, P.G. (2005) Lithospheric evolution of the Andean
Bermejo foreland basin, Argentina. J. Soc. Am. Earth Sci., 14, fold-thrust belt, Bolivia, and the origin of the central Andean
775^798. plateau.Tectonophysics, 399, 15^37.
Julivert, M. (1963) Los rasgos tectonicos de la Sabana de Miall, A. (1978) Lithofacies types and vertical prole models
Bogota y los mecan| smos de formacion de las estructuras. in braided river deposits: a summary. In: Fluvial Sedimentology
Boletin de Geologia Universidad Industrial de Santander, 13^14, (Ed. by A.D. Miall), Mem. Can. Soc. Petrol. Geol. 5, 597^604.
5^102. Miall, A. (1996) The Geology of Fluvial Deposits. Springer-Verlag,
Julivert, M. (1970) Cover and basement tectonics in the Cordil- New York.
lera Oriental of Colombia, South America, and a comparison Mohrig, D., Heller, P.L., Paola, C. & Lyons,W.J. (2000) Inter-
with some other folded chains. Geol. Soc. Am. Bull., 81, 3623^ preting avulsion process from ancient alluvial sequences: Guada-
3646. lope-Matarranya system (northern Spain) and Wasatch
Kammer, A. & Sanchez, J. (2006) Early Jurassic rift structures Formation (western Colorado). Geol. Soc. Am. Bull., 112, 1787^1803.
associated with the Soapaga and Boyaca faults of the Eastern Molina, A.C., Cordani, U.G. & MacDonald, W.D. (2006)
Cordillera, Colombia: Sedimentological inferences and regio - Tectonic correlations of pre-Mesozoic crust from the northern
nal implications. J. S. Am. Earth Sci., 21, 412^422. termination of the Colombian Andes, Caribbean region. J. S.
Kimberley, M.M. (1980) Paz-de-Rio oolitic inland- sea iron Am. Earth Sci., 21, 337^354.
Formation. Econ. Geol., 75, 97^106. Mora, A., Gaona,T., Kley, J., Montoya, D., Parra, M., Quir-
Kraus, M.J. (1999) Paleosols in clastic sedimentary rocks: Their oz, L.I., Reyes, G. & Strecker, M.R. (2009) The role of in-
geologic applications. Earth-Sci. Rev., 47, 41^70. herited extensional fault segmentation and linkage in
Kraus, M.J. & Gwinn, B. (1997) Facies and facies architecture of contractional orogenesis: A reconstruction of lower Cretac-
Paleogene oodplain deposits, Willwood Formation, Bighorn eous inverted rift basins in the Eastern Cordillera of Colombia.
basin,Wyoming, USA. Sediment. Geol., 114, 33^54. Basin Res., 21, 111^137.
Kraus, M.J. & Wells, T. (1999) Recognizing avulsion deposits Mora, A., Parra, M., Strecker, M.R., Kammer, A., Dimate, C.
in the ancient stratigraphical record. In: Fluvial Sedimentology & Rodriguez, F. (2006) Cenozoic contractional reactivation of
VI (Ed. by N. Smith & J. Rovers), Spec. Publ. Int. Assoc. Sedimen- Mesozoic extensional structures in the Eastern Cordillera of
tol., Oxford. 28, 251^268. Colombia.Tectonics, 25, TC2010, doi: 10.1029/2005TC001854.
Lacombe, O. & Mouthereau, F. (2002) Basement-involved Mora, A., Parra, M., Strecker, M.R., Sobel, E.R., Hoo-
shortening and deep detachment tectonics in forelands of oro - ghiemstra, H., Torres,V. & Jaramillo, J.V. (2008) Climatic
gens: insights from recent collision belts (Taiwan, western forcing of asymmetric orogenic evolution in the Eastern Cor-
Alps, Pyrenees).Tectonics, 21, doi: 10.1029/2001TC901018. dillera of Colombia. Geol. Soc. Am. Bull., 120, 930^949.

r 2010 The Authors


400 Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Cenozoic partitioning of the Floresta basin, Colombia

Moreno, C.J. (2010) Paleogene sedimentation patterns and Restrepo-Pace, P.A. (1992) Petrotectonic characterization of
basin evolution during Andean orogenesis, Middle Magdalena the central Andean terrane, Colombia. J. S. Am. Earth Sci., 5,
Valley basin, Colombia. MS thesis, The University of Texas at 97^116.
Austin, Austin. Restrepo-Pace, P.A., Ruiz, J., Gehrels, G. & Cosca, M.
Mortimer, E., Carrapa, B., Coutand, I., Schoenbohm, L., (1997) Geochronology and Nd isotopic data of Grenville-age
Sobel, E.R. & Gomez, J.S. (2007) Fragmentation of a foreland rocks in the Colombian Andes: New constraints for late Proter-
basin in response to out- of- sequence basement uplifts and ozoic- early Paleozoic paleocontinental reconstructions of the
structural reactivation: El Cajon-Campo del Arenal basin, Americas. Earth Planet. Sci. Lett., 150, 427^441.
NW Argentina. Geol. Soc. Am. Bull., 119, 637^653. Rodriguez, A.J. & Solano, O. (2000) Mapa geoloogico del de-
Mutti, E., Tinterri, R., Benevelli, G., di Biase, D. & partamento de Boyaca , memoria explicativa, Ingeominas.
Cavanna, G. (2003) Deltaic, mixed and turbidite sedi- Bogota.
mentation of ancient foreland basins. Mar. Petrol. Geol., 20, Santos, C., Jaramillo, C., Bayona, G., Rueda, M. & Torres,
733^755. V. (2008) Late Eocene marine incursion in north-western
Nie, J., Horton, B.K., Mora, A., Saylor, J.E., Housh, T.B., South America. Paleogeogr. Paleoclimatol. Paleoecol., 264,
Rubiano, J. & Naranjo, J.A.O. (2010) Tracking evolution of 140^146.
Andean ranges bounding the Middle Magdalena Valley basin, Sarmiento -Rojas, L. (2001) Mesozoic rifting and Cenozoic
Colombia. Geology, 38, 451^454. basin inversion history of the Eastern Cordillera, Colombian
Orton, G.J. & Reading, H.G. (1993) Variability of deltaic pro - Andes: Inferences from Tectonic Models. PhD thesis, Vrije
cesses in terms of sediment supply, with particular emphasis Universiteit, Amsterdam.
on grain- size. Sedimentology, 40, 475^512. Sarmiento -Rojas, L.F., Van Wess, J.D. & Cloetingh, S.
Pardo-Tujillo, A. (2004) Paleocene^Eocene Palynology and (2006) Mesozoic transtensional basin history of the Eastern
Palynofacies from Northeastern Colombia and Western Vene- Cordillera, Colombian Andes: Inferences from tectonic mod-
zuela. PhD thesis, Universite de Liege, Liege. els. J. Soc. Am. Earth Sci., 21, 383^411.
Parra, M., Mora, A., Jaramillo, C., Strecker, M., Sobel, E., Scisciani,V. & Montefalcone, R. (2006) Coexistance of thin-
Quiroz, L., Rueda, M. & Torres, V. (2009a) Orogenic and thick- skinned tectonics: An example from the central
wedge advance in the northern Andes: Evidence from the Oli- Apennines, Italy. In: Styles of Continental Contraction (Ed. by
gocene^Miocene sedimentary records of the Medina basin, S. Mazzoli & R. Butler), Geol. Soc. Am. Spec. Pap., Boulder, 414,
Eastern Cordillera, Colombia. Geol. Soc. Am. Bull., 121, 780^ 33^54.
800. Seager, W.R., Mack, G.H. & Lawton, T.F. (1997) Structural
Parra, M., Mora, A., Jaramillo, C.,Torres,V., Zeilinger, G. kinematics and depositional history of a Laramide uplift-
& Strecker, M.R. (2010) Tectonic controls on Cenozoic basin pair in southern New Mexico: Implications for develop -
foreland basin development in the north- eastern Andes, ment of intraforeland basins. Geol. Soc. Am. Bull., 109, 1389^
Colombia. Basin Res. doi: 10.1111/j.1365-2117.2009.00459.x. 1401.
Parra, M., Mora, A., Sobel, E.R., Strecker, M.R. & Gonza- Shagam, R., Kohn, B.P., Banks, P.O., Dasch, L.E., Vargas,
lez, R. (2009b) Episodic orogenic front migration in the R., Rodriguez, G.I. & Pimentel, N. (1984) Tectonic implica-
northern Andes: Constraints from low-temperature thermo - tions of Cretaceous^Pliocene ssion-track ages from rocks of
chronology in the Eastern Cordillera, Colombia. Tectonics, 28 the circum-Maracaibo basin region of western Venezuela and
TC4004, doi:10.1029/2008TC002423. eastern Colombia. In: The Caribbean-South American Plate
Price, R.A. (1973) Large- scale gravitational ow of supracrustal Boundary and Regional Tectonics (Ed. by W.E. Bonini, R.B.
rocks, southern Canadian Rockies. In: Gravity and Tectonics Hargraves & R. Shagam), Geol. Soc. Am. Mem., Boulder, 162,
(Ed. by K. DeJong & R. Scholten), pp. 491^502. Wiley, New 385^412.
York. Sinclair, H.D. (1997) Tectonostratigraphic model for under-
Pulido, M., Rodriguez, G.I.,Torres,V.,Vargas, M., Jaramil- lled peripheral foreland basins: An Alpine perspective. Geol.
lo, C., Fiorini, F., Cardona, A., Bayona, G., Rueda, M., Soc. Am. Bull., 109, 324^346.
Rincon, D. & Arenas, J. (2006) Proyecto cronolog| a de Sinclair, H.D., Coakley, B.J., Allen, P.A. & Watts, A.B.
las secuencias bioestratigra cas del cenozoico del piedemonte (1991) Simulation of foreland basin stratigraphy using a diu-
Llanero, Smithsonian Tropical Research Institute & Instituto sion-model of mountain belt uplift and erosion: An example
Colombiano del Petroleo. Panama-Bucaramanga, 105. from the central Alps, Switzerland.Tectonics, 10, 599^620.
Reading, H.G. & Collinson, J. (1996) Clastic coasts. In: Sedi- Smith, D.G. (1986) Anastomosing river deposits, sedimentation
mentary Environments: Processes, Facies and Stratigraphy (Ed. by rates and basin subsidence, Magdalena river, northwestern
H.G. Reading), pp. 154^231. Blackwell Science, Oxford. Colombia, South America. Sediment. Geol., 46, 177^196.
Reineck, H.E., Howard, J.D., Frey, R.W., Dorjes, J. & Hert- Smith, D.G. & Smith, N.D. (1980) Sedimentation in anasto -
weck, G. (1970) Characteristics of shoreface sediments. Am. mosed river systems: Examples from alluvial valleys near
Assoc. Petrol. Geol. Bull., 54, 866. Ban, Alberta. J. Sediment. Petrol., 50, 157^164.
Reinson, G.E. (1992) Transgressive barrier island and estuarine Smith, M.E., Carroll, A.R. & Singer, B.S. (2008) Synoptic
systems. In: Facies Models: Response to Sea Level Change (Ed. by reconstruction of a major ancient lake system: Eocene Green
R.G. Walker & N. James), pp. 179^194. Geological Association River Formation, western United States. Geol. Soc. Am. Bull.,
of Canada, Ontario. 120, 54^84.
Restrepo-Moreno, S.A., Foster, D.A., Stockli, D.F. & Par- Stacey, J. & Kramers, J. (1975) Approximation of terrestrial lead
ra-Sanchez, L.N. (2009) Long-term erosion and exhuma- isotope evolution by aTwo- stage model. Earth Planet. Sci. Lett.,
tion of the Altiplano Antioqueno, northern Andes 26, 207^221.
(Colombia) from apatite (U-Th)/He thermochronology. Earth Strecker, M.R., Alonso, R.N., Bookhagen, B., Carrapa, B.,
Planet. Sci. Lett., 278, 1^12. Hilley, G.E., Sobel, E.R. & Trauth, M.H. (2007) Tectonics

r 2010 The Authors


Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 401
J. E. Saylor et al.

and climate of the southern central Andes. Annu. Rev. Earth. national Conference on Thermochronometry (Ed. by J.I. Garver &
Planet. Sci., 35, 747^787. M.J. Montario), pp. 253^255. Anchorage, Alaska.
Tagami, T. (2005) Zircon ssion-track thermochronology and Villamil, T. (1999) Campanian-Miocene tectonostratigraphy,
applications to fault studies. In: Low-Temperature Thermochro- depocenter evolution and basin development of Colombia
nology: Techniques, Interpretations, and Applications (Ed. by P.W. and western Venezuela. Paleogeogr. Paleoclimatol. Paleoecol., 153,
Reiners & T.A. Ehlers), Rev. Mineral. Geochem. 58, 95^122. 239^275.
Tassinari, C.C.G. & Macambira, M.J.B. (1999) Geochro - Vinasco, C.J., Cordani, U.G., Gonzalez, H.,Weber, M. & Pe-
nological provinces of the Amazonian craton. Episodes, 22, laez, C. (2006) Geochronological, isotopic, and geochemical
174^182. data from Permo -Triassic granitic gneisses and granitoids
Teixeira,W.,Tassinari, C.C.G., Cordani, U.G. & Kawashita, of the Colombian central Andes. J. Soc. Am. Earth Sci., 21,
K. (1989) A review of the geochronology of the Amazonian 355^371.
craton -Tectonic implications. Precambrian Res., 42, 213^227. Warren, E.A. & Pulham, A.J. (2002) Anomalous porosity and
Ulloa, C., Rodriguez, E., Fuquen, J. & Acosta, J. (2001a) permeability preservation in deeply buried Tertiary and Me-
Geologia de la plancha 192 Laguna de Tota, INGEOMINAS. sozoic sandstones in the Cusiana eld, Llanos foothills, Co -
Bogota, Map with Report. lombia - reply. J. Sediment. Res., 72, 445^448.
Ulloa, C.E., Rodriguez, E. & Rodriiguez, G.I. (2001b) Geo - Wijninga,V.M. (1996a) Neogene ecology of the Salto deTequen-
logia de la plancha 172 Paz de Rio, INGEOMINAS. Bogota, dama site (2475 m altitude, Cordillera Oriental, Colombia):The
Map with Report. paleobotanical record of montane and lowland forests. Rev.
van der Wiel, A.M. (1991) Uplift and volcanism of the SE Colom- Palaeobot. Palynol., 92, 97^156.
bian Andes in relation to Neogene sedimentation of the Upper Wijninga,V.M. (1996b) Palynology and paleobotany of the early
MagdalenaValley. PhD thesis, Free University Wageningten. Pliocene section Rio Frio 17 (Cordillera Oriental, Colombia):
Van Hinte, J.E. (1978) Geo -history analysis: Application of mi- Biostratigraphical and chronostratigraphical implications.
cropaleontology in exploration geology. Am. Assoc. Petrol. Geol. Rev. Palaeobot. Palynol., 92, 329^350.
Bull., 62, 201^222. Zapata, T.R. & Allmendinger, R.W. (1996) Growth stratal re-
Villago mez, D. (2010) Thermochronology, Geochronology and cords of instantaneous and progressive limb rotation in the
Geochemistry of the Western and Central Cordilleras and precordillera thrust belt and Bermejo basin, Argentina. Tec-
Sierra Nevada de Santa Marta, Colombia:TheTectonic Evolu- tonics, 15, 1065^1083.
tion of NW South America. PhD thesis, Universite de Gene' ve,
Gene' ve.
Villago mez, D., Spikings, R., Seward, D., Magna, T. &
Winkler, W. (2008) New thermochronological constraints Manuscript received 26 April 2010; in revised form 1 October
on the tectonic history of western Colombia. In: The 11th Inter- 2010; Manuscript accepted 4 October 2010.

r 2010 The Authors


402 Basin Research r 2010 Blackwell Publishing Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists

You might also like