ArticlePDF Available

Neotectonic basin and landscape evolution in the Eastern Cordillera of NW Argentina, Humahuaca Basin (~24°S)

Authors:

Abstract

Intermontane basins are important repositories for sediment storage, documenting the tectonic, climatic, and sedimentary evolution of orogens in time and space. The Quebrada de Humahuaca in the Southern Central Andes of NW Argentina is an ideal setting to study the evolution of an intermontane depositional environment, which is located within a very narrow section of the Eastern Cordillera between 23-24°S. Ongoing shortening, associated surface uplift and progressive aridification, due to orographic barrier uplift, have helped accommodate and retain a variety of late Miocene to Quaternary sedimentary deposits in the Humahuaca basin. These deposits record the transition from an originally open foreland to an intermontane basin depositional environment. Here, we present new U-Pb zircon ages of intercalated volcanic ashes and paleocurrent data, suggesting that prior to ~4.3 Ma, the Qda. de Humahuaca was an integral part of a contiguous foreland basin, whose fluvial network was subsequently decoupled from the foreland due to basement-block uplifts in the course of an eastward-migrating deformation front. Subsequently, the basin experienced at least three cycles of hydrologic isolation and filling followed by headward erosion, renewed fluvial connectivity and erosion, respectively. The ultimate basin-filling episode between ~800 ka and
Neotectonic basin and landscape evolution in the
Eastern Cordillera of NW Argentina, Humahuaca
Basin (~24°S)
Heiko Pingel,*Manfred R. Strecker,*Ricardo N. Alonsoand Axel K. Schmitt
*DFG Leibniz Center for Surface Process and Climate Studies, Institut f
ur Erd- und Umweltwissenschaften,
Universit
at Potsdam, Potsdam, Germany
Departamento de Geolog
ıa, Universidad Nacional de Salta, Salta, Argentina
Department of Earth and Space Sciences, UC Los Angeles, Los Angeles, CA, USA
ABSTRACT
The intermontane Quebrada de Humahuaca Basin (Humahuaca Basin) in the Eastern Cordillera of
the southern Central Andes of NW Argentina (23°24°S) records the evolution of a formerly contig-
uous foreland-basin setting to an intermontane depositional environment during the late stages of
Cenozoic Andean mountain building. This basin has been and continues to be subject to shortening
and surface uplift, which has resulted in the establishment of an orographic barrier for easterly
sourced moisture-bearing winds along its eastern margin, followed by leeward aridification. We pres-
ent new UPb zircon ages and palaeocurrent reconstructions suggesting that from at least 6 Ma until
4.2 Ma, the Humahuaca Basin was an integral part of a largely contiguous depositional system that
became progressively decoupled from the foreland as deformation migrated eastward. The Humahu-
aca Basin experienced multiple cycles of severed hydrological conditions and subsequent re-captured
drainage, fluvial connectivity with the foreland and sediment evacuation. Depositional and structural
relationships among faults, regional unconformities and deformed landforms reveal a general pattern
of intrabasin deformation that appears to be associated with different cycles of alluviation and basin
excavation in which deformation is focused on basin-internal structures during or subsequent to
phases of large-scale sediment removal.
INTRODUCTION
To understand the spatiotemporal evolution of tectoni-
cally active range fronts in mountain belts, it is essen-
tial to unravel the relationships between styles and
rates of tectonic deformation, surface uplift, and the
distribution of precipitation and surface processes that
reflect relief and local climatic conditions. This evolu-
tion may be partially recorded by the sedimentary
deposits preserved in intermontane basins in the
peripheral sectors of an orogen, in the compartmental-
ized basins within broken forelands, or farther away in
adjacent foreland basins.
Sediment accumulation within contiguous foreland
basins is predominantly determined by the flexural
response of the crust to the topographic load from an
adjacent fold-and-thrust belt and from the sediments
derived from the orogen (e.g. Beaumont, 1981; De-
Celles & Giles, 1996). As a fundamental characteristic
of foreland basins, protracted deformation and coeval
deposition progressively extend into the previously
undeformed, distal foreland regions, with the orogenic
deformation front and associated sedimentary facies
patterns advancing in a systematic spatiotemporal pat-
tern (e.g. DeCelles & Giles, 1996). In contrast, broken
foreland basins may develop in regions where shorten-
ing is accommodated along reactivated high-angle struc-
tures inherited from former tectonic regimes (Jordan &
Allmendinger, 1986; Jordan & Alonso, 1987), often
leading to highly diachronous and spatially disparate
basement uplifts (reviewed in Strecker et al., 2011).
Isolated range uplifts promote much more subdued
flexural subsidence with accommodation space that is
limited to the margins of the individual ranges (Strec-
ker et al., 2011). If these tectonic characteristics are
paired with arid climate conditions, headwater basins
can become isolated from the downstream fluvial net-
work, leading to sediment accumulation between uplift-
ing ranges (Meyer et al., 1998; Sobel et al., 2003;
Hilley & Strecker, 2005). Over geological timescales, an
array of isolated to variably connected and laterally
restricted depocentres may develop, forming a land-
scape similar to the partially coalesced basins observed
in Cenozoic orogenic plateaus and their flanks, such as
the southern part of the Andean Altiplano-Puna Plateau
or the Tibetan Plateau and the adjacent Qilian Shan in
Asia (Meyer et al., 1998; Sobel et al., 2003; Alonso
et al., 2006; Strecker et al., 2009).
Correspondence: Heiko Pingel, Institut fu
¨r Erd- und Umwelt-
wissenschaften, Universita
¨t Potsdam, Karl Liebnecht-Str. 24,
D-14476 Potsdam-Golm, Germany E-mail: heiko.pingel@geo.
uni-potsdam.de
©2012 The Authors
Basin Research ©2012 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
554
Basin Research (2013) 25,554–573, doi: 10.1111/bre.12016
EAGE
The dynamics of basin hydrology in these environ-
ments may depend on a type of competition between
uplift of downstream ranges, which favours fragmentation
of the fluvial network, and headward incision, which pro-
motes the persistence or re-integration of rivers draining
the periphery of these orogens (Sobel et al., 2003; Hilley
& Strecker, 2005; Garcia-Castellanos, 2007). As a result,
changes in climatic conditions or tectonic rates may cause
these marginal basins along the flanks of an orogenic pla-
teau to transition between conditions in which fluvial con-
nectivity is promoted or basins are hydrologically isolated
from their downstream watersheds. These alternations in
the fluvial network are expected to influence the rate and
tempo of sediment removal from the interior of the oro-
gen to the unrestricted parts of foreland basins.
The Andean broken foreland areas of the northern
Sierras Pampeanas, the Santa Barbara System and parts
of the Eastern Cordillera of NW Argentina (Fig. 1)
illustrate the complex morphology resulting from tec-
tonic uplift, basin formation and basin excavation along
the eastern flanks of the orogenic Altiplano-Puna Pla-
teau. While field studies in this region show that rivers
connecting intermontane basins with the foreland are
often interrupted due to spatial and temporal changes
in deformation, climate and the erodibility of exposed
bedrock (e.g. Hilley & Strecker, 2005), predicting the
combinations of driving factors responsible for alternat-
ing states of basin isolation and fluvial connectivity is
still difficult. Deposits preserved in the intermontane
basins are often deformed, faulted and frequently show
syntectonic growth as a result of initial foreland frag-
mentation and out-of-sequence deformation, demon-
strating clear tectonic influence on basin sedimentation
(for a summary see Strecker et al., 2011; Hilley et al.,
2005). Moreover, regional unconformities show that
large volumes of sediment have been rapidly removed
from these basins once they have been re-captured
(Hilley & Strecker, 2005; Strecker et al., 2009). These
superposed processes raise an interesting and as yet
unexplored possible feedback between the re-integration
of these intermontane basins with the foreland through
fluvial connectivity and renewed faulting within the
orogenic realm following the removal of sedimentary
loads. While these relationships among tectonics,
sedimentation and erosion have been suspected to exist
in many intermontane basins of the NW Argentine
Andes (Strecker et al., 1989, 2009; Hilley & Strecker,
2005; Alonso et al., 2006), the timescales for individual
filling and excavation cycles have remained poorly con-
strained.
A rich record of frequently intercalated volcanic ashes
in the sediments of the southern Humahuaca Basin of
NW Argentina (~23.5°S, Fig. 1) provides the requisite
chronology to quantify temporal associations among tec-
tonics, climate and sedimentation on the scale of a single
intermontane basin. In our study, we present new chrono-
stratigraphic (
206
Pb/
238
U zircon and AMS
14
C), struc-
tural and sedimentological data for various preserved
conglomeratic basin fills within the Humahuaca Basin and
document that (a) the transition between a largely contin-
uous foreland depositional environment and a subsequent
fault-bounded intermontane basin is related to more pro-
nounced surface uplifts to the east after ~4.2 Ma; (b) the
intermontane basin stage has been characterized by multi-
ple cycles of basin filling and subsequent sediment
removal; and, although speculative, (c) out-of-sequence
reactivation of faults within the basin may be closely
linked with sediment evacuation.
REGIONAL AND GEOLOGICAL SETTING
The Humahuaca Basin (Jujuy Province) is the northern-
most intermontane sedimentary basin in an array of
reverse-fault bounded basins within the Eastern Cordil-
lera of NW Argentina along the eastern Puna Plateau
margin, the southern extension of the Bolivian Altiplano
(Fig. 1). The Humahuaca Basin is surrounded by high-
elevation mountain ranges exceeding 5,000 m a.s.l. that
consist of smaller reverse and thrust fault-bounded
blocks. The Sierra Alta separates the basin from the inter-
nally drained and arid Puna Plateau to the west, while the
Tilcara ranges constitute the boundary with the humid
foreland depositional system to the east. At present, the
Humahuaca Basin is connected to the foreland via a nar-
row, fault-bounded bedrock gorge to the south, through
which the Rıo Grande exits the basin (Figs 2 and 9).
Here, the course of the Rıo Grande firstly follows and
then obliquely crosses the trace of the west-dipping
reverse fault that bounds the southern sector of the Til-
cara ranges.
Basement blocks constitute the basin-bounding ranges
that have been uplifted along north to northnortheast
striking, bivergent thrust and reverse-fault systems
(Rodrıguez-Fernandez et al., 1999; Kley et al., 2005;
Fig. 2). At the latitude of the Humahuaca Basin, the east-
ern margin of the neighbouring Altiplano-Puna Plateau
records a middle Eocene to Oligocene deformation his-
tory, influenced by pre-existing crustal heterogeneities
and structures that were reactivated during Cenozoic
compression (Coutand et al., 2001, 2006; Deeken et al.,
2006; Hongn et al., 2007; Insel et al., 2012). Between 10
and 8 Ma, the eastern plateau margin apparently attained
sufficient elevation and relief to intercept moisture-bear-
ing easterly winds; this topography constituted a major
orographic barrier to atmospheric circulation on a hemi-
spheric scale, resulting in the aridification of the orogen
interior and the establishment of humid conditions on the
eastern flanks (Allmendinger et al., 1997; Kleinert &
Strecker, 2001; Starck & Anzotegui, 2001; Strecker et al.,
2007; Uba et al., 2007; Carrapa et al., 2008; Mulch et al.,
2010; Vezzoli et al., 2012). During the Mio-Pliocene,
deformation migrated into the present-day Eastern Cor-
dillera, where the formerly contiguous foreland basin was
partitioned by regional range uplifts. This deformation is
spatially disparate, highly diachronous (reviewed in
©2012 The Authors
Basin Research ©2012 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 555
Basin & landscape evolution in the E Cordillera
Strecker et al., 2007, 2009) and ongoing (Bevis & Isacks,
1984; Cahill et al., 1992; USGS/NEIC PDE-catalogue).
The basement rocks exposed along the flanks of the
Humahuaca Basin comprise tightly folded late Protero-
zoic to early Palaeozoic low-grade metasediments of the
Puncoviscana Formation (Turner, 1960; Omarini, 1983).
These units are unconformably overlain by Cambro-
Ordovician sandstones and quartzites of the Meson and
Santa Victoria groups (Moya, 1988; Sanchez & Salfity,
1999; Ace~nolaza, 2003). An angular unconformity sepa-
rates these sediments from the late Cretaceous to
Palaeogene Salta Group related to the Cretaceous Sal-
ta Rift (Salfity, 1982; Galliski & Viramonte, 1988; Mar-
quillas et al., 2005). The most prominent strata of these
sequences exposed in the southern Humahuaca Basin are
continental red beds of the Pirgua Subgroup, white sand-
stones and yellow-coloured marine carbonates of the
Lecho Formation and the stromatolitic Yacoraite Forma-
tion (Balbuena Subgroup), respectively, and fluvial
deposits of the Lumbrera Formation (Santa Barbara Sub-
group). For detailed reviews, see Marquillas et al. (2005)
and Sanchez & Marquillas (2010).
These lithologies are typically overlain by early Ceno-
zoic foreland sediments such as the Quebrada de los Col-
orados Formation (middle EoceneOligocene), the
deposits of the Oran Group (MiocenePliocene) or equiv-
alent strata (Gebhard et al., 1974; Russo & Serraiotto,
1978; Dıaz & Malizzia, 1983; Vergani & Starck, 1989;
Coutand et al., 2001) in the Puna and present-day
foreland regions to the east of the study area. However,
these sediments have mostly been removed in the highly
exhumed Eastern Cordillera (Jordan & Alonso, 1987;
Kley et al., 2005). A regional exception to this general
pattern is a ca. 6-km thick succession of middle Eocene to
Pliocene foreland and intermontane basin deposits in the
Cianzo Basin of the Eastern Cordillera, 20 km east of the
town of Humahuaca (details in Siks & Horton, 2011).
Strata overlying the Salta Group in the Humahuaca Basin
largely consist of weakly consolidated, mainly
conglomeratic deposits that reflect a complex history of
deposition, erosion and deformation that spans the late
Miocene and Quaternary.
Three major units have previously been described:
the Maimara Formation (Salfity et al., 1984), the Uquıa
Formation (Castellanos, 1950; Marshall et al., 1982;
Walther et al., 1998) and thick conglomeratic fills of
Quaternary age (Tchilinguirian & Pereyra, 2001;
Robinson et al., 2005; Strecker et al., 2007; Sancho
et al., 2008). In the following sections, we will refine
this stratigraphic framework for the southern part of
the Humahuaca Basin and focus our attention on its
distinct volcanic ash-bearing conglomerates and sand-
stones that document sustained deposition, deformation
and erosion in the basin, and which provide excellent
stratigraphic markers to assess the late Cenozoic basin
evolution.
METHODS
We used stratigraphic and structural analysis, together
with detailed geological mapping of exposed units
(Fig. 3), regional unconformities, sediment provenance,
lateral facies pinch-outs and lithological contacts to docu-
ment the tectono-sedimentary history of the southern
Humahuaca Basin. A similar approach is used in an
ongoing chronostratigraphic study of the northern Huma-
huaca Basin (Streit et al., 2012).
Fig. 1. Topography and east-draining river network of the
South Central Andes in NW Argentina and S Bolivia based on
SRTM-GTOPO30. Structural information is taken from Sobel
et al. (2003), Mortimer et al. (2007), Hain et al. (2011), Mulch
et al. (2010), Uba et al. (2007), Allmendinger & Zapata (2000),
Kley et al. (1997) and Carrapa et al. (2006). Dashed lines delin-
eate morphotectonic domains. Letters relate to major intermon-
tane basins in NW Argentina discussed in the text: H, Quebrada
de Humahuaca; T, Quebrada del Toro; L, Lerma Basin; C,
Calchaquı Basin; SM, Santa Marıa Basin; CA, El Cajon-Campo
Arenal Basin; F, Bolson de Fiambala Basin. Black box indicates
location of Fig. 2.
©2012 The Authors
Basin Research ©2012 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
556
H. Pingel et al.
Despite apparent similarities between the various sy-
norogenic lithologies, spatiotemporal changes in the
sediment sources result in distinct differences among
the conglomeratic fill units in the Humahuaca Basin.
We characterized the compositional differences of the
fills by counting at least 100 clasts from within a 0.25-
m
2
grid in key stratigraphic units. To deduce sediment
provenance and transport directions for ancient river
systems, we measured ~1,600 imbricated clasts at 33
localities. Where possible, we measured the orientation
of at least 50 clasts per site, applied corrections for
structural dip and displayed them in unidirectional rose
diagrams using OSXStereonet software (by N. Cardozo
& R. Allmendinger).
(a)
(b)
Fig. 2. (a) Simplified geology of the Humahuaca Basin and surroundings and (b) geological cross section after Rodrıguez-Fernandez
et al. (1999) and own data. Black box indicates the area mapped in detail (Fig. 3a).
©2012 The Authors
Basin Research ©2012 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 557
Basin & landscape evolution in the E Cordillera
(a) (b)
(c)
Fig. 3. (a) Geological map of the central study area between Tilcara and Purmamarca in the southern Humahuaca Basin. Triangles
represent UPb zircon sample locations (see Table S1) and thick black lines show the position of measured stratigraphic sections from
Fig. 6. Map is rotated anti-clockwise by 20°. (b) Simplified late Cenozoic chronostratigraphy of the study area. Unless otherwise
indicated, values represent averaged UPb zircon ages from this study.
a40
Ar/
39
Ar-biotite (Strecker et al., 2007);
b
OSL-quartz
(Robinson et al., 2005; Sancho et al., 2008);
c
AMS14C (this study). (c) Subsurface interpretations of severely deformed strata. Shown
are pseudo-fault-plane solutions calculated from fault-kinematic indicators documenting thrust kinematics during the Pliocene. Pv,
Puncoviscana Fm; Sa, Salta Group; M, Maimara Fm; T, Tilcara Fm; Q, Quaternary gravels.
©2012 The Authors
Basin Research ©2012 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
558
H. Pingel et al.
To provide a chronological base for different tectonic
and sedimentological events, we dated 12 volcanic ash
deposits interbedded in the Mio-Pleistocene basin strata
using UPb zircon geochronology. Samples were
crushed, sieved and treated with standard heavy-liquid
and magnetic separation techniques to isolate zircon
crystals. About 30 crystals per sample were handpicked,
mounted in epoxy, polished and cleaned, and then
gold-coated for microprobe analysis. Crystals free of
inclusions, or cracks were selected for UPb analysis
using the CAMECA IMS 1270 ion microprobe at the
University of California in Los Angeles, following pro-
tocols described in Schmitt et al. (2003) and Grove
et al. (2003). The
206
Pb/
238
U ages have been corrected
for common Pb and initial disequilibrium. The uncer-
tainties in UPb ages, estimated from the reproducibil-
ity of standard AS3 zircons (1,099.1 Ma; Paces &
Miller, 1993), were 2.2% and 2.7% (1 standard devia-
tion) for the analytical sessions in July 2009 and June
2010, respectively.
RESULTS
UPb zircon geochronology
Most analysed samples show complex zircon age distribu-
tions. This could be due to protracted pre-eruptive crystal
residences (e.g. Schmitt et al., 2003), or post-eruptive
reworking in which case mixing and contamination with
detrital crystals during emplacement would lead to the
presence of multiple age populations. We therefore sys-
tematically omitted older ages from our calculations of an
average zircon crystallization age. The statistically uni-
form younger age population was then used as an approxi-
mation for the depositional age, while acknowledging that
this is likely to overestimate the eruption age because of
pre-eruptive zircon crystallization (e.g. by ~0.1 Ma for
the large-volume Atana ignimbrite; Schmitt et al., 2001).
Most samples yielded consistent
206
Pb/
238
U ages for the
majority of crystals, as indicated by near-unity values for
the mean square of weighted deviates (MSWDs), suggest-
ing only minor reworking. In some cases, however, only a
small percentage of crystals defined a coherent young
population; in these cases, we have interpreted the
206
Pb/
238
U zircon age as the maximum age for deposition.
Results are shown in Figs 35, and summarized in Table
S1.
Late Miocene to Pleistocene stratigraphy
Maimara Formation
The ochre to yellow beds of the Maimara Formation
unconformably overlie the older lithologies exposed in
the basin, including the Proterozoic Puncoviscana For-
mation. The Maimara Formation generally comprises
arkosic sandstones and interbedded cobble conglomer-
ates, and is at least 250-m thick (Fig. 6). The matrix-
(a) (b) (c)
(d) (e) (f)
Fig. 4.
207
Pb/
206
Pb vs.
238
U/
206
Pb zircon data for Maimara samples, uncorrected for common Pb and regression lines with a fixed
y-axis intercept corresponding to common Pb (
207
Pb/
206
Pb =0.83). Concordia segment (ages in Ma) is plotted for initial disequilib-
rium D
230
Th/
238
U=0.2 and D
231
Pa/
235
U=3. Number of analyses excluded from regression is given in parentheses.
©2012 The Authors
Basin Research ©2012 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 559
Basin & landscape evolution in the E Cordillera
to clast-supported conglomerates are composed of well-
rounded pebbles and cobbles, and occasionally boul-
ders, mainly from Proterozoic (23.5% P
V
) and Palaeo-
zoic (64.5% P
Z
) sources (Puncoviscana Fm. & Meson
Group). This unit also contains clasts of limestones
and sandstones from the Cretaceous to Palaeogene Salta
Group (12% C
R
) and a minor proportion of lithic-rich,
late Miocene to Pliocene ignimbrites with the nearest
known exposures confined to the Puna Plateau to the
west (e.g. Riller et al., 2001).
The most complete section is exposed in the Quebrada
de Maimara, west of the town of Maimara (Figs 3 and 6),
where the succession has been thrust eastward over Plio-
cene conglomerates. Fossil-rich clay beds and siltstones
dominate the basal 50 m of this section and contain
freshwater ostracods (Limnocythere sp.; Fig. 8b) and intact
calcic encrustations of charophyte oogonia. The following
200-m thick sequence of interbedded sandstones and
conglomerates is intercalated with at least seven volcanic
ash layers.
About 75 m of faulted strata of the Maimara Formation
is also exposed at Incahuasi, located about 10 km south of
the Quebrada de Maimara. This section unconformably
overlies palaeo-relief developed in the Proterozoic to Pal-
aeozoic basement and comprises several metre-thick
banks of moderately consolidated fine-grained arkosic
sandstones that are frequently intercalated with rhyolitic
ash layers, debris-flow deposits and conglomeratic
channel fills (Fig. 6). Further exposures of the Maimara
Formation exist east of the town of Tilcara, along the Rıo
Huasamayo, where the Maimara
´Formation has been
thrust over late Pleistocene conglomerates (e.g. Salfity
et al., 1984; Marrett et al., 1994; Fig. 8a).
The presence of ignimbrite clasts sourced in the Puna
supports the notion of an eastward fluvial transport across
the present-day Sierra Alta, the major mountain range
that now constitutes the eastern margin of the Altiplano-
Puna Plateau west of the basin. The notion of a western
provenance is in agreement with our palaeocurrent esti-
mates that demonstrate an eastsoutheast-directed palae-
o-drainage system at that time (Fig. 7). We therefore
conclude that the former fluvial network must have
drained eastward across both ranges that now delimit the
basin.
Our UPb zircon ages from various volcanic ashes
(08HUM03; 08HUM05; 08HUM07; 09HUM12;
10HUM02; 10HUM21; 10HUM23; Figs 4 and 6; Table
S1) constrain that this depositional setting existed at least
between 5.92 0.12 Ma (MSWD =1.4; n=14) and
4.18 0.11 Ma (MSWD =0.68; n=10).
Tilcara Formation
In the southern Humahuaca Basin, the Maimara Forma-
tion is overlain by a series of metre-thick interbedded
conglomerate, fanglomerate and sandstone beds, at least
250-m thick, which also contain volcanic ash layers. The
transition to subsequent lithologies is always character-
ized by a pronounced regional unconformity, rendering
all measurements of total sediment thickness minimum
estimates (Figs 8c, d). In contrast to the Maimara Forma-
tion, the poorly consolidated strata comprise well-
rounded and well-imbricated pebble- to boulder-sized
clasts with only a minor quantity of Salta Group clasts,
(a)
(b)
(c)
(d)
Fig. 5.
207
Pb/
206
Pb vs.
238
U/
206
Pb zircon data for Tilcara and
Quaternary gravel samples, uncorrected for common Pb and
regression lines with a fixed y-axis intercept corresponding to
common Pb (
207
Pb/
206
Pb =0.83). See caption of Fig. 4 for addi-
tional information.
©2012 The Authors
Basin Research ©2012 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
560
H. Pingel et al.
but no ignimbrites (27.5% P
V
; 71% P
Z
; 1.5% C
R
).
Carbonate cementation of conglomerates occurs in metre-
thick beds. The nearly filled pore spaces, within such
beds, suggest advanced pedogenic K-horizon formation
(e.g. Gile et al., 1966; Machette, 1985), a common feature
associated with conglomerate deposits in the semi-arid
environments of the southern Central Andes (e.g. Strec-
ker et al., 1989). Clast-imbrication measurements record
a change to a southerly direction of sediment transport
(Figs 6 and 7). While on average, this change demon-
strates a rotation of ca. 20°towards the south (Fig. 7),
palaeocurrent directions measured along continuous sec-
tions reflect a dramatic reorganization of the fluvial system
by more than 90°(Fig. 6).
Two ashes sampled in the lower part of the section
(08HUM01 and 08HUM08) yielded overlapping UPb
zircon ages of 3.66 0.20 (MSWD =1.6; n=6) and
3.52 0.08 Ma (MSWD =0.44; n=8; Fig. 5; Table
S1). Because these age determinations are statistically
indistinguishable, we consider them to represent the same
ash horizon. A second ash layer from the upper section
(09HUM05) yielded a significantly younger
206
Pb/
238
U
zircon age of 2.50 0.10 Ma (MSWD =1.4; n=9;
Fig. 5; Table S1).
Previous stratigraphic and palaeontological studies
have shown that in the northern Humahuaca Basin, strata
of the Maimara Formation are overlain by the fossil-bear-
ing fluvial Uquıa Formation. This unit comprises mud
and sandstones with occasional conglomeratic beds and
interbedded volcanic ash horizons (Castellanos, 1950;
Reguero et al., 2007). Clast counts for the Uquıa Forma-
tion at a limited number of outcrops reveal ~10% Prote-
rozoic rocks (Puncoviscana Fm.) and ~90% Palaeozoic
rocks (Meson Group). Age determinations of a basal vol-
canic ash from the Uquıa Formation (3.54 0.04 Ma;
Marshall et al., 1982) and from within the upper third of
the section (zircon fission-track age ~2.5 Ma; Walther
et al., 1998) are in good agreement with our chronology
of the Tilcara Formation. Furthermore, palaeomagnetic
results from the Uquıa Formation (Marshall et al., 1982)
imply that its uppermost strata may be as young as
~1.5 Ma. As the topmost section of the Tilcara Formation
is not preserved, we are unable to determine an upper
depositional age limit. However, we infer that the timing
of deposition was similar in both sub-basins. We consider
this assumption to be valid because (a) we do not find any
evidence of other deposits between 2.5 and >1 Ma in the
southern Humahuaca Basin, and (b) the oldest
Quaternary sediments that cover both formations uni-
formly may be as old as ~1 Ma. It is therefore quite possi-
ble that the (unpreserved) top of the Tilcara Formation in
the southern basin is temporally equivalent to the ~1.5-
Myr old top of the Uquıa Formation to the north.
Although chronostratigraphic investigations in the
northern Humahuaca Basin are still ongoing (Streit et al.
2012), we have sufficient evidence for lithological differ-
ences between the corresponding Plio-Pleistocene depos-
its in the two sub-basins to establish a new lithological
unit in the southern basin: the Tilcara Formation. Our
radiometric ages show that deposition of this unit took
place between ~4.2 and 2.5 Ma and by correlation may
have lasted until approximately 1.5 Ma.
Landslide deposits
We identified multiple voluminous landslide deposits
south of Maimara village that unconformably cover older
units along the eastern basin margin (Fig. 3a). These
deposits predominantly consist of Palaeozoic and Creta-
ceous to Eocene rocks, and overlie previously exhumed
Palaeozoic rocks that dip steeply westward. Multiple, per-
vasively shattered rock sheets with no stratigraphic con-
text are located 60 m above the valley floor and appear to
be sourced in the eastern basin-bounding range. In places,
the landslide deposits have been subsequently covered by
sub-horizontally bedded conglomerates of Pleistocene
age. The depositional age of the landslide deposits can,
therefore, be only crudely constrained to be older than the
Pleistocene conglomerates.
Fig. 6. Measured stratigraphic sections located near Incahuasi
and in the Quebrada de Maimara. Detailed locations are given in
Fig. 3a.
©2012 The Authors
Basin Research ©2012 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 561
Basin & landscape evolution in the E Cordillera
To the west of the Rıo Grande and north of Incahuasi
(Fig. 3a), a voluminous Quaternary landslide deposit that
extends for approximately 3.5 km to the north has been
preserved covering an erosional palaeo-landscape in the
previously deposited and folded units. The deposit com-
prises two distinct source lithologies: (a) basal conglomer-
ates whose clast composition, size and general appearance
match the conglomeratic sections of the Maimara Forma-
tion; and (b) Proterozoic basement, found only in the
upper section of the landslide. In both parts, the rocks are
heavily sheared and fractured, with the degree of
cataclasis increasing with depth, culminating at the bot-
tom where underlying undifferentiated sediments have
been injected upward into the fully disintegrated rocks.
The contact between the two units comprising the land-
slide deposit is very sharp and resembles a thrust fault
identical to the relationships that can be observed along
the present-day basin-bounding Sierra Alta to the west.
The landslide deposit is covered by ash-bearing fanglom-
erates that have been dated to ~1 Ma (see section below),
which were subsequently tilted by faulting.
Quaternary gravels
The youngest deposits in the southern Humahuaca Basin
constitute thick gravel fills covering palaeotopography in
the previously deformed and eroded units. Two lithologi-
cal units can be distinguished on the basis of clast compo-
sitions. The first unit comprises dark grey to black
fanglomerates that were derived exclusively from source
regions to the west. These sediments consist predomi-
nantly of angular to subangular clasts of the Puncoviscana
Formation, together with less abundant clasts from the
Meson and Salta groups. We thus interpret these strata to
be proximal alluvial-fan deposits sourced from the Sierra
Alta. At several locations corresponding to more distal
sectors of the inferred alluvial fans, the fanglomerates in-
terfinger with well-stratified layers of pebble conglomer-
ates, graded sands and unconsolidated silty clay. The
layers are characterized by lateral, eastwest-oriented
pinch-outs. By analogy with the present-day depositional
environment of the Rıo Grande and from the geometry of
the pinch-outs, we infer that the former fluvial system
also drained southward. The youngest zircons from a tra-
chy-dacitic ash deposit in the upper part of a deformed
succession 2 km northwest of Incahuasi yielded mid-
Pleistocene ages (1.06 0.10 Ma; n=2; 08HUM11;
Fig. 5; Table S1), which we infer to represent a maxi-
mum depositional age.
The second unit in the Quaternary gravels comprises
a group of grey alluvial-fan deposits and fluvial conglom-
erates (Fig. 8e) that are widely distributed within the
Humahuaca Basin, its tributary valleys and in the basin
outlet region to the north of Volcan village (Figs 2 and
9). These gravels typically consist of approximately
equal proportions of Proterozoic and Palaeozoic clasts
with minor contributions from Cretaceous lithologies,
representing the present-day exposure of rock types in
the surrounding source areas. The oldest conglomeratic
fill in this unit forms an abandoned geomorphic surface
to the east of the town of Tilcara, which is up to 400 m
above the present-day baselevel and has an
40
Ar/
39
Ar-
biotite age of ~800 ka, taken from a volcanic ash layer in
the lower third of the section (Strecker et al., 2007).
Other investigators have further differentiated these
gravels and document at least one additional, separate fill
unit at lower elevations, corresponding to a third basin-
filling episode between 93.8 7.9 and 65 4ka
(Tchilinguirian & Pereyra, 2001; Robinson et al., 2005;
Sancho et al., 2008). In the south of Tilcara, this younger
fill unit has been episodically downcut during the last
~65 ka, which has resulted in fluvial terrace surfaces at
successively lower elevations (Fig. 8f). In the tributary
Quebrada de Purmamarca in the southwest of the study
area, even younger deposits, up to 250-m thick, constitute
massive basin fills that have been dated 47.6 2.8 ka
(OSL, Robinson et al., 2005) and, although at the limit of
the dating method, at 49.55 1.7 ka BP (23°40.9S, 65°
34.5W; AMS
14
C this study).
Structures
Three east-vergent, basement-involved fault systems
define the structural framework of the Humahuaca Basin:
(a) the basin-bounding Purmamarca Thrust Fault to the
west; (b) a set of thrust-and-reverse faults within the Til-
cara ranges to the east; and (c) the Tumbaya Fault in the
basin centre, close to the Rıo Grande (Fig. 9).
The Purmamarca Thrust Fault juxtaposes Proterozoic
basement of the Sierra Alta over Meso-Cenozoic sedi-
ments along the western basin margin and has developed
a characteristic deformation pattern in the overthrust
lithologies, involving steep eastward dips, or overturned
strata. Moreover, multiple splays have developed from
that fault extending into the basin, offsetting the Cenozoic
strata (Figs 2 and 3). The Sierra Alta belongs to a set of
subparallel basement ranges that have been uplifted along
bivergent thrust and reverse faults. Apatite fission track
Fig. 7. Palaeoflow directions in the Humahuaca Basin derived from clast-imbrications measured in late Miocene to Pleistocene basin
strata. Directions are presented on unidirectional rose diagrams for each site. Panels (a), (b) and (c) show the development of the
fluvial network in the Maimara, Tilcara and Quaternary gravel formations respectively. Panel (b) also shows data from the Uquıa
Formation in the northern Humahuaca Basin. A combination of all clast imbrications for each unit is given in panel (d) highlighting a
significant change in transport directions after 4.2 Ma associated with changes in the topographic boundary conditions. Palaeoflow
directions in the Quaternary gravels are biased by a lack of properly distributed sample sites; locations are mainly along the eastern
margins of the basin, thus more prone to reflect westerly drainages. Results of provenance analyses are shown in histogram plots
depicting mean clast compositions of conglomeratic units.
©2012 The Authors
Basin Research ©2012 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
562
H. Pingel et al.
(a) (b)
(c) (d)
©2012 The Authors
Basin Research ©2012 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 563
Basin & landscape evolution in the E Cordillera
(a)
(b) (c)
(e)(d)
(f)
Fig. 8. (a) Thrust fault near Tilcara along the Rıo Huasamayo, juxtaposing late Miocene Maimara deposits against Quaternary
gravels. Arrows and dashed lines indicate the vertical displacement (up to 20 m) of a formerly contiguous terrace surface. (b) SEM
image of Limnocythere sp. from the lower beds of the Maimara Formation. (c) Deformed bedding-parallel erosion surfaces in the Til-
cara Formation preserved below Quaternary landslide deposits. Tectono-sedimentary relationships suggest an episode of river incision
and excavation prior to deformation. (d) Another example of the marked regional unconformity between the Tilcara Formation and
Quaternary gravels. (e) Thick conglomeratic fill unit (Quaternary gravels) near the town of Tumbaya. (f) Panoramic view towards the
east showing well developed geomorphic surfaces along the western flanks of the Tilcara ranges at successively lower elevation.
©2012 The Authors
Basin Research ©2012 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
564
H. Pingel et al.
thermal modelling suggests that exhumation of the east-
ernmost ranges constituting the Sierra Alta began
between 15 and 10 Ma (Deeken et al., 2004).
Present-day elevations of the eastern basin-bounding
Tilcara ranges are attributed to shortening, crustal thick-
ening and block rotation along multiple major east-verg-
ing thrust faults located within the range (Rodrıguez-
Fernandez et al., 1999; Fig. 2). Because fault activity
was mainly restricted to areas east of the Humahuaca
Basin, the Proterozoic basement and overlying Palaeozo-
ic to Mesozoic strata along the western flanks of the
range were affected by westward tilting of this basement
block.
At Molle Punco, near Tumbaya (Figs 2 and 9), Prote-
rozoic rocks of the Puncoviscana Formation are thrust
over the Palaeozoic successions of the Meson Group along
the Tumbaya Fault. The Tumbaya Fault can be traced
into the northern Humahuaca Basin, intersecting the
course of the Rıo Grande repeatedly. In the southern Hu-
mahuaca Basin, this fault is responsible for an uplifted
central range (Fig. 3a) that causes the narrowing of the
basin. The surface expression of the fault and associated
basement exposure gradually decrease northward, but the
fault location can still be inferred from west-dipping
Cenozoic basin strata in its hanging wall.
The sedimentary strata and landforms in the southern
Humahuaca Basin attest to protracted deformation during
the Plio-Pleistocene (see section below and Fig. 3c). The
deformation includes thrusting of Precambrian to
Mesozoic rocks over the conglomerates of the Maimara
Formation along the Purmamarca Thrust Fault and asso-
ciated folding, and the subsequent tilting of late Miocene
to Pleistocene basin sediments in the hanging wall of the
Tumbaya Fault until after 1 Ma.
Shortening within the basin was further accommodated
by a number of mesoscale structures, mostly affecting the
Maimara and Tilcara formations. Tight to open folding
and the development of generally east-vergent and shal-
low-dipping thrust systems led to multiple stacked repeti-
tions of the Miocene-Pliocene strata within the basin
(Fig. 3). These thrusts are often associated with the
development of drag folds, and shortening is in places
accommodated by antithetic west-vergent faults. This
pronounced Plio-Pleistocene strain accommodation in the
southern Humahuaca Basin is mainly observed in the west
of the Rıo Grande between the Purmamarca and
Tumbaya faults.
Fig. 10. Schematic cross section (~250 km) of the Andean margin in NW Argentina at ~23.5°S showing times of pronounced Ceno-
zoic deformation. (1) Letcher (2007); (2) Deeken et al. (2005); (3) A. Deeken personal communication (2012); (4) Deeken et al. (2004);
(5) Siks & Horton (2011); (6) Reynolds et al. (1994, 2000).
Fig. 9. Simplified structural map of the location and strike of
major fault systems responsible for the basin geometry. Indicated
are (a) the Purmamarca Thrust Fault, (b) the Tilcara Thrust
System and (c) the central Tumbaya Fault. Solid thick white line
represents the watershed between the internally drained Puna
Plateau and the presently externally drained Eastern Cordillera;
dashed lines delineate the Humahuaca Basin and adjacent sub-
basins (Tres Cruces and Cianzo basins); also shown is the
location of the Zapla anticline east of San Salvador de Jujuy.
©2012 The Authors
Basin Research ©2012 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 565
Basin & landscape evolution in the E Cordillera
Tectono-sedimentary relationships and
deformation of basin sediments
The Maimara Formation rests unconformably on
deformed sandstones of the Palaeogene Lumbrera Forma-
tion (upper Salta Group) and Proterozoic to Palaeozoic
basement rocks. This unconformity and onlap relation-
ships between the Maimara Formation and the underly-
ing, irregularly shaped basement surface suggests
deposition on palaeotopography that was sculpted into
these rocks prior to ~6 Ma. To date, no evidence has been
observed supporting an unconformable relationship
between the Maimara Formation and the intermontane
Tilcara Formation. The Tilcara Formation is always cut
by a marked regional unconformity that in many places
has been subsequently altered and obscured by continued
deformation, erosion and deposition. Lateral correlation
with the Uquıa Formation in the northern Humahuaca
Basin suggests that deformation following the initial inci-
sion occurred after ~1.5 Ma. Near the Tumbaya Fault at
Incahuasi (Figs 3a and 9), this deformed unconformity
resembles fluvial erosion surfaces sculpted into the
Tilcara Formation that are tilted at ~2030°W, parallel to
bedding (Fig. 8c). These unconformable relationships
and the vestiges of an erosional palaeotopography in the
Tertiary sedimentary rocks are well preserved under a
>1.06 0.1 Ma conglomeratic fill unit that dips 10°
15°W. Basal remnants of this earlier fill unit have a depo-
sitional age of about 0.8 Ma.
A subsequent conglomerate unit (~9465 ka at Tilcara
and <50 ka near Purmamarca) filled palaeotopography,
covering the channel of the former trunk stream, which is
now being re-excavated. This >200-m thick fill unit ter-
minates in a smooth terrace surface that is connected with
the mountain fronts. At Purmamarca, these gravels are
generally not affected by fault displacement or other
deformation. Near Tilcara, however, a prominent thrust
fault close to the eastern margin of the basin (Fig. 8a)
juxtaposes the Maimara Formation with young
terrace-forming sections of the Quaternary strata (Salfity
et al., 1984; Marrett et al., 1994). Here, deformation
resulted in vertical offsets between 15 and 20 m, which
translates into an average vertical displacement rate of
0.27 0.04 mm/a during the last 65 kyr. Horizontal
displacements of ~40 m (Sancho et al., 2008) suggest
shortening rates of 0.62 0.04 mm/a. Although these
rates are only approximations, they emphasize the impor-
tance of protracted tectonic activity in the Humahuaca
Basin within the interior of the orogen, which is compati-
ble with the characteristics of regional shallow crustal
seismicity (e.g. Bevis & Isacks, 1984; Cahill et al., 1992).
It is noteworthy that this youngest deformation followed a
major phase of gradual basin evacuation.
DISCUSSION
Foreland basin fragmentation and orographic
barrier development
Crustal deformation corresponding to the region of the
present-day interior of the Puna Plateau initiated in
Eocene to Oligocene time (e.g. Kraemer et al., 1999; Car-
rapa et al., 2005; Deeken et al., 2006; Hongn et al., 2007;
Letcher, 2007; Fig. 10). While individual ranges were
uplifted, thick synorogenic strata buried the region that
now constitutes the Eastern Cordillera and adjacent
regions to the east (Reynolds et al., 2001; Deeken et al.,
2006; DelPapa et al., in review). Today, these sediments
are only rarely preserved in the uplifted Eastern Cordil-
lera and the south-eastern flanks of the present-day Puna
Plateau (Jordan & Alonso, 1987; Bossi et al., 2001; Cout-
and et al., 2001; Kley et al., 2005; Mortimer et al., 2007)
and none appear to have been retained in the southern
Humahuaca Basin. A regional exception is a ca. 6-km
thick succession of middle Eocene to Pliocene foreland
and intermontane basin deposits in the Cianzo Basin
(Fig. 9), about 20 km east of the town of Humahuaca,
where deformation and severed drainage conditions have
been documented at ca. 10 Ma (Siks & Horton, 2011).
Similarly, the Tres Cruces Basin in the Puna (Fig. 9) to
the west has retained thick Cenozoic deposits (e.g. Boll &
Hernandez, 1986; Coutand et al., 2001). In both cases,
major reverse faults enclosing the basins have helped to
preserve the Cenozoic sedimentary record. These regional
relationships support the notion of widely distributed
early Tertiary sediments in the area of the Eastern Cordil-
lera and regions farther east. Their general absence in
these high-elevation sectors of the orogen thus suggests
their removal during regional exhumation in the realm of
the Eastern Cordillera at about 1510 Ma (Deeken et al.,
2004, 2006; Coutand et al., 2006; Siks & Horton, 2011).
The earliest synorogenic strata recognized in the south-
ern Humahuaca Basin belong to the Maimara Formation
and were deposited on a palaeotopography of exposed
Proterozoic and Palaeozoic basement. Deposition of Mai-
mara sediments after a prolonged period of exhumation
and deformation along the former orogenic flanks clearly
documents that by ca. 6 Ma (and possibly some time
(a)
(b)
Fig. 11. Conceptual model of foreland-basin fragmentation. (a)
The Maimara Formation is deposited into an exposed basement
palaeotopography in a largely continuous depositional system
since ca. 6 Ma. (b) Surface uplifts to the east led to re-routing of
the fluvial network and deposition of the Tilcara Formation into
an intermontane basin after 4.2 Ma.
©2012 The Authors
Basin Research ©2012 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
566
H. Pingel et al.
earlier), new topographic conditions had evolved that
would have promoted sedimentation in the study area.
Our palaeocurrent and provenance data including ignim-
brite clasts from currently isolated areas in the Puna inte-
rior and clasts from the Eastern Cordillera document that
these sediments were sourced from the west. As the
Humahuaca Basin and the Puna are not connected any-
more, this observation confirms that thrusting in the
Sierra Alta had not completely interrupted eastward-
draining rivers at about 6 Ma and that the final disruption
of the fluvial network must have occurred later.
We do not record any sedimentary evidence for uplift
of the Tilcara ranges at the time when the Maimara For-
mation was deposited. This suggests that the Humahuaca
Basin was part of an unrestricted foreland, and that sur-
face uplift of the eastern ranges started later, coupled with
the deposition of the Tilcara Formation. Alternatively,
one could envision a scenario with an eastward-directed
antecedent fluvial network that may have traversed the
uplifting Tilcara ranges and transported sediment
towards the regions farther east. In both settings, more
pronounced surface uplift in Plio-Pleistocene time would
have ultimately forced the full establishment of intermon-
tane basin conditions and the formation of an efficient
drainage divide, with rivers routing sediments towards
the south. Although not documented in this study, a lat-
eral correlation between the Maimara Formation and the
upper sections of the Oran Group east of the ranges can-
not be excluded. Sediments of the Oran Group may yield
additional information to further elaborate these interpre-
tations in future.
Nevertheless, Oran Group sediments exposed in the
Zapla anticline to the south-east of the Tilcara ranges
(Fig. 9), record a major pulse in deformation prior to
5 Ma (Reynolds et al., 1994, 2000; reviewed in Kley &
Monaldi, 2002). This is consistent with our palaeocurrent
data from the Tilcara Formation that record a distinct
early Pliocene changeover of fluvial transport towards a
northsouth-oriented drainage system, which is also doc-
umented in the Uquıa Formation of the central and
northern Humahuaca Basin (Fig. 7). Based on the cur-
rently available data, we conclude that full intermontane
basin conditions were achieved during the deposition of
the Tilcara Formation, but that an antecedent drainage
system traversing a proto-Tilcara range may have existed
prior to that (Fig. 11).
Palaeo-environmental evidence from the Miocene-Pli-
ocene palaeontological and sedimentary record prior to
pronounced uplift indicates that climatic conditions dur-
ing that time were relatively humid. While the Uquıa
Formation is well known for its fossil assemblage indicat-
ing subtropical to tropical warm, humid conditions
(Alonso et al., 2006; Reguero et al., 2007; Reguero &
Candela, 2008), ostracods analysed in this study addition-
ally attest to the existence of permanent freshwater bodies
in the lower Maimara Formation, clearly indicating that
topography and relief conditions must have been sub-
dued. Calcic rhizoconcretions and nodules in the upper
section of the Maimara Formation indicate regional wet
dry seasonality, a characteristic, which is maintained in
the overlying Tilcara and Uquıa formations. In stark con-
trast are the more arid conditions and the efficient oro-
graphic barrier to the east observed in the Humahuaca
Basin today. Stratigraphic relationships and inferred pal-
aeoclimatic indicators suggest that aridification may be a
relatively young phenomenon, related to the uplift of the
Tilcara ranges.
The timing of tectonic uplift and associated aridifica-
tion appears to have been diachronous and basin-specific
throughout the southern Central Andes, and determined
by the individual behaviour of the basin-bounding faults.
Although such environmental shifts are a hallmark of vir-
tually all intermontane basins along the eastern flank of
the Puna Plateau and generally took place during the Plio-
cene (Bossi et al., 2001; Kleinert & Strecker, 2001; Starck
& Anzotegui, 2001; Coutand et al., 2006; Hain et al.,
2011), our observations demonstrate that in the Humahu-
aca Basin, the change to drier conditions probably
occurred as late as Plio-Pleistocene time.
Basin -fill evolution and deformation in the
Humahuaca Basin
The abrupt change in palaeoflow directions during the
deposition of the Tilcara Formation at ~4.2 Ma and the
associated facies change from distal to rather proximal
sources (upward coarsening; Figs 3 and 6) record the tec-
tonically induced reorganization of east-flowing river net-
works that formerly traversed the Tilcara ranges
(Fig. 11). By analogy with neighbouring intermontane
basins (e.g. the Toro and Lerma basins), the strata of the
Tilcara Formation are interpreted as heralding the attain-
ment of an intermontane basin stage. At that time, deposi-
tion in the Humahuaca Basin must have taken place
under conditions of restricted fluvial connectivity with
the foreland; otherwise, the deposition of over 250 m of
sediment within the basin may not have been possible.
Others have suggested that protracted internal drain-
age (e.g. Sobel et al., 2003) or restricted external drainage
(e.g. Hilley & Strecker, 2005) appear to be related to the
combined effects of high uplift rates, exposure of resistant
rocks and aridity. In this context, the following mecha-
nisms for basin aggradation between 4.2 and ~1.5 Ma
may be envisioned: (1) increased surface-uplift rates in
the Tilcara ranges, associated with activity along a reverse
fault that obliquely crosses the outlet region, and/or
exposure of more resistant rock types reducing the fluvial
transport efficiency within the basin; and (2) a change in
global and/or regional climatic conditions towards
increased aridity, reduced runoff and transport capacity.
Reduced runoff by increased aridity appears unlikely
given the record of humid palaeoclimatic conditions.
Moreover, the Tilcara Formation belongs to a group of
spatially widespread Miocene-Pliocene conglomerates in
NW Argentina commonly known as Punaschotter (Penck,
1920). These deposits are more likely related to individual
©2012 The Authors
Basin Research ©2012 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 567
Basin & landscape evolution in the E Cordillera
range uplifts, because their diachronous deposition in var-
ious intermontane basins and foreland regions excludes a
coeval sedimentary response to regional climate change
(McPherson, 2008; Schoenbohm et al., 2008). As glacia-
tion in the Central Andes has not been documented prior
to 3.5 Ma (Clapperton, 1979) and subsequent glaciations
have been minor due to limited moisture availability
(Haselton et al., 2002), it seems also unlikely that the style
and deposition of the Tilcara conglomerates in the Hu-
mahuaca Basin results from glacial erosion at high eleva-
tion. Therefore, although a global trend towards colder
and drier climates since mid-Miocene (e.g., Zachos et al.,
2001) may have favoured basin isolation, it is unlikely to
have initiated basin filling here at 4.2 Ma. We therefore
favour the first scenario as the most likely mechanism for
initializing partial hydrological isolation of the Humahu-
aca Basin, which is consistent with the inferred tectonic
forcing and severing of the fluvial system. The transition
towards coarser grain sizes with the onset of deposition of
the Tilcara Formation is thus best explained by erosion of
more proximal sources in the uplifting Sierra Alta and
Tilcara ranges.
Based on the depositional age estimates from the Tilcara
and Uquıa formations (i.e. ~4.21.5 Ma) and the uncon-
formably overlying Quaternary gravel (<1.06 0.10 Ma),
it appears that fluvial connectivity with the foreland must
have been re-established in the intervening time period.
Although our data are currently insufficient to resolve the
processes that led to basin excavation, incision rates at the
valley outlet must have been sufficient to preserve external
drainage, and sufficient transport capacity must have
existed to allow sediment from upstream to bypass the
basin while large volumes of sediment were removed.
These processes, however, do not necessarily imply more
availability of moisture in the Humahuaca Basin itself, but
may instead represent efficient headward erosion at the
location of the valley outlet, where current precipitation
rates abruptly decline upstream towards the Humahuaca
Basin (Strecker et al., 2007).
Although later surface processes have often altered
the unconformity in the Tilcara Formation, preserved
vestiges of deformed erosion surfaces suggest that flu-
vial incision predated an episode of deformation in the
basin. These events were followed by a prolonged
phase of restricted hydrological connectivity with the
foreland associated with basin filling of up to 400-m
thick fluvial and alluvial gravel between >1.06 and
<0.8 Ma. Whether this filling episode was linked to
tectonically induced basin isolation farther downstream,
regional uplift of the Tilcara ranges or the result of
climatic forcing remains uncertain. It is, however, con-
ceivable that the ongoing surface uplift of the Tilcara
ranges exceeded a regional threshold elevation to form
an efficient orographic barrier to cause basin-wide
semi-arid conditions (i.e. reduced runoff and transport
capacities) as observed today.
These 400-m thick gravels were largely removed from
the basin during basin excavation, followed by the
deposition of a second major fill, which reached a thick-
ness of approximately 250 m between at least about 94
and 65 ka in the Humahuaca Basin (Robinson et al.,
2005; Sancho et al., 2008) and until after ~50 ka in the
tributary Quebrada de Purmamarca (this study). This fill
unit and all previous deposits now form the substrate for
gravel-covered pediments and multiple fluvial terraces
that have been sculpted into these deposits. These terrace
systems were formed at successively lower elevations
(Fig. 8f), indicating renewed incision that may record
recent reductions in channel gradients and/or changes in
local baselevel with respect to the Andean foreland. These
periods of sediment removal are episodic and have
occurred some time between 65 and 50 ka and the present
day. Faulting that has affected these deposits is an expres-
sion of renewed deformation within the basin during, or
shortly after an episode of extensive basin excavation in
the Pleistocene (Fig. 8a).
Although the last basin-fill episode in the Humahuaca
Basin coincides to some extent with more humid phases
documented in the Altiplano-Puna Plateau (e.g. Bobst
et al., 2001; Placzek et al., 2006), a direct correlation can-
not be observed. Due to limited age control on the youn-
ger Quaternary deposits and landforms in the
Humahuaca Basin, we cannot entirely exclude such a pos-
sible correlation. However, aggradation during humid
phases requires that sediment generation from hillslopes
must increase with precipitation more rapidly than the
increase in the transport capacity of rivers that would
result from enhanced discharge in rivers during this time.
Interestingly, our observations point towards a sys-
tematic behaviour among erosion, deposition and tec-
tonic processes in the Humahuaca Basin, an intriguing
relationship that can also be found in other intermon-
tane basins of the NW Argentine Andes (Strecker
et al., 1989; Hilley & Strecker, 2005). According to
these observations, basin-internal/out-of-sequence
deformation occurs during or following episodes of
enhanced basin excavation. This is documented at least
three times in the stratigraphic record of the southern
Humahuaca Basin. First, after ~1.5 Ma, the Tilcara
Formation was partly removed from the basin followed
by major out-of-sequence thrusting along the basin-
internal Tumbaya Fault. Here, deformation ceased
prior to 1 Ma, which is documented by the onlap of
~1-Myr old gravels. Second, deformed remnants of
these gravels, which lack evidence of syntectonic depo-
sition, might also indicate that deformation occurred in
association with the removal of the strata. Third,
renewed, but locally limited deformation of young river
terraces near Tilcara is observed right after the removal
of the youngest Quaternary gravels. Although this phe-
nomenon and the associated mechanisms will require
more detailed studies in the future, it is conceivable
that the removal of basin fills and the resulting reduc-
tion in lithostatic stresses on formerly locked thrust
and reverse faults could ultimately result in the reacti-
vation of these faults.
©2012 The Authors
Basin Research ©2012 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
568
H. Pingel et al.
Regional context of foreland fragmentation
and basin-fill evolution
Many intermontane basins along the Puna margin record
the partial or complete loss of fluvial connectivity with the
foreland through the presence of thick conglomeratic fills
(Strecker et al., 2007). Periodic reconnection to the fore-
land is documented by marked fluvial incision and excava-
tion of these basin fills, reflected in distinct regional
unconformities and complex onlap relationships. Virtu-
ally all basins in the northern Sierras Pampeanas, the
Santa Barbara System and the Eastern Cordillera of
north-western Argentina exhibit similar, but diachronous,
patterns of basin-fill and erosion throughout their history
(reviewed in Strecker et al., 2007, 2009). The unifying
characteristic in the history of these basins is that follow-
ing the reorganization of fluvial systems by surface uplift,
the establishment of orographic barriers results in pro-
gressive aridification, interrupted drainages and reduced
fluvial transport capacity (Sobel et al., 2003). While the
basins in the arid interior of the orogen (the Puna Plateau)
have maintained internal drainage systems, resulting in
thick sediment accumulations (Alonso et al., 1991; Van-
dervoort et al., 1995), the intermontane basins along the
eastern flank have alternated between restricted external
and transient internal drainages, and fully integrated flu-
vial systems connected with the foreland.
At present, virtually all of the intermontane basins to
the east of the Puna are hydrologically connected to the
foreland, often via narrow, deeply incised bedrock gorges
(Fig. 1). For example, the Toro Basin at 24.5°S was cut
off from the foreland between 8 and 6 Ma and has subse-
quently experienced at least two cycles of basin filling and
excavation (Marrett & Strecker, 2000; Hilley & Strecker,
2005). Similarly, between 5.2 and 2.4 Ma, the uplift of an
orographic barrier to the east of the present-day Calc-
haquı Valley at ~25.5°S caused aridification associated
with deposition of the conglomeratic San Felipe Forma-
tion, which was subsequently incised, deformed and
finally overlain by another conglomeratic gravel after
2.4 Ma (Coutand et al., 2006). After renewed sediment
removal, a subsequent gravel unit with complex onlap
relationships was deposited that once formed a continu-
ous surface of coalesced alluvial fans and river gravel.
River superposition, incision and removal of most of the
gravels in the lower Calchaquı Basin document the ongo-
ing erosion of this unit (Coutand et al., 2006; Strecker
et al., 2007). The Santa Marıa Basin at ~26.5°S is a result
of foreland compartmentalization related to basement
uplift to the east that occurred after 6 Ma (Kleinert &
Strecker, 2001; Bossi et al., 2001;Sobel & Strecker, 2003),
which was followed by aridification, severed fluvial con-
nection to the foreland, deformation, erosion and finally,
the deposition of thick conglomerates after 2.9 Ma (Strec-
ker et al., 1989). These units were subsequently incised,
as documented by successively lower pediments and flu-
vial terraces. Further examples of this type of basin devel-
opment are provided by the El Cajon and Fiambala
basins, at 27°S and 27.5°S, respectively (Mortimer et al.,
2007; Carrapa et al., 2008), emphasizing the similarities
between the processes and depositional facies that control
the evolution of intermontane basins within this environ-
ment.
We suggest that the tectono-sedimentary history of the
Humahuaca Basin and other intermontane basins in NW
Argentina can be best explained through a threshold pro-
cess described in Sobel et al. (2003), in which active uplift
of downstream topographic barriers steepens the channels
that traverse these ranges, while aggradation upstream
must keep pace with the associated uplift of the channel.
As rates of rock uplift increase relative to the transport
efficiency (related to precipitation) and bedrock erodibil-
ity, internal drainage is favoured. Conversely, low rates of
rock uplift in the downstream basement ranges, a high
level of erodibility of exposed rock types and pronounced
rainfall gradients all promote incision, headward erosion
and basin capture. All of these processes ultimately con-
tribute to sustaining fluvial connectivity with the fore-
land. This, however, is only possible if the narrow outlets
of the arid basins are in close proximity to steep rainfall
and run-off gradients or if the structural setting is condu-
cive to funnelling moist air into the orogen interior during
protracted moist episodes. In turn, this condition would
increase precipitation, run-off and erosion, which would
ultimately help to achieve or re-establish external drain-
age conditions. Indeed, on the basis of our chronology,
observations and interpretations, we are able to show that
the intermontane Humahuaca Basin is the result of pro-
gressive rock uplift and associated surface processes in the
Eastern Cordillera that commenced at 1510 Ma and sub-
sequently led to deposition of the Maimara Formation at
about 6 Ma. The largely continuous depositional system
finally became dismembered after 4.2 Ma when topogra-
phy of the Tilcara ranges deflected the fluvial network
into rang-parallel drainage. Until about 1.5 Ma, the basin
was characterized by restricted fluvial conditions most
likely related to ongoing regional uplifts. Afterwards, the
basin was rapidly excavated and internally deformed. The
ensuing regional palaeotopography in the basin was sub-
sequently refilled between >1.06 and <0.8 Ma. The rea-
son for the initial excavation is not known, but this could
have been related to headward erosion and fluvial connec-
tivity, followed by re-established hydrological isolation.
The resulting basin fill, then, was largely removed,
deformed and replaced by a younger fill that periodically
has been excavated some time after ~65 and 50 ka. Exca-
vation, again, seems to have been accompanied by basin-
internal deformation, recorded by offset fluvial terraces
near Tilcara, while currently the basin is aggrading (Riv-
elli & Flores, 2009).
SUMMARY AND CONCLUSIONS
In this study, we have combined new provenance and
palaeocurrent data from the sedimentary record of the
©2012 The Authors
Basin Research ©2012 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 569
Basin & landscape evolution in the E Cordillera
intermontane Humahuaca Basin of the Eastern Cordillera
of the southern Central Andes with 12 new
206
Pb/
238
U
zircon age estimates from intercalated volcanic ash depos-
its to assess its spatiotemporal evolution. This enabled us
to improve our understanding of the neotectonic basin
and landscape evolution of an intermontane setting in the
immediate vicinity of the intra-orogenic Puna Plateau, the
world’s second largest plateau, and an important barrier
to atmospheric circulation and surface-process regimes.
We suggest that the coupled tectonic, erosion and sedi-
mentary processes and associated landscape development
in the Humahuaca Basin reflect an environment whose
evolution is relevant for the assessment of intermontane
basins worldwide, including the North American Lara-
mide province, and the Tien Shan and Qilian Shan base-
ment uplifts in Asia. From our analysis, we draw the
following conclusions:
(1) In comparison with the previously developed chro-
nostratigraphy for the Humahuaca Basin, our new
UPb zircon dates extend the lower boundary of
the Maimara Formation into late Miocene, older
than 5.92 0.12 Ma, confirms existing ages from
the northern basin, and reveals Quaternary fills as
old as ~1 Ma.
(2) On the basis of provenance, lithology and spatial
distribution, we introduced a new stratigraphic
unit, the Tilcara Formation, in the southern
Humahuaca Basin that is apparently coeval with
the radiometrically and palaeontologically con-
strained Uquıa Formation from the central and
northern sectors of the basin. The Tilcara Forma-
tion highlights the different depositional environ-
ments and source areas between the southern and
northern parts of the basin.
(3) The sedimentary units in the Humahuaca Basin
record a transition between a partially segmented
foreland basin and a fault-bounded intermontane
basin in the course of surface uplift to the east.
This resulted in a change in fluvial connectivity
and the re-arrangement of the formerly eastward-
draining river network into an axial, south-directed
drainage after ~4.2 Ma.
(4) Repeated hydrological disconnection from the fore-
land due to tectonism and ensuing aridification in
the lee of rising topography repeatedly resulted in
restricted fluvial connectivity and possibly transient
fluvial isolation and accumulation of at least three
basin-fill units during Plio-Pleistocene times. After
episodic re-capture, these fills were partially evacu-
ated by fluvial incision. This is similar to other inter-
montane basins along the eastern flank of the Puna
where surface uplift resulted in the tectonic defeat
of fluvial networks, hydrological isolation and basin
aggradation, until renewed river incision exceeded
rock uplift at tectonically active basin outlets.
(5) We furthermore observed that major deformation
events in the Humahuaca Basin apparently fol-
lowed a phase of enhanced removal of basin-fill
units, a scenario that has been observed in other
intermontane basins of NW Argentina. We
speculate that this behaviour is related to the
reduction in lithostatic stresses acting on subsur-
face structures during major phases of basin
excavation.
ACKNOWLEDGEMENTS
Strecker, Alonso and Pingel acknowledge funding by the
DFG-Leibniz Center for Earth Surface Process and Cli-
mate Studies (DFG grant STR 373/19-19) and funds
provided by DFG grant STR373/32-1. Schmitt acknowl-
edges the use of the ion microprobe facility at UCLA, in
part supported by a grant from the Instrumentation and
Facilities Program, Division of Earth Sciences, U.S.
National Science Foundation. We are grateful for con-
structive reviews by F. Schlunegger, F. Davila, J. Kley
and editorial support from P. van der Beek.
SUPPORTING INFORMATION
Additional Supporting Information may be found in the
online version of this article:
Table S1. Summary of UPb zircon analytical data of
volcanic ash samples in the Humahuaca Basin using
CAMECA IMS 1270 ion microprobe at UCLA.
REFERENCES
ACEN
˜OLAZA, G.F. (2003) The Cambrian System in Northwest-
ern Argentina: stratigraphical and palaeontological frame-
work. Geol. Acta,1(1), 2339.
ALLMENDINGER, R.W. & ZAPATA, T. (2000) The footwall ramp
of the Subandean decollement, northernmost Argentina, from
extended correlation of seismic reflection data. Tectonophysics,
321(1), 3755.
ALLMENDINGER, R.W., JORDAN, T.E., KAY, S.M. & ISACKS, B.L.
(1997) The evolution of the Altiplano-Puna Plateau of the
Central Andes. Annu. Rev. Earth Planet. Sci.,25(1), 139174.
ALONSO, R.N., JORDAN, T.E., TABBUTT, K.T. & VANDERVOORT,
D.S. (1991) Giant evaporite belts of the Neogene central An-
des. Geology,19(4), 401404.
ALONSO, R.N., BOOKHAGEN, B., CARRAPA, B., COUTAND, I., HAS-
CHKE, M., HILLEY, G.E., SCHOENBOHM, L., SOBEL, E.R.,
STRECKER, M.R., TRAUTH, M.H. & VILLANUEVA, A. (2006)
Tectonics, climate, and landscape evolution of the Southern
Central Andes: The Argentine Puna Plateau and adjacent
regions between 22 and 30°S. In: The Andes,22 (Ed. by O.
Oncken, G. Chong, G. Franz, P. Giese, H.-J. Gotze, V. A.
Ramos, M.R. Strecker & P. Wigger), pp. 265283. Frontiers
in Earth Sciences, Springer, Berlin, Heidelberg.
BEAUMONT, C. (1981) Foreland basins. Geophys. J. Roy. Astron.
Soc.,65(2), 291329.
©2012 The Authors
Basin Research ©2012 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
570
H. Pingel et al.
BEVIS,M.&ISACKS, B.L. (1984) Hypocentral trend surface anal-
ysis: probing the geometry of Benioff Zones. J. Geophys. Res.,
89 (B7), 61536170.
BOBST, A.L., LOWENSTEIN, T.K., JORDAN, T.E., GODFREY, L.V.,
KU, T.L. & LUO, S. (2001) A 106ka paleoclimate record from
drill core of the Salar de Atacama, northern Chile. Palaeogeo-
graphy, Palaeoclimatology, Palaeoecology,173, 21–42.
BOLL,A.&HERNA
´NDEZ, R.M. (1986) Interpretacion estructural
del area Tres Cruces. Bol. Inf. Pet.,7,214.
BOSSI, G., GEORGIEFF, S., GAVRILOFF, I., IBAN
˜EZ,L.&MURU-
AGA, C. (2001) Cenozoic evolution of the intramontane Santa
Marıa Basin, Pampean Ranges, northwestern Argentina. J. S.
Am. Earth Sci.,14(7), 725734.
CAHILL, T., ISACKS, B.L., WHITMAN, D., CHATELAIN, J.L.,
PEREZ,A.&MING CHIU, J. (1992) Seismicity and tectonics in
Jujuy Province, northwestern Argentina. Tectonics,11(5), 944
959.
CARRAPA, B., ADELMANN, D., HILLEY, G.E., MORTIMER, E., SO-
BEL, E.R. & STRECKER, M.R. (2005) Oligocene range uplift
and development of plateau morphology in the southern
central Andes. Tectonics,24, TC4011, doi:10.1029/
2004TC001762
CARRAPA, B., STRECKER, M.R. & SOBEL, E.R. (2006) Cenozoic
orogenic growth in the Central Andes: evidence from sedi-
mentary rock provenance and apatite fission track thermo-
chronology in the Fiambala Basin, southernmost Puna
Plateau margin (NW Argentina). Earth Planet. Sci. Lett.,247
(12), 82100.
CARRAPA, B., HAUER, J., SCHOENBOHM, L., STRECKER, M.R.,
SCHMITT, A.K., VILLANUEVA,A.&SOSA GOMEZ, J. (2008)
Dynamics of deformation and sedimentation in the northern
Sierras Pampeanas: an integrated study of the Neogene Fiam-
bala basin, NW Argentina. Geol. Soc. Am. Bull.,120(1112),
15181543.
CASTELLANOS, A. (1950) El Uquiense, Sedimentos Neogenos de
Uquıa (Senador Perez) en la Provincia de Jujuy (Argentina).
Facultad de Ciencias Matematicas Fısicas y Quımicas y Natu-
rales de la Universidad Nacional del Litoral, 36, Serie Tecnic-
o Cientıfico, 55 p., Rosario.
CLAPPERTON, C.M. (1979) Glaciation in Bolivia before
3.27 Myr. Nature,277(5695), 375377.
COUTAND, I., COBBOLD, P.R., de URREIZTIETA,M.,GAUTIER,P.,
CHAUVIN, A., GAPAIS,D.,ROSSELLO,E.A.&LO
´PEZ-GAMUND,
O. (2001) Style and history of Andean deformation, Puna Pla-
teau, Northwestern Argentina Tectonics,20(2), 210234.
COUTAND, I., CARRAPA, B., DEEKEN, A., SCHMITT, A.K., SOBEL,
E.R. & STRECKER, M.R. (2006) Propagation of orographic bar-
riers along an active range front: insights from sandstone
petrography and detrital apatite fission-track thermochronol-
ogy in the intramontane Angastaco Basin, NW Argentina
Basin Res.,18(1), 126.
DECELLES, P.G. & GILES, K.A. (1996) Foreland basin systems.
Basin Res.,8(2), 105123.
DEEKEN, A., SOBEL, E.R., HASCHKE,M.&RILLER, U. (2004)
Age of initiation and growth pattern of the Puna Plateau,
NW-Argentina, constrained by AFT thermochronology. 19th
Colloquium on Latin American Geosciences. Terra Nostra, Pots-
dam, Germany, Abstract Volume, 05(1), 39.
DEEKEN, A., SOBEL, E.R., HASCHKE,M.&RILLER, U. (2005)
Age of initiation and growth pattern of the Puna plateau,
NW-Argentina, constrained by AFT thermochtonology. 19th
Colloquium on Latin American Geosciences, GeoForschungs-
Zentrum Potsdam, Potsdam, Germany, 18–20 April.
DEEKEN, A., SOBEL, E.R., COUTAND, I., HASCHKE, M., RILLER,
U. & STRECKER, M.R. (2006) Development of the Southern
Eastern Cordillera, NW Argentina, constrained by apatite fis-
sion track thermochronology: from Early Cretaceous exten-
sion to Middle Miocene shortening. Tectonics,25, TC6003,
doi: 10.1029/2005TC001894
DEL PAPA, C.E., HONGN, F.D., POWELL, J., DOCAMPO, M.,
STRECKER, M.R., PETRINOVIC, I.A., SCHMITT, A.K. & PER-
EYRA, R. (in review) Middle Eocene-Oligocene broken fore-
land evolution in the Andean Valle Calchaqui, NW
Argentina: insights from stratigraphic, structural, and prove-
nance studies. Unpublished Manuscript.
DıAZ, J.I. & MALIZZIA, D.C. (1983) Estudio Geologico y sedi-
mentologico del Terciario Superior del Valle Calchaquı. Uni-
versidad Nacional de Tucuman, Facultad de Ciencias Naturales,
Boletın Sedimentologico,2(1), 828.
GARCIA-CASTELLANOS, D. & (2007) The role of climate during
high plateau formation. Insights from numerical experiments.
Earth Planet. Sci. Lett.,257, 372390.
GALLISKI,M.&VIRAMONTE, J.G. (1988) The Cretaceous paleor-
ift in Northwestern Argentina: a petrologic approach. J. S.
Am. Earth Sci.,1(4), 329342.
GEBHARD, J., GUIDICE,A.&GASCON, J. (1974) Geologıa de la co-
marca entre el Rıo Juramento y Arroyo Las Tortugas, Provin-
cias de Salta y Jujuy, Republica Argentina. Revista de la
Asociacion Geologica Argentina,29(3), 359375.
GILE, L.H., PETERSON, F.F. & GROSSMAN, R.B. (1966) Morpho-
logical and genetic sequences of carbonate accumulation in
desert soils. Soil Sci.,101(5), 347360.
GROVE, M., JACOBSON, C.E., BARTH, A.P. & VUCIC, A. (2003)
Temporal and spatial trends of Late Cretaceous - early Ter-
tiary underplating Pelona and related schist beneath southern
California and southwestern Arizona. Geol. Soc. Am. Spec.
Pap.,374, 381406.
HAIN, M., STRECKER, M.R., BOOKHAGEN, B., ALONSO, R.N., PIN-
GEL,H.&SCHMITT, A.K. (2011) Neogene to Quaternary bro-
ken-foreland formation and sedimentation dynamics in the
Andes of NW Argentina (25°S). Tectonics,30, TC2006,
doi:10.1029/2010TC002703
HASELTON, K., HILLEY, G.E. & STRECKER, M.R. (2002a) Aver-
age Pleistocene climatic patterns in the southern Central An-
des: controls on mountain glaciation and paleoclimate
implications. J. Geol.,110(2), 211226.
HILLEY, G.E. & STRECKER, M.R. (2005) Processes of oscillatory
basin filling and excavation in a tectonically active orogen.
Quebrada del Toro Basin, NW Argentina. Geol. Soc. Am.
Bull.,117(78), 887901.
HILLEY, G.E., BLISNIUK, P.M. & STRECKER, M.R. (2005)
Mechanics and erosion of basement-cored uplift provinces.
J. Geophys. Res.,110 (B12), B12409, doi:10.1029/
2005JB003704
HONGN, F.D., DELPAPA, C., POWELL, J., PETRINOVIC, I., MON,
R. & DERACO, V. (2007) Middle Eocene deformation and sedi-
mentation in the Puna-Eastern Cordillera transition (23°-
26°S): control by pre-existing heterogeneities on the pattern
of initial Andean shortening. Geology,35 (3), 271274.
GARCIA-CASTELLANOS, D. (2007) The role of climate during
high plateau formation. Insights from numerical experiments.
Earth Planet Sci. Lett.,257, 372–390. doi:10.1016/j.epsl.
2007.02.039
INSEL, N., GROVE, M., HASCHKE, M., BARNES, J.B., SCHMITT,
A.K. & STRECKER, M.R. (2012) Paleozoic to early Cenozoic
cooling and exhumation of the basement underlying the east-
©2012 The Authors
Basin Research ©2012 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 571
Basin & landscape evolution in the E Cordillera
ern Puna plateau margin prior to plateau growth. Tectonics,
31, TC6006. doi:10.1029/2012TC003168
JORDAN, T.E. & ALLMENDINGER, R.W. (1986) The Sierras Pam-
peanas of Argentina: a modern analogue of Rocky Mountain
foreland deformation. Am. J. Sci.,286 , 737764.
JORDAN, T.E. & ALONSO, R.N. (1987) Cenozoic stratigraphy and
basin tectonics of the Andes Mountains, 20-28°S lat. AAPG
Bull.,71 (1), 4964.
KLEINERT,K.&STRECKER, M.R. (2001) Climate change in
response to orographic barrier uplift: paleosol and stable isotope
evidence from the Late Neogene Santa Marıa Basin, Northwes-
tern Argentina. Geol. Soc. Am. Bull.,113(6), 728742.
KLEY,J.&MONALDI, C. (2002) Tectonic inversion in the Santa
Barbara System of the central Andean foreland thrust belt,
northwestern Argentina. Tectonics,21, 1061, doi:10.1029/
2002TC902003.
KLEY, J., MuLLER, J., TAWACKOLI, S., JACOBSHAGEN,V.&MA-
NUTSOGLI, U. (1997) Pre-Andean and Andean-age deforma-
tion in the Eastern Cordillera of southern Bolivia. J. S. Am.
Earth Sci.,10(1), 119.
KLEY, J., ROSSELLO, E.A., MONALDI, C.R. & HABIGHORST,B.
(2005) Seismic and field evidence for selective inversion of
Cretaceous normal faults, Salta Rift, Northwestern Argen-
tina. Tectonophysics,399(14), 155172.
KRAEMER, B., ADELMANN, D., ALTEN,M.&SCHNURR, W. (1999)
Incorporation of the Paleogene foreland into the Neogene
Puna plateau: the Salar de Antofalla area, NW Argentina. J.
S. Am. Earth Sci.,12(2), 157182.
LETCHER, A.J. (2007) Deformation history of the Susques basin
(~23°S, 66°W), Puna Plateau, NW Argentina: new constraints
by apatite (U-Th)/He thermochronology and
40
Ar/
39
Ar geo-
chronology, and implications for plateau formation in the cen-
tral Andes. MSc Thesis, Stanford University, USA.
MACHETTE, M.N. (1985) Calcic soils of the southwestern United
States. In: Soils and Quaternary Geology of the Southwestern
United States,203 (Ed. by D.L. Weide & M.L. Faber), pp. 1
21. Geol. Soc. Am. Special Paper, Denver, CO.
MARQUILLAS, R.A., DELPAPA,C.&SABINO, I.F. (2005)
Sedimentary aspects and paleoenvironmental evolution of a
rift basin: salta Group (Cretaceous-Paleogene), northwestern
Argentina. Int. J. Earth Sci. (Geol.Rundsch.),94 (1), 94113.
MARRETT, R.A. & STRECKER, M.R. (2000) Response of intracon-
tinental deformation in the Central Andes to Late Cenozoic
reorganization of South American plate motions. Tectonics,19
(3), 452467.
MARRETT, R.A., ALLMENDINGER, R.W., ALONSO, R.N. & DRAKE,
R.E. (1994) Late Cenozoic tectonic evolution of the Puna Pla-
teau and adjacent foreland, Northwestern Argentine Andes.
J. S. Am. Earth Sci.,7(2), 179207.
MARSHALL, L.G., BUTLER, R.F., DRAKE, R.E. & CURTIS, G.H.
(1982) Geochronology of Type Uquian (Late Cenozoic) Land
Mammal Age. Argent. Sci.,216(4549), 986989.
MCPHERSON, H.M. (2008) Climate and tectonic controls on sedi-
mentation and deformation in the Fiambala Basin of the
southern Puna Plateau, Northwest Argentina. MSc Thesis,
Ohio State University, USA.
MEYER, B., TAPPONNIER, P., BOURJOT, L., METIVIER, F., GAUDE-
MER, Y., PELTZER, G., SHUNMIN,G.&ZHITAI, C. (1998) Crus-
tal thickening in Gansu-Qinghai, lithospheric mantle
subduction, and oblique, strike-slip controlled growth of the
Tibet plateau. Geophys. J. Int.,135(1), 147.
MORTIMER,E.,CARRAPA,B.,COUTAND,I.,SCHOENBOHM,L.,SO-
BEL, E.R., SOSA GOMEZ,J.&STRECK ER, M.R. (2007) Fragmenta-
tion of a foreland basin in response to out-of-sequence basement
uplifts and structural reactivation: El Cajon-Campo del Arenal
Basin, NW Argentina. Geol. Soc. Am. Bull.,119(56), 637653.
MOYA, M.C. (1988) Lower Ordovician in the southern part of
the argentine Eastern Cordillera. In: The Southern Central An-
des,17, (Ed. by H. Bahlburg, C. Breitkreuz & P. Giese), pp.
5569. Frontiers in Earth Sciences, Springer Berlin Heidel-
berg.
MULCH, A., UBA, C.E., STRECKER, M.R., SCHOENBERG,R.&
CHAMBERLAIN, C.P. (2010) Late Miocene climate variability
and surface elevation in the Central Andes. Earth Planet. Sci.
Lett.,290(12), 173182.
OMARINI, R. (1983) Caracterizacion Litologica, Diferencion y
Genesis de la Formacion Puncoviscana Entre el Valle de Lerma y
la Faja Eruptiva de la Puna. Unpublished thesis, Universidad
Nacional de Salta, Argentina.
PACES, J.B. & MILLER, J.D. (1993) Precise U-Pb ages of Duluth
Complex and related mafic intrusions, northeastern Minne-
sota: geochronological insights into physical, petrogenetic,
paleomagnetic and tectonomagnetic processes associated with
the 1.1 Ga mid-continent rift system. J. Geophys. Res.,98,
1399714013.
PENCK, W. (1920) Der Südrand der Puna de Atacama (NW-Ar-
gentinien): Ein Beitrag zur Kenntnis des Andinen
Gebirgstypus und der Frage der Gebirgsbildung. Abhandlun-
gen der Sachsischen Akademie der Wissenchaften,1,3420.
PLACZEK, C., QUADE,J.&PATCHETT, P.J. (2006) Geochronology
and stratigraphy of late Pleistocene lake cycles on the southern
Bolivian Altiplano: implications for causes of tropical climate
change. Geol. Soc. Am. Bull.,118(56), 515532.
REGUERO, M.A. & CANDELA, A.M. (2008) Bioestratigrafıa de las
secuencias neogenas tardıas de la Quebrada de Humahuaca,
Provincia de Jujuy. Implicancias paleoambientales y paleobio-
geograficas. Relatorio del XVII Congreso Geologico Argentino,
1, 286296.
REGUERO, M.A., CANDELA, A.M. & ALONSO, R.N. (2007) Bio-
chronology and biostratigraphy of the Uquıa Formation (Plio-
cene-Early Pleistocene, NW Argentina) and its significance in
the Great American Biotic Interchange. J. S. Am. Earth Sci.,
23(1), 116.
REYNOLDS, J.H., IDLEMAN, B. D., HERNA
´NDEZ,R.M.&NAESER,
C. W. (1994) Preliminary chronostratigraphic constraints on
Neogene tectonic activity in the Eastern Cordillera and Santa
Barbara System, Salta Province, NW Argentina. GSA
Abstracts,26 (7), A503.
REYNOLDS,J.,GALLI,C.,HERNA
´NDEZ,R.,IDLEMAN,B.,KOTILA,
J., HILLIARD,R.&NAESER, C. (2000) Middle Miocene tectonic
development of the transition zone, Salta Province, northwest
Argentina: magnetic stratigraphy from the Metan Subgroup,
Sierra de Gonzalez. Geol. Soc. Am. Bull.,112(11), 17361751.
REYNOLDS, J., HERNA
´NDEZ, R.M., GALLI, C.I. & IDLEMAN,
B.D. (2001) Magnetostratigraphy of the Quebrada La Por-
celana section, Sierra de Ramos, Salta Province, Argentina:
age limits for the Neogene Oran Group and uplift of the
southern Sierras Subandinas. J. S. Am. Earth Sci.,14(7),
681692.
RILLER, U., PETRINOVIC, I., RAMELOW, J., STRECKER,M.&
ONCKEN, O. (2001) Late Cenozoic tectonism, collapse caldera
and plateau formation in the central Andes. EPSL,188(34),
299311.
RIVELLI,F.R.&FLORES, E.M. (2009) Proteccio
´ndema
´rgenes en
el
´o grande, tramo las quebradas trancas - Tilcara. Cuarto
Simposio Regional sobre Hidrı
´ulica de
´os, Salta, Argentina. 4p.
©2012 The Authors
Basin Research ©2012 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
572
H. Pingel et al.
ROBINSON, R.A.J., SPENCER, J.Q.G., STRECKER, M.R., RICHTER,
A. & ALONSO, R.N. (2005) Luminescence dating of alluvial
fans in intramontane basins of NW Argentina. In: Alluvial
Fans: Geomorphology, Sedimentology, Dynamics,251 (Ed. by
A. Harvey, A. Mather & M. Stokes) pp. 153168, Geol. Soc.
London Special Pub.
RODRI
´GUEZ-FERNA
´NDEZ, L.R., HEREDIA,N.,SEGGIARO,R.E.&
GONZA
´LEZ, M. (1999) Estructura andina de la Cordillera
Oriental en el area de la Quebrada de Humahuaca, Provincia
de Jujuy, NO de Argentina. Trabajos de Geologıa,21,321332.
RUSSO,A.&SERRAIOTTO, A. (1978) Contribucion al conocimien-
to de la estratigrafıa Terciaria en el Noroeste Argentino. Actas
del VII Congreso Geologico Argentino,1, 715730.
SALFITY, J.A. (1982) Evolucion paleogeografica del Grupo Salta
(Cretacieo-Eogenico), Argentina. Actas del V Congreso Latino-
american Geologıa,1,1126.
SALFITY, J.A., BRANDA
´N, E.M., MONALDI, C.R. & GALLARDO,
E.F. (1984) Tectonica compresiva Cuaternaria en la Cordil-
lera Oriental Argentina latitud de Tilcara (Jujuy). Actas del
IX Congreso Geologico Argentino,2, 427434.
SA
´NCHEZ, M.C. & MARQUILLAS, R.A. (2010) Facies y ambientes
del grupo Salta (cretacico-paleogeno) en Tumbaya, Quebrada
de Humahuaca, provincia de Jujuy. Revista de la Asociacion
Geologica Argentina,67(3), 383391.
SA
´NCHEZ,M.&SALFITY, J.A. (1999) La cuenca Cambrica del
Grupo Meson en el Noroeste Argentino: desarrollo
estratigrafico y paleogeografico. Actas del Geologica Hispanica,
34(23), 123139.
SANCHO,C.,PENA,J.L.,RIVELLI,F.,RHODES,E.&MUNOZ,A.
(2008) Geomorphological evolution of the Tilcara alluvial fan (Ju-
juy Province, NW Argentina): tectonic implications and paleoen-
vironmental considerations. J. S. Am. Earth Sci.,26(1), 6877.
SCHMITT, A.K., de SILVA, S.L., TRUMBULL, R.B. & EMMER-
MANN, R. (2001) Magma evolution in the Purico ignimbrite
complex, N. Chile: evidence for zoning of a dacitic magma by
injection of rhyolitic melts following mafic recharge. Contrib.
Mineral. Petrol.,140, 680700.
SCHMITT, A.K., GROVE, M., HARRISON, T.M., LOVERA, O.,
HULEN,J.&WALTERS, M. (2003) The Geysers-Cobb Moun-
tain magma system, California (Part 1): U-Pb zircon ages of
volcanic rocks, conditions of crystallization and magma resi-
dence times. Geochim. Cosmochim. Ac.,67(18), 34233442.
SCHOENBOHM, L., CARRAPA, B., BYWATER, S., MCPHERSON,H.&
PRATT, J. (2008) Climatic and tectonic controls on deposition
of the Punaschotter conglomerate in Neogene marginal basins
of the Puna Plateau (NW Argentina): evidence from zircon
U-Pb geochronology. AGU Fall Meeting Abstracts, T41C04,
2008, San Francisco, USA.
SIKS, B.C. & HORTON, B.K. (2011) Growth and fragmentation
of the Andean foreland basin during eastward advance of
fold-thrust deformation, Puna plateau and Eastern Cordillera,
northern Argentina. Tectonics,30, TC6017, doi:10.1029/
2011TC002944
SOBEL, E.R. & STRECKER, M.R. (2003) Uplift, exhumation and
precipitation: tectonic and climatic control of Late Cenozoic
landscape evolution in the northern Sierras Pampeanas,
Argentina. Basin Res.,15(4), 431451.
SOBEL, E.R., HILLEY, G.E. & STRECKER, M.R. (2003) Formation
of internally drained contractional basins by aridity-limited
bedrock incision. J. Geophys. Res.,108 (B7), 2344, doi:10.
1029/2002JB001883
STARCK,D.&ANZoTEGUI, L. (2001) The late Miocene climatic
change - persistence of a climatic signal through the orogenic
stratigraphic record in northwestern Argentina. J. S. Am.
Earth Sci.,14(7), 763774.
STRECKER, M.R., CERVENY, P., BLOOM,A.&MALIZIA, D. (1989)
Late Cenozoic tectonism and landscape development in the
foreland of the Andes: Northern Sierras Pampeanas (2628°
S), Argentina. Tectonics,8(3), 517534.
STRECKER, M.R., ALONSO, R.N., BOOKHAGEN, B., CARRAPA, B.,
HILLEY, G.E., SOBEL, E.R. & TRAUTH, M.H. (2007) Tectonics
and climate of the Southern Central Andes. Annu. Rev. Earth
Planet. Sci.,35(1), 747787.
STRECKER, M.R., ALONSO, R.N., BOOKHAGEN, B., CARRAPA, B.,
COUTAND, I., HAIN, M., HILLEY, G.E., MORTIMER, E., SCHO-
ENBOHM,L.&SOBEL, E.R. (2009) Does the topographic distri-
bution of the Central Andean Puna Plateau result from
climatic or geodynamic processes? Geology,37(7), 643646.
STRECKER, M.R., HILLEY, G.E., BOOKHAGEN,B.&SOBEL, E.R.
(2011) Structural, geomorphic, and depositional characteris-
tics of contiguous and broken foreland basins: examples from
the eastern flanks of the central Andes in Bolivia and NW
Argentina. In: Tectonics of Sedimentary Basins: Recent
Advances (Ed. by C. Busby & A. Azor), John Wiley & Sons,
Ltd, Chichester, UK. doi: 10.1002/9781444347166.ch25
STREIT, R.L., BURBANK, D.W., STRECKER, M.R. & ALONSO,
R.N. (2012) Neogene-quaternary tectonics, sedimentation,
and erosion of the Quebrada de Humahuaca and Casa Grande
Basin, Intermontane Basins on the margin of the Puna Plateau
(23-24˚S). In: NW Argentina, AGU Fall Meeting, December
2012, San Francisco, USA.
TCHILINGUIRIAN,P.&PEREYRA, F.X. (2001) Geomorfologıa del
sector Salinas Grandes-Quebrada de Humahuaca, provincia de
Jujuy. Revista de la Asociacion Geologica Argentina,56,315.
TURNER, J.C.M. (1960) Estratigrafıa de la Sierra de Santa Victo-
ria y adyacencias. Bol. Acad. Nac. de Ciencias Cordoba,41(2),
163196.
UBA, C.E., STRECKER, M.R. & SCHMITT, A.K. (2007)
Increased sediment accumulation rates and climatic forcing
in the central Andes during the late Miocene. Geology,35
(11), 979982.
VANDERVOORT, D.S., JORDAN, T.E., ZEITLER, P.K. & ALONSO,
R.N. (1995) Chronology of internal drainage development
and uplift, southern Puna plateau, Argentine Central Andes.
Geology,23(2), 145148.
VERGANI,G.&STARCK, D. (1989) Aspectos Estructurales del
Valle de Lerma al Sur de la Ciudad de Salta. Boletın Inform.
Petrol.,20,(29). Buenos Aires
VEZZOLI, L., ACOCELLA, V., OMARINI,R.&MAZZUOLI, R. (2012)
Miocene sedimentation, volcanism and deformation in the
Eastern Cordillera (24°30S, NW Argentina): tracking the
evolution of the foreland basin of the Central Andes. Basin
Res.,24(6), 637663.
WALTHER, A., ORGEIRA, M., REGUERO, M., VERZI, D., VILAS, J.,
ALONSO, R.N., GALLADO, E., KELLY,S.&JORDAN, T.E.
(1998) Estudio Paleomagnetico, Paleontologico y Radimetrico
de la Formacion Uquıa (Plio-Pleistoceno) en Esquina Blanca
(Jujuy). Actas del X Congreso Latinoamericano de Geologıa y
VI Congreso Nacional de Geologıa Economica,1, 77.
ZACHOS, J., PAGANI, M., SLOAN, L., THOMAS,E.&BILLUPS,K.
(2001) Trends, rhythms, and aberrations in global climate
65 Ma to present. Science,292(5517), 686693.
Manuscript received 26 June 2012; In revised form 26
November 2012; Manuscript accepted 07 December 2012.
©2012 The Authors
Basin Research ©2012 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 573
Basin & landscape evolution in the E Cordillera
... The Río Iruya may have become disconnected from the Puna Plateau and begun to establish its modern extent based on the near disappearance of Neogene zircons from the sediment. axial geometry and triggering deposition of the fluvial Tilcara Formation by roughly 4.2 Ma (Fig. 3) 44 . Likewise, the nearby Sierra Alta experienced a pulse of uplift at around 4.3 Ma (Fig. 3) 45 . ...
... These differing scenarios, it should be noted, only result in minor changes to the absolute magnitude of calculated paleo-erosion rates from 1.8 Ma to 4 Ma and not the observed patterns or cyclicity. The transport contribution (N T ) to measured cosmogenic radionuclide concentrations in the analysis is assumed to be negligible based on: (1) the close proximity of the depositional basin to a relatively steep source area starting at~4 Ma; 32,44 and (2) good agreement between 26 Al: 10 Be sample ratios (for 26 Al AMS errors <35%) with expected 26 Al: 10 Be ratios assuming no significant burial during transport from~2-3 Ma (Fig. 2c). Our calculations assume a relative contribution by nucleon spallation, negative muon capture, and fast muon reactions at the surface equal to 98.15%, 1.2%, and 0.65%, respectively 73 Whereas the low-spatial resolution of the model is incapable of resolving orographic and vegetative forcings associated with the Andes (and thereby associated regional systems like the South American low level jet), CLIMBER-2 has been shown to reliably replicate both large-scale climate reconstructions from the Last Glacial Maximum, as well as more complex model simulations for the Pliocene 79,80 . ...
Article
Full-text available
It has long been hypothesized that climate can modify both the pattern and magnitude of erosion in mountainous landscapes, thereby controlling morphology, rates of deformation, and potentially modulating global carbon and nutrient cycles through weathering feedbacks. Although conceptually appealing, geologic evidence for a direct climatic control on erosion has remained ambiguous owing to a lack of high-resolution, long-term terrestrial records and suitable field sites. Here we provide direct terrestrial field evidence for long-term synchrony between erosion rates and Milankovitch-driven, 400-kyr eccentricity cycles using a Plio-Pleistocene cosmogenic radionuclide paleo-erosion rate record from the southern Central Andes. The observed climate-erosion coupling across multiple orbital cycles, when combined with results from the intermediate complexity climate model CLIMBER-2, are consistent with the hypothesis that relatively modest fluctuations in precipitation can cause synchronous and nonlinear responses in erosion rates as landscapes adjust to ever-evolving hydrologic boundary conditions imposed by oscillating climate regimes.
... that lead to the present-day topographic configuration of the adjacent Eastern Cordillera commenced at ca. 15-10 Ma (Deeken et al., 2005;Siks & Horton, 2011;Pingel et al., 2013Pingel et al., , 2014Lapiana, 2021). ...
... be absent and tectonic activity shifted eastward to the present-day Eastern Cordillera DeCelles et al., 2011;Siks & Horton, 2011;Pearson et al., 2013;Pingel et al., 2013Pingel et al., , 2020Rahl et al., 2018;Lapiana, 2021) and the Sierras Pampeanas (Strecker et al., 1989;Mortimer et al., 2007;Löbens et al., 2013) and, to some extent, the southern Santa Bárbara System Arnous et al., 2020). Interestingly, the eastern boundary of the Eastern Cordillera deformed synchronously at ca. 10 Ma along its entire length from the Tilcara Ranges (23.5°S, ...
Article
Full-text available
The Andean Plateau of north‐western Argentina (Puna) at a mean elevation of ca. 4.2 km constitutes the southern continuation of the Altiplano; it is a compressional basin‐and‐range province comprising fault‐bounded, high‐elevation mountain ranges and largely internally drained basins with often thick sedimentary and volcaniclastic fill. Growing sedimentological and structural evidence supports the notion that the north‐western Argentine Andes between 22° and 26°S developed from an initial extensive broken‐foreland system that extended across the present‐day eastern Andean flank during the early to middle Eocene. However, compelling evidence of the tectonic history of this region is still missing. Here, we present new apatite fission track and zircon (U–Th)/He thermochronological data and U–Pb zircon ages from intercalated volcanic ash deposits from the Pastos Chicos Basin (23.5°S, 66.5°W) to constrain basin formation and the timing of major crustal deformation in the northern Puna. Inverse thermal modeling of the thermochronological data provides further temporal constraints on the late Cenozoic cooling history of the crust in this region and, by inference, on the timing of upper‐crustal shortening, range uplift, and basin formation in the northern sector of the present‐day Puna Plateau. Specifically, we argue for plateau‐wide distributed deformation in the Eocene between 23° and 24°S, followed by spatially disparate and diachronous deformation (Oligocene to Pliocene).
... North of 22°S, major deformation in the western Eastern Cordillera ceased by 9-10 Ma, based on undeformed strata that overlie folded Miocene sediments above a geomorphic surface (Allmendinger & Zapata, 2000;Gubbels et al., 1993). Farther south, the Andean thrust front moved to the east of the Humahuaca valley (blue star in Figure 1a) by at least 15-10 Ma (Henríquez et al., 2023;Pingel et al., 2013Pingel et al., , 2014Siks & Horton, 2011). Inversion of the Hornocal fault is proposed to have commenced in the middle to late Miocene (Henríquez et al., 2023;Lapiana, 2021;Siks & Horton, 2011), based on growth strata in upper Miocene syntectonic deposits (Siks & Horton, 2011). ...
Article
Full-text available
The present‐day deformation style of the Eastern Cordillera in NW Argentina is strongly influenced by the inversion of pre‐existing Paleozoic and Mesozoic structures. In particular, the extensional faults and lithological contrasts resulting from the Cretaceous–Paleogene Salta Rift phase form heterogeneities that were preferentially reactivated during the Andean orogeny. Constraining the timing and characteristics of reactivation is a key to understanding the interplay between tectonics and inherited crustal anisotropies. In this study, we combine structural and sedimentological field data with a low‐temperature thermochronology data set from the area surrounding the fault‐bounded Cianzo basin. The southeastern boundary is formed by the inverted Hornocal fault, which was the basin‐bounding normal fault of the Lomas de Olmedo sub‐basin (Salta Rift basin). Lacustrine deposits of the Yacoraite Formation overspill on the footwall of this fault and mark tectonic quiescence during the post‐rift phase of the Salta Rift. Apatite (U‐Th‐Sm)/He and fission track ages in the Hornocal fault hanging wall show an onset of rapid cooling interpreted to be concomitant with fault inversion between the latest Oligocene and middle Miocene (∼24–15 Ma). Low‐temperature thermochronology data also constrain the timing of major folding in the eastern limb of the Cianzo syncline to pre‐10 Ma, whereas the western limb started tilting post‐10 Ma. Characterization of exhumation patterns related to fault activity surrounding the Cianzo basin emphasizes the influence of the pre‐existing structural framework on deformation in fold‐and‐thrust belts.
... Por encima de esta unidad y en pasaje transicional se encuentra la Formación Río Grande del Mioceno medio (Siks y Horton, 2011;Lapiana et al., 2023), en paraconformidad se apoya la Formación Pisungo (Plioceno; González et al., 2004;Siks y Horton, 2011). Finalmente, completan la columna en discordancia los depósitos de la Formación Mal Paso de edad Plio-Pleistocena (Pingel et al., 2013;Fig. 2). ...
Article
Full-text available
Soft-sediment deformation structures (SSDS) develop when sediment, temporarily in a fluid-viscous state, deforms, resulting in the alteration of the original sedimentary structures (Lowe, 1975; Allen, 1982; Owen, 1987; Owen and Moretti, 2011). Deformation can occur rapidly, near the surface, during or after deposition (Allen, 1977; Allen, 1982). The main mechanisms through which a saturated granular material, initially in a solid state, transitions into a fluid-viscous state capable of mobilizing and dragging particles are referred to as liquefaction and fluidization (Lowe, 1975; Allen, 1982). The resulting structures by these processes include convolute bedding, load structures, clastic dikes, disturbed bedding, water escape structures, sand pipes and dish-and-pillar structures (Lowe, 1975; Owen, 1987; Rodríguez-Pascua et al., 2000; Jolly and Lonergan, 2002; Hurst et al., 2003; Schillizzi et al., 2010; Berra and Felletti, 2011; Sánchez et al., 2013). These structures are typically confined within discrete sedimentary layers, bounded by unaltered strata (Lowe, 1975; Hurst et al., 2003). The mechanisms that induce liquefaction and fluidization processes can be 'external' (allogenic) triggers, such as earthquakes, or 'internal' (autogenic) triggers, such as water waves, rapid sediment deposition, or groundwater movements (Allen and Banks, 1972; Obermeier, 1996; Owen and Moretti, 2011). The characteristics of the SSDS and their context provides insights into the mechanisms that initiated the processes associated with their formation. Furthermore, the stratigraphic distribution of these structures enables the identification of time intervals with recurring or non-recurring triggering agents. In the Casa Grande Formation (middle Eocene; Eastern Cordillera, Jujuy, Argentina), a suite of sedimentary structures has been described at different scales, interpreted as responses to contemporaneous tectonic deformation processes along with sedimentation. Among these structures, some have been recognized as products of sediment liquefaction and fluidization processes, as well as progressive folding of the substrate (Montero López et al., 2018). This lithostratigraphic unit, characterized by red to reddish-brown siltstones and sandstones, crops out along the Casa Grande valley, bounded to the east by the Sierra de Mal Paso and to the west by the Espinazo del Diablo in the northern sector of the Argentine Eastern Cordillera and its transition to the Puna region (Fig. 1). An intraformational unconformity (Figs. 1c, 2), led to propose a new subdivision into: Casa Grande I and Casa Grande II sequences (Montero López et al., 2018). The aim of this contribution is to present a comprehensive description and interpretation of specific discordant structures identified as sand pipes, within the Casa Grande II sequence (Figs. 3, 4). Additionally, we aim to discuss potential geological processes and factors that may have led to their formation. The sand pipes occur within two sand beds and are characterized by a vertical or subvertical disposition that cross-cut the depositional bedding. They exhibit an inverted funnel geometry in side view and tabular in plan view and are notable for their dimensions, which span from 0.72 to 2.10 m at each analyzed level, exhibiting consistent lateral distribution (Figs. 4, 5). The internal material of the structures is the same as that of the host rock, and based on their characteristics, two distinct zones can be identified (Figs. 5, 6, 7): a) central zone: corresponds to the internal part of the structures, characterized by medium-sized sands exhibiting a massive structure and, b) marginal zone: outer layer composed of finely laminated sandy material and, in some cases areas with finer material, displaying either a distinct vertical lamination or minor disturbances with folding features (Fig. 6). Considering the characteristics of the structures, the mechanism responsible for their formation are interpreted as liquefaction and fluidization. Similar attributes observed in other deposits are interpreted as the consequence of forceful upward water fluidization and dewatering from underlying induced by overpressurization (e.g., Chan et al., 2019; Espinoza et al., 2020). The fluid injection occurs as pressurized water is discharged from the lower layers, fluidizing the sediments and the area around them, in multiple points, along its upward path (Draganits et al., 2003; Wheatley et al., 2016; Chan et al., 2019; Fig. 8a). The development of a central and marginal zone is interpreted as a result of a flow velocity gradient during the injection (Ross et al., 2011, 2014; Fig. 8b). The presence of a marginal zone is consistent with shear between the fluidized flow and the surrounding sediment (Deynoux et al., 1990; Draganits and Janda, 2003; Draganits et al., 2003; Chan et al., 2019) which induced grain alignment parallel with the direction of the flow; on the other hand, the central zone is massive which is consistent with high flow velocities (Fig. 8c). Among the potential triggers for these processes a seismic event is interpreted following the criteria proposed by Owen and Moretti (2011). These criteria include vertical repetition, lateral continuity, morphology comparable with structures formed by recent earthquakes and the geological context (i.e., proximity to faults).
... A provenance change occurred at 4.5 Ma, that is interpreted to have arisen due to the uplift resulting as a result of activation of the Pescado fault(Echavarria et al., 2003;Hernández et al., 1996). Accelerated uplift and erosion has been interpreted in Río Iruya and across the region in nearby sites such as the Humauacauna region, between 4-2.3 Ma due to out of sequence faulting(Pingel et al., 2013, Amidon et al., 2015. The Proto-Río Iruya river system gained a Puna source during this time interval based on detrital zircon analysis(Amidon et al., 2015). ...
Thesis
Full-text available
The impact of climate change on vegetation dynamics and fire regimes has been the subject of much interest in determining how anthropogenic climate change might alter ecological communities. Here, I study the C4 grassland expansion during the late Miocene-Pliocene epoch in N.W. Argentina from paleosols from three terrestrial sedimentary basins- Palo Pintado, La Viña and Río Iruya. I hypothesize that increased seasonality associated with the strengthening of the South American Summer Monsoon drove the spread of C4 grasses and increased fire frequency, leading to a positive feedback between increased C4 cover and fire regimes. We test this hypothesis using multiproxy analysis using novel biomarkers (i.e. polycyclic aromatic hydrocarbons) as a proxy for fire frequency and vegetation type, carbon isotopic analysis of bulk organic matter as a proxy for C4 vegetation and bulk elemental geochemical analysis of paleosols as a proxy of mean annual precipitation and temperature. We observe the appearance of ~50% C4 cover in La Viña and Río Iruya prior to the Miocene-Pliocene boundary. Precipitation appears to play an important role in the spread of C4 cover in both La Viña with decreasing precipitation resulting in the greater spread of C4 cover. We do not observe a significant change in fire regimes during this spread in C4 vegetation at La Viña and Río Iruya. We establish that the lack C4-fire feedback is either due to the lack of high enough C4 cover, or the absence of this feedback mechanism in late Miocene-Pliocene South America. However, we observe trends in gymnosperm proliferation with changes in fire regimes, with one proxy showing a decrease in gymnosperm proliferation with increasing fire frequency and two proxies showing an increase in gymnosperm proliferation with decreasing fire frequency.
Preprint
Full-text available
Theory suggests that the response time of alluvial channel long-profiles to perturbations in climate is related to the magnitude of the forcing and the length of the system. Shorter systems may record a higher frequency of forcing compared to longer systems. Empirical field evidence that system length plays a role in the climate periodicity preserved within the sedimentary record is, however, sparse. The Toro Basin in the Eastern Cordillera of NW Argentina provides an opportunity to test these theoretical relationships as this single source-to-sink system contains a range of sediment deposits, located at varying distances from the source. A suite of eight alluvial fan deposits is preserved along the western flanks of the Sierra de Pascha. Farther downstream, a flight of cut-and-fill terraces have been linked to eccentricity-driven (100-kyr) climate cycles since ca. 500 ka. We applied cosmogenic radionuclide (10Be) exposure dating to the fan surfaces to explore (1) how channel responses to external perturbations may or may not propagate downstream, and (2) the differences in landscape response to forcing frequency as a function of channel length. We identified two generations of fan surfaces: the first (G1) records surface activity and abandonment between ca. 800 and 500 ka and the second (G2) within the last 100 kyr. G1 fans record a prolonged phase of net incision, which has been recognised throughout the Central Andes, and was likely triggered by enhanced 100-kyr global glacial cycles following the Mid-Pleistocene Transition. Relative fan surface stability followed, while 100-kyr cut-and-fill cycles occurred downstream, suggesting a disconnect in behaviour between the two channel reaches. G2 fans record higher frequency climate forcing, possibly the result of precessional forcing of climate (ca. 21/40-kyr timescales). The lack of a high-frequency signal farther downstream provides field support for theoretical predictions of a filtering of high-frequency climate forcing with increasing channel length. We show that multiple climate periodicities can be preserved within the sedimentary record of a single basin. Differences in the timing of alluvial fan and fluvial terrace development in the Toro Basin appears to be associated with how channel length affects fluvial response times to climate forcing as well as local controls on net incision, such as tectonic deformation.
Article
Full-text available
We describe new Upper Miocene mammalian specimens from the Maimara Formation (Late Miocene to Early Pliocene) exposed at Humahuaca Basin (23°–24°S), northwestern Argentina (NWA), and analyse their taxonomy and relevance for our understanding of the initial stages of the Great American Biotic Interchange (GABI). The stratigraphical and geochronological control of the studied specimens indicates a time window of c. 6.6–5.8 Ma. These data are crucial for establishing the oldest records of the Holarctic immigrants of the GABI. The first record of cricetid rodents from the Maimara Formation is reported. Moreover, with an age of c. 6 Ma it is the first appearance datum (FAD) of these rodents in South America. The age of the procyonid Cyonasua recorded in this unit is estimated between c. 6.6 and c. 6.4 Ma. The record of procyonids and cricetids in the same continuous sedimentary sequence suggests that the time interval between the dispersion of both groups into the continent during GABI was c. 1 myr. Of the autochthonous mammals from the Maimara Formation, the first records of litopterns, chlamyphorid and mylodontine xenarthrans, and caviomorph rodents (including new species Pithanotomys? solisae and Palaeocavia humahuaquense) are described. FADs of immigrant and autochthonous mammals in NWA suggest a major faunal turnover during GABI at the Messinian, probably linked to global climatic changes and the tectonic activity that affected the Humahuaca Basin at that time. Hypsodonty, a dominant feature among Maimaran mammals, may be related to a landscape strongly influenced by the Andean uplift.
Article
Full-text available
The Andes orogenic belt is one of the most tectonically active zones located between the convergent margins of the South American plate and the Pacific plate. From the dissected western belt to the expanded alluvial plain of Amazon basin, the Madeira watershed bears recent tectonic imprints. The tectonic signature is well documented by the spatial and linear characteristics of the five sub watersheds along with the Madeira trunk river and channels. The level of tectonic activeness has been assessed on the sub watersheds of the Madeira River, which is one of the main right-hand tributaries of the Amazon River. On the five defined sub-watersheds and the main channel watershed of Madeira, several morphometric indices, viz., asymmetry factor (AF), tilt angle (β), sinuosity index (SI), channel concavity (ɵ), elongation ratio (Re), circularity ratio (Rc), transverse topographic symmetry factor (T) and basin shape index (Bs), have been used. To rank the watersheds based on tectonic activity, we applied the Technique for Order Preference by Similarity to Ideal Solution (TOPSIS) The active tectonics was also tested by a comparative study of linear indicators as Hack profiles, longitudinal profiles and segment-wise stream gradient index (SL) of watersheds 2 (rank 1) and 3 (rank 6). The results reveal that the most tectonically active watershed 2 is located in a high seismic magnitude zone with several deep earthquake epicenters. The least tectonically active watershed 6 is located in a low seismic zone.
Preprint
Full-text available
The Andean foreland is divided into morphotectonic provinces characterized by diverse deformation styles and seismogenic behavior partially stemming from distinct geological histories that preceded the current phase of subduction. The transition between the high Andes and the eastern foreland is exposed to numerous natural hazards and contains critical economic infrastructure, yet we know relatively little about regional active tectonics due to few geophysical investigations. Here we use waveforms collected during a 15-month-long seismic network deployment in the Santa Bárbara System (SBS) of northwest Argentina following the 2015 Mw 5.7 El Galpón earthquake to determine the distribution and magnitude of local earthquakes, obtain a regional 1D seismic velocity model, and improve our overall understanding of SBS neotectonics. Of the nearly 1200 recorded earthquakes, ~700 occurred in the crust with half of the moment release associated with events deeper than 25 km. The depth extent of seismicity supports the notion that the SBS upper and middle crust are homogeneous and that the lower crust is composed of granulites. These conditions likely formed during Paleozoic mountain building and Salta Rift-related Cretaceous magmatism, which dehydrated the crust. We find no clear indications that a shallow, low-angle detachment fault inferred to have been active during Cretaceous rifting exerts a strong control on modern deformation in contrast to the active décollement beneath the adjacent fold-and-thrust belt of the Subandes to the north. It remains unclear how active, inverted normal faults in the SBS shallow crust connect to the deeper zones of seismicity.
Article
Geomorphology of the Salinas Grandes-Quebrada de Humahuaca area, Jujuy Province. Landforms in the area between Salinas Grandes and Quebrada de Humahuaca (23°-24°S and 65°-66°W), Jujuy Province, are related to fluvial process and, to a lesser extent, also to mass wasting and aeolian process. In higher zones, above 4500 m, moraines and other glacigenic features, such as moraines, glacial valleys and cryogenic structures are also evident. The relationships between regional geomorphological evolution, the tectonic environment and the dynamic evolution of the latter were a particular focus of study, likewise the impact of regional climatic oscillations on geomorphological development. The climatic and structural conditions show a great variation for which reason, the area is particularly interesting for examining their interrelationships in the process of landscape evolution.
Article
The red deposits of Salta Group (Cretaceous-Paleogene) which outcrops in the quebrada de Tumbaya, right margin of Quebrada de Humahuaca (province of Jujuy), they are characterized by facies that differs from the typical ones in their places of the Salta Group basin. They are proximal facies, characterized by arid and semi-arid conditions whose accumulation advanced on the N-NW side of the Salto-Jujuy high structure. Salta Group is separated from the underlying Cambrian Meson Group by an erosive discordance. An angular discordance at the top of the Group separates it from the superposed Quaternary deposits. In Tumbaya, the two main units of Salta Group, the synrift and post-rift, are represented. The research and study of sedimentary facies allowed to identify and to characterize the three Subgroups (Pirgua, Balbuena and Santa Bárbara) and to differentiate their formations. The Pirgua Subgroup consists of eolian sandstones; it is not possible to differentiate its three formation units, therefore it is mentioned as Pirgua Formation. The Balbuena Subgroup is subdivided into two formations, Lecho and Yacoraite Formations. The Lecho Formation is characterized by sheet floods facies and the Yacoraite Formation displays transitional, continental and marine litoral-sublitoral carbonatic environmental conditions influenced by tidal currents and storms. The three formations of Santa Barbara Subgroup are present: (Mealla, Maíz Gordo and Lumbrera), which represent a fining-upward fluvial succession with development of multiple paleosoils; there have also been identified sheet floods and floodplain facies.