ArticlePDF Available

Abstract and Figures

We explore canyon incision history of the western margin of the Andean (Altiplano-Puna) plateau in the central Andes as a proxy for surface uplift. (U-Th)/He apatite data show rapid cooling beginning at ca. 9 Ma and continuing to ca. 5.1 Ma in response to incision. A minimum of 1.0 km of incision took place during that interval. The youngest apatite date and a volcanic fl ow perched 125 m above the present valley fl oor dated at 2.261 ± 0.046 Ma (40Ar/39Ar) show that an additional ~1.4 km of incision occurred between ca. 5.1 and 2.3 Ma. Thus, we infer that a total of at least 2.4 km, or 75% of the present canyon depth was incised after ca. 9 Ma. (U-Th)/He zircon data collected along the same transect imply that the western margin of the plateau was warped upward into its present monoclinal form, rather than uplift being accommodated on major surface-breaking faults.
Content may be subject to copyright.
GEOLOGY, June 2007 523
Geology, June 2007; v. 35; no. 6; p. 523–526; doi: 10.1130/G23532A.1; 5 fi gures; Data Repository item 2007124.
© 2007 The Geological Society of America. For permission to copy, contact Copyright Permissions, GSA, or editing@geosociety.org.
ABSTRACT
We explore canyon incision history of the western margin of the
Andean (Altiplano-Puna) plateau in the central Andes as a proxy for
surface uplift. (U-Th)/He apatite data show rapid cooling beginning
at ca. 9 Ma and continuing to ca. 5.1 Ma in response to incision. A
minimum of 1.0 km of incision took place during that interval. The
youngest apatite date and a volcanic fl ow perched 125 m above the
present valley fl oor dated at 2.261 ± 0.046 Ma (40Ar/39Ar) show that an
additional ~1.4 km of incision occurred between ca. 5.1 and 2.3 Ma.
Thus, we infer that a total of at least 2.4 km, or 75% of the present
canyon depth was incised after ca. 9 Ma. (U-Th)/He zircon data col-
lected along the same transect imply that the western margin of the
plateau was warped upward into its present monoclinal form, rather
than uplift being accommodated on major surface-breaking faults.
Keywords: Altiplano, geochronology, tectonics, Peru, geomorphology,
helium.
INTRODUCTION
Understanding the development of the Central Andean plateau is
crucial to evolutionary models of both Andean geodynamics and regional
climate patterns. (Note: We use “Central Andean plateau” as defi ned by
Allmendinger et al., 1997, to represent the region above the 3 km eleva-
tion contour between 13°S and 27°S. This includes the Altiplano and Puna
plateaus and portions of the Western and Eastern Cordilleras.) Although
episodes of central Andean deformation are constrained in many regions,
it is often diffi cult or impossible to discern the magnitude of plateau uplift
based on deformation history alone. Different approaches for estimating
paleoelevation or the existence of high topography have led to a broad
range of proposed uplift histories, but precise constraints are lacking. Oligo-
cene uplift probably generated less than half of the central Andean relief
seen today (e.g., Gubbels et al., 1993; Kennan, 2000). Numerous lines
of evidence from the plateau and its eastern margin point to additional
surface uplift starting at ca. 10 Ma, with magnitudes ranging from at least
1 km to as much as 3.5 km (e.g., Kennan et al., 1997; Lamb and Hoke,
1997; Barke and Lamb, 2006; Garzione et al., 2006; Ghosh et al., 2006).
In northern Chile, Hoke (2006) and Nestor et al. (2006) estimate 1–1.4 km
of western margin uplift after 10 Ma, and Wörner et al. (2000) argue for
termination of uplift by 2.7 Ma. More recently, interpretations of oxygen
and clumped isotope data have led to estimates that the plateau reached
its present height by ca. 6 Ma (Garzione et al., 2006; Ghosh et al., 2006).
Uncertainties in all these estimates require that independent measures of
surface uplift be made before we can accurately constrain uplift history.
In southwestern Peru, large rivers cut deep canyons through the west-
ern margin of the Central Andean plateau. Cotahuasi-Ocoña Canyon is
the deepest of these, incising more than 3 km below the plateau surface
(Fig. 1). The western margin of the plateau has been characterized by
a semiarid-hyperarid climate for at least the past 15 m.y. (e.g., Hartley,
2003; Rech et al., 2006). As expected, given the arid climate and gentle
slopes of the region, Kober et al. (2007) found that millennial-scale inter-
uve erosion rates are uniformly low, despite a gradual increase in erosion
rate with elevation (<0.001–0.05 mm/yr). As predicted from these studies,
and as seen in the fi eld, little surface erosion has occurred on canyon inter-
uves in our study area, making canyon incision an excellent proxy for
surface uplift. Incision depth provides a minimum estimate for the amount
of surface uplift, while the onset of incision provides a minimum age for
the timing of uplift. We use 40Ar/39Ar dates of volcanic fl ows and bedrock
(U-Th)/He thermochronologic data to explore canyon evolution. Detailed
information regarding geochronologic data and their interpretation are
provided in the GSA Data Repository1 (Tables DR1–DR4).
VALLEY-FILLING VOLCANICS
Dates of valley-fi lling volcanic fl ows give a minimum time by which
the canyon had incised to the depth of the sampled fl ow. In the middle
reaches of Cotahuasi-Ocoña Canyon, a whole-rock 40Ar/39Ar date of 2.261
± 0.046 Ma for a basaltic andesite fl ow sampled 125 m above the present
valley fl oor (05TS38, Fig. 1) shows that ~96% of the incision in that sec-
tion of the canyon (3.2 km total local incision) happened before 2.3 Ma.
Farther upvalley, an ignimbrite (05TS25) perched ~400 m above the pres-
ent valley shows that ~75% of the canyon depth (1.6 km total local inci-
sion) was cut before 3.825 ± 0.016 Ma. Thouret et al. (2005) obtained a
40Ar/39Ar date of 3.76 ± 0.14 Ma for the same fl ow and interpreted it as a
maximum age for most of the valley incision.
LOW-TEMPERATURE THERMOCHRONOLOGY
In this setting where background erosion rates are very slow, we
expect thermochronometers to yield very old dates. However, canyons are
sites of localized, potentially rapid exhumation. When a canyon is incised,
perturbations to near-surface isotherms result in rapid cooling of bedrock
below the canyon bottom (Fig. 2). Because the closure temperature iso-
therm for the apatite (U-Th)/He system is only ~2–3 km below the sur-
face (assuming a geothermal gradient of 20–30 °C/km, Farley, 2000), we
expect to fi nd young, rapidly cooled apatites below this level in the deep-
est reaches of Cotahuasi-Ocoña Canyon. Their young ages should refl ect
both downward movement of the closure isotherm and relative upward
advection of rock due to erosion of the surrounding surface, though the
latter component is small in this setting as discussed above. The oldest
date among this young suite of ages should refl ect the initiation of the
thermal response to canyon incision.
We collected most samples for apatite (U-Th)/He dating along a
75 km, canyon-bottom transect, far from the infl uence of volcanic activity
(Figs. 1 and DR1). Additional samples were collected from the upper-
1GSA Data Repository item 2007124, Figure DR1 (map of volcanic units
and generation of paleosurface), Table DR1 (apatite-He data), Table DR2 ( zircon-
He data), Table DR3 (40Ar/39Ar data), Table DR4 (40Ar/39Ar summary), and
Table DR5 (thermal model parameters and results), is available online at www.
geosociety.org/pubs/ft2007.htm, or on request from editing@geosociety.org or
Documents Secretary, GSA, P.O. Box 9140, Boulder, CO 80301, USA.
Uplift of the western margin of the Andean plateau revealed
from canyon incision history, southern Peru
Taylor F. Schildgen Department of Earth, Atmospheric, and Planetary Sciences, Massachusetts Institute of Technology,
Cambridge, Massachusetts 02139, USA
Kip V. Hodges
Kelin X Whipple School of Earth and Space Exploration, Arizona State University, Tempe, Arizona 85287, USA
Peter W. Reiners Department of Geosciences, University of Arizona, Tucson, Arizona 85721, USA
Malcolm S. Pringle Department of Earth, Atmospheric, and Planetary Sciences, Massachusetts Institute of Technology,
Cambridge, Massachusetts 02139, USA
524 GEOLOGY, June 2007
most catchment where there are abundant volcanic fl ows dated at ca. 1.4–
3.8 Ma (Thouret et al., 2005). We constrained canyon depth by measur-
ing the distance below a paleosurface that is preserved today as bedrock
remnants beneath the regionally blanketing 14–16 Ma Huaylillas Ignim-
brite (Thouret et al., 2005, and dates reported here). Positions of remnants
exposed in valley walls were digitized and a spline surface warped to fi t
the points, giving an approximation for the bedrock surface (Fig. DR1).
Samples were collected from ~1 km below the paleosurface near the coast,
down to ~3.1 km below in the deepest reaches of the canyon.
The resulting dates are ca. 60 Ma near the coast and rapidly decrease
to ca. 9 Ma in the middle reaches of the canyon. Farther upvalley, there
is a much more gradual change in dates with distance, with the youngest
sample away from the volcanic-dominated upper catchment dated at 5.1 Ma
(Fig. 3). Samples from the uppermost catchment (05TS07 and samples
upstream from there) range from <1 to 3.8 Ma. Because these samples are
as young as the abundant volcanic fl ows in that region, we suspect their ages
have been reset or partially reset by recent volcanism. For this reason, we
focus the remainder of our discussion on samples from the middle and lower
catchment, which we believe have not been affected by reheating.
A plot of canyon depth versus sample cooling age (Fig. 4) can be inter-
preted the way that vertical profi les of thermochronologic data are typically
interpreted: as indicative of how exhumation rates changed over the time
interval represented by the measured ages. Although such an interpretation
of Figure 4 may be compromised by variations in bedrock thermal structure,
there is very little evidence of large lateral variations in upper-crustal tem-
peratures in this region of southern Peru. For example, reported variations
in measured surface heat fl ow range only from 32 to 44 mW/m2 (Henry and
Pollack, 1988; Hamza and Muñoz, 1996; Springer and Förster, 1998).
Figure 1. Location map and sample sites. Topography is shown as a
gray-scale digital elevation model (DEM) draped over shaded relief
from 90-m-resolution Shuttle Radar Topography Mission (SRTM) data.
Inset shows location relative to South American geography. (U-Th)/He
samples are collected from near river level. 40Ar/39Ar samples are from
elevations noted in GSA Data Repository (see footnote 1).
Figure 2. Schematic interpretation of incision history based on (U-Th)/He
apatite data. Gray zones are rapidly cooled due to migration of the
closure temperature isotherm (~70 °C) in response to incision. At least
1.0 km of incision occurred between 9 and 5.1 Ma, since this is the
thickness of the rapidly cooled zone between those ages on the age-
depth plot (Fig. 4). Post–5.1 Ma incision is equal to the depth of the
closure temperature isotherm below the valley bottom at 5.1 Ma.
Figure 3. Zircon and apatite results plotted as distance measured
perpendicular to coastline. Errors in age show 2
σ
uncertainty.
Figure 4. Apatite and zircon (U-Th)/He ages collected along valley at
river level. Mean ages are weighted by the inverse of the variance
of individual crystal ages. Errors in age are plotted as 2
σ
analyti-
cal errors only. Errors in depth below paleosurface are estimated at
±100 m. Inset shows apatite data over the last 25 m.y.
GEOLOGY, June 2007 525
Ignoring the minor infl uences of such variations in heat fl ow, our
apatite data imply slow background erosion of only ~0.7 km from ca. 60 to
9 Ma (0.01 km/m.y.) and a change to rapid incision of at least 1.0 km from
ca. 9 to 5.1 Ma (0.26 km/m.y.). The latter is a minimum estimate because
isotherm movement is damped compared to changes in surface topogra-
phy. Additional incision must have occurred after 5.1 Ma to exhume this
youngest apatite. Because a 2.3 Ma volcanic fl ow is perched 125 m above
the valley fl oor in the middle reaches of the canyon, we know that the
young apatite must have been exhumed from its closure depth to the near
surface between 5.1 and 2.3 Ma. We explored probable erosion rates and
depths to the closure temperature for this scenario using M. Brandon’s
computer code AGE2EDOT (summarized in Ehlers et al., 2005). Assum-
ing typical thermal properties of the crust (Table DR5) and an initially
uncompressed geothermal gradient between 20 and 30 °C/km, exhuma-
tion of the youngest apatite requires incision between 0.7 and 0.5 km/m.y.,
an effective closure temperature of 72 °C, and corresponding depths of
2.0–1.4 km to the closure temperature. Because we expect even greater
compression of isotherms below a canyon compared to the compression
expected from simply increasing erosion rate over fl at topography as
AGE2EDOT assumes, we choose the lower end of this range as an esti-
mate. Thus, we infer that ~1.4 km of material was removed in the middle
reaches of the canyon between 5.1 and 2.3 Ma (average 0.5 mm/yr) to
exhume the youngest apatite.
Collectively, the data suggest that the middle section of Cotahuasi-
Ocoña Canyon was incised at least 1.0 km between ca. 9 and 5.1 Ma and
~1.4 km between 5.1 and 2.3 Ma. This represents incision of at least 75%
of the total canyon depth since 9 Ma, and 44% since 5.1 Ma.
INTERPRETING SURFACE UPLIFT FROM
THERMOCHRONOLOGIC DATA
Deciphering surface uplift from thermochronologic interpretations
of canyon incision is complicated by several factors. First, incision can
refl ect a transient response to tectonically driven uplift and/or climatically
driven changes in sediment supply and fl uvial discharge, provided prior
surface uplift had produced suffi cient topographic relief. We interpret the
latest phase of incision to be a response to uplift because it is the simplest
explanation, and there is no evidence for a climate change that could induce
a pulse of incision at this time. The interpreted drying of climate starting at
ca. 15 Ma (e.g., Hartley, 2003; Rech et al., 2006) should have the opposite
effect of deterring incision, rather than generating a pulse of incision into
pre-existing topography. Also, our estimate for the onset of incision cor-
relates well with tectonic evidence in northern Chile for surface uplift after
ca. 10 Ma (Nestor et al., 2006). Second, although incision magnitude is
limited by uplift magnitude, incision rates are not necessarily closely tied
to uplift rates. The incision history we interpret from thermochronologic
data is offset from tectonically driven uplift by lag times defi ned by both
the geomorphic response time to uplift and the thermal response time to
incision. Constraining these response times is critical in order to quantita-
tively extract uplift history from canyon incision.
By “geomorphic response time,” we mean the time scale over which
rivers respond to a pulse of uplift. When uplift occurs, a river steepens
at its outlet. Over time, this steep segment migrates upstream at a rate
that depends principally on whether incision is transport- or detachment-
limited (Whipple and Tucker, 2002). In a detachment-limited system, the
pulse of incision will move through the system vertically at a rate that
approximates the new uplift rate (Niemann et al., 2001), which appears
to be ~0.5 mm/yr in Cotahuasi-Ocoña Canyon. In transport-limited sys-
tems, the incision pulse migrates upstream much more rapidly, resulting
in essentially uniform onset of incision throughout the catchment on geo-
logic time scales (Whipple and Tucker, 2002). Because we sampled from
the trunk stream at large drainage area and thus both relatively low in ele-
vation and most likely to approach transport-limited conditions, we expect
incision across the sampled region to have a relatively short response time.
Even if incision were detachment-limited, the highest bedrock sampled
for apatite (852 m) would imply a maximum lag time of 2 m.y..
We investigated the thermal response time to incision using M. Bran-
don’s RESPTIME computer code (summarized in Ehlers et al., 2005).
Although this code was written to evaluate thermal fi eld transients result-
ing from changes in broad-scale erosion rates, it is also useful for exploring
general effects of localized incision. We tested the thermal response time
to a sudden increase in erosion rate from 0.01 mm/yr to 0.5 mm/yr. Using
typical values for thermal properties of the crust (Table DR5), results show
that after an initial slow response, the migration rate of the apatite-He
closure temperature isotherm reaches 75% of the incision rate 0.7 m.y.
after incision starts. Thus, isotherm migration is signifi cantly slower than
incision for the fi rst ~1 m.y.
Response time estimates show two ways in which our thermo-
chronologic data do not directly refl ect uplift. First, the age of the “kink”
in the age-depth profi le (Fig. 4) is likely to lag behind onset of uplift pri-
marily as a result of geomorphic response time, which we expect to be a
maximum of 2 m.y. Second, the damped response of isotherm movement
in response to incision means that we derive only a minimum estimate of
uplift rate and magnitude. Given these limitations, our estimate for the
onset of uplift is likely good to ~2 m.y., and our estimate for uplift magni-
tude should be a robust minimum.
STYLE OF TECTONIC DEFORMATION
Compared to apatite, the higher closure temperature to helium dif-
fusion in zircon (~180 °C, Reiners et al., 2004) means that (U-Th)/He
zircon ages are set much deeper in the crust (~7 km) where they are not
signifi cantly affected by localized, near-surface thermal effects of canyon
incision. The pattern of zircon cooling ages should therefore refl ect only
regional exhumation and postclosure bedrock deformation. Zircon dates
of samples collected along the same valley-bottom transect show a regular
but scattered progression from older dates near the coast to younger dates
farther upvalley (Fig. 3).
If the latest pulse of uplift recorded by canyon incision were accom-
modated through localized, surface-breaking fault movement, we would
expect to see a step-change in the pattern of zircon cooling ages (Fig. 5).
Although we see no evidence for a large step, the present data set cannot
rule out the possibility of uplift accommodated on a series of small faults
distributed over a wide region. We have not seen distributed faulting in the
eld, but we mapped a focused shear zone at the range front. A 16.12 ±
0.04 Ma (40Ar/39Ar) undeformed ignimbrite crossing what appears to be
the equivalent shear zone 100 km to the southeast of the fi eld area suggests
that signifi cant movement on this structure predates the inferred ca. 9 Ma
uplift. These thermochronologic data lead us to interpret the latest pulse
of uplift as resulting from monoclinal warping of the western margin, con-
sistent with interpretations presented by Isacks (1988).
Surface
topog-
raphy
Figure 5. Predictions for patterns of (U-Th)/He zircon ages for sam-
ples collected along river profi le in various tectonic settings, assum-
ing initially fl at isochrons. The pattern of ages presented in this
study best matches the prediction resulting from a monocline.
526 GEOLOGY, June 2007
GEODYNAMIC IMPLICATIONS AND CONCLUSIONS
One of the most important characteristics of the data presented here
is the evidence they provide for onset of at least 2.4 km of river incision
at ca. 9 Ma, in broad agreement with the magnitude and timing of recent
uplift of the Central Andean plateau estimated by different methods. If,
as we suggest, this incision corresponds to a phase of surface uplift, what
can the age range tell us about plausible causes of uplift? Upper-crustal
thickening from shortening at the plateau margins mostly occurred prior
to 10 Ma (e.g., Gubbels et al., 1993; Victor et al., 2004; Farías et al., 2005;
Elger et al., 2005). The argument for uplift through lithospheric delamina-
tion is well supported at the southern end of the plateau, where thinned
lithosphere and widespread mafi c volcanism point to delamination as a
plausible cause (Kay and Kay, 1993; Schurr et al., 2006). In the north,
however, strong evidence is lacking. Seismic data (Beck and Zandt, 2002)
suggest there may be thinned lithosphere beneath the eastern portion of
the plateau, but not under the western region. There is also no evidence
for a volcanic fl are-up close to the time of uplift (James and Sacks, 1999).
Ignimbrite volcanism in southern Peru from ca. 16 to 14 Ma may be related
to a small delamination event, although we expect canyon incision would
not lag more than 2 m.y. behind signifi cant uplift. Ductile thickening of
the middle to lower crust provides a likely alternative, possibly triggered
by displacement of material due to underthrusting of the Brazilian Shield
beneath the eastern margin, as proposed by Isacks (1988). The ca. 10 Ma
initiation of underthrusting (Gubbels et al., 1993) correlates well with the
latest phase of uplift inferred from our data. For these reasons, although
the data we present on uplift timing and monocline deformation are con-
sistent with either lithospheric delamination or lateral fl ow of lower crust,
we favor the latter hypothesis to explain this latest phase of uplift.
ACKNOWLEDGMENTS
We thank T. Smith, K. Cornell, and J. Bradley for help with fi eldwork;
S. Nicolescu for help with (U-Th)/He analyses; G. Wörner for discussions; T. Sempere,
O. Oncken, and an anonymous reviewer for constructive reviews. This work is
supported by National Science Foundation Tectonics Division grant EAR-0409359.
REFERENCES CITED
Allmendinger, R.W., Jordan, T.E., Kay, S.M., and Isacks, B.L., 1997, The evolu-
tion of the Altiplano-Puna plateau of the central Andes: Annual Review of
Earth and Planetary Sciences, v. 25, p. 139–174.
Barke, R., and Lamb, S., 2006, Late Cenozoic uplift of the Eastern Cordillera,
Bolivian Andes: Earth and Planetary Science Letters, v. 249, p. 350–367,
doi: 10.1016/j.epsl.2006.07.012.
Beck, S.L., and Zandt, G., 2002, The nature of orogenic crust in the central Andes:
Journal of Geophysical Research, v. 107, B10, doi: 10.1029/2000JB000124.
Ehlers, T.A., Chaudhri, T., Kumar, S., Fuller, C.W., Willett, S.D., Ketcham,
R.A., Brandon, M.T., Belton, D.X., Kohn, B.P., Gleadow, A.J.W., Dunai,
T.J., and Fu, F.Q., 2005, Computational tools for low-temperature thermo-
chronometer interpretation: Reviews in Mineralogy and Geochemistry,
v. 58, p. 589–622, doi: 10.2138/rmg.2005.58.22.
Elger, K., Oncken, O., and Glodny, J., 2005, Plateau-style accumulation of deformation,
southern Altiplano: Tectonics, v. 24, TC4020, doi: 10.1029/2004TC001675.
Farías, M., Charrier, R., Compte, D., Martinod, J., and Hérail, G., 2005, Late
Cenozoic deformation and uplift of the western fl ank of the Altiplano: Evi-
dence from the depositional, tectonic, and geomorphic evolution and shal-
low seismic activity (northern Chile at 19º30S): Tectonics, v. 24, TC4001,
doi: 10.1029/2004TC001667.
Farley, K.A., 2000, Helium diffusion from apatite: General behavior as illustrated
by Durango fl uorapatite: Journal of Geophysical Research, v. 105, no. B2,
p. 2903–2914, doi: 10.1029/1999JB900348.
Garzione, C.N., Molnar, P., Libarkin, J.C., and MacFadden, B.J., 2006, Rapid
late Miocene rise of the Bolivian Altiplano: Evidence for removal of mantle
lithosphere: Earth and Planetary Science Letters, v. 241, p. 543–556, doi:
10.1016/j.epsl.2005.11.026.
Ghosh, P., Garzione, C.N., and Eiler, J.M., 2006, Rapid uplift of the Altiplano
revealed through 13C-18O bonds in paleosol carbonates: Science, v. 311,
p. 511–514, doi: 10.1126/science.1119365.
Gubbels, T.L., Isacks, B.L., and Farrar, E., 1993, High-level surfaces, plateau
uplift, and foreland development, Bolivian central Andes: Geology, v. 21,
p. 695–698, doi: 10.1130/0091-7613(1993)021<0695:HLSPUA>2.3.CO;2.
Hamza, V.M., and Muñoz, M., 1996, Heat fl ow map of South America: Geo-
thermics, v. 25, no. 6, p. 599–646, doi: 10.1016/S0375-6505(96)00025-9.
Hartley, A.J., 2003, Andean uplift and climate change: Geological Society
[ London] Journal, v. 160, p. 7–10.
Henry, S.G., and Pollack, H.N., 1988, Terrestrial heat fl ow above the Andean sub-
duction zone in Bolivia and Peru: Journal of Geophysical Research, v. 93,
no. B12, p. 15,153–15,162.
Hoke, G.H., 2006, The infl uence of climate and tectonics on the geomorphology
of the western slope of the central Andes, Chile and Peru [Ph.D. thesis]:
Ithaca, New York, Cornell University, 283 p.
Isacks, B.L., 1988, Uplift of the Central Andean plateau and bending of the Boliv-
ian Orocline: Journal of Geophysical Research, v. 93, no. B4, p. 3211–3231.
James, D.E., and Sacks, I.S., 1999, Cenozoic formation of the central Andes: A
geophysical perspective, in Skinner, B.J., ed., Geology and ore deposits of the
central Andes: Society of Economic Geology Special Publication 7, p. 1–26.
Kay, R.W., and Kay, S.M., 1993, Delamination and delamination magmatism:
Tectonophysics, v. 219, p. 177–189, doi: 10.1016/0040-1951(93)90295-U.
Kennan, L., 2000, Large-scale geomorphology of the Andes: Interrelationships
of tectonics, magmatism and climate, in Summerfi eld, M.A., ed., Geomor-
phology and global tectonics: New York, John Wiley, p. 167–199.
Kennan, L., Lamb, S., and Hoke, L., 1997, High-altitude palaeosurfaces in the
Bolivian Andes: Evidence for late Cenozoic surface uplift, in Widdowson,
M., ed., Palaeosurfaces: recognition, reconstruction and paleoenvironmen-
tal interpretation: Geological Society [London] Special Publication 120,
p. 307–324.
Kober, F., Ivy-Ochs, S., Schlunegger, F., Baur, H., Kubik, P.W., and Wieler, R.,
2007, Denudation rates and a topography-driven rainfall threshold in north-
ern Chile: Multiple cosmogenic nuclide data and sediment yield budgets:
Geomorphology, v. 83, p. 97–120, doi: 10.1016/j.geomorph.2006.06.029.
Lamb, S., and Hoke, L., 1997, Origin of the high plateau in the central
Andes, Bolivia, South America: Tectonics, v. 16, no. 4, p. 623–649, doi:
10.1029/97TC00495.
Nestor, P.L., Jordan, T.E., Blanco, N., Hoke, G.D., and Tomlinson, A.J., 2006,
Evidence for late Miocene uplift by long-wavelength rotation of western
ank of Altiplano segment of central Andes 20º30–20º30S, Chile, in Pro-
ceedings, Geological Society of America and Asociación Geológica Argen-
tina, Backbone of the Americas Conference, 3-7 April 2006, Mendoza,
Argentina.
Niemann, J.D., Gasparini, N.M., Tucker, G.E., and Bras, R.L., 2001, A quantita-
tive evaluation of Playfair’s law and its use in testing long-term stream ero-
sion models: Earth Surface Processes and Landforms, v. 26, p. 1317–1332,
doi: 10.1002/esp.272.
Rech, J.A., Currie, B.S., Michalski, G., and Cowan, A.M., 2006, Neogene climate
change and uplift in the Atacama Desert, Chile: Geology, v. 34, p. 761–764,
doi: 10.1130/G22444.1.
Reiners, P.W., Spell, T.L., Nicolescu, S., and Zanetti, K.A., 2004, Zircon
(U-Th)/He thermochronology: He diffusion and comparisons with
40Ar/39Ar dating: Geochimica et Cosmochimica Acta, v. 68, no. 8, p. 1857–
1887, doi: 10.1016/j.gca.2003.10.021.
Schurr, B., Rietbrock, A., Asch, G., Kind, R., and Oncken, O., 2006, Evidence for
lithospheric detachment in the central Andes from local earthquake tomog-
raphy: Tectonophysics, v. 415, p. 203–223.
Springer, M., and Förster, A., 1998, Heat-fl ow density across the Central Andean
subduction zone: Tectonophysics, v. 291, p. 123–139, doi: 10.1016/S0040-
1951(98)00035-3.
Thouret, J.-C., Wörner, G., Singer, B., and Finizola, A., 2005, The central Andes
in Peru: “Old” valleys in a “young” mountain range?, in Proceedings, 6th
International Symposium on Andean Geodynamics, Barcelona, 12-14 Sep-
tember 2005, Extended Abstracts, p. 726–729.
Victor, P., Oncken, O., and Glodny, J., 2004, Uplift of the western Altiplano
plateau: Evidence from the Precordillera between 20 and 21°S (northern
Chile): Tectonics, v. 23, TC4004, doi: 10.1029/2003TC001519.
Whipple, K.X, and Tucker, G.E., 2002, Implications of sediment-fl ux-dependent
river incision models for landscape evolution: Journal of Geophysical
Research, v. 107, 2039, doi: 10.1029/2000JB000044.
Wörner, G., Hammerschmidt, K., Henjes-Kunst, F., Lezaun, J., and Wilke, H.,
2000, Geochronology (40Ar/39Ar, K-Ar and He-exposure ages) of Cenozoic
magmatic rocks from northern Chile (18–22°S): Implications for magma-
tism and tectonic evolution of the central Andes: Revista Geológica de
Chile, v. 27, no. 2, p. 131–137.
Manuscript received 10 November 2006
Revised manuscript received 18 January 2007
Manuscript accepted 25 January 2007
Printed in USA
... The Neogene uplift of the Andes has been studied and quantified principally in the Altiplano region. The area between the Altiplano and the coast (forearc) has also been uplifted, but probably to a lesser degree (e.g., Schildgen et al., 2007;Jordan et al., 2010;Evenstar et al., 2015). The coastal part of the Central Andes forearc currently seems to be uplifting almost everywhere, except in central Peru between 8 and 14 • S (Regard et al., 2010;Melnick, 2016;Martinod et al., 2016b;Saillard et al., 2017), as indicated by uplifted marine terraces and rasas (e.g., Hsu et al., 1989;Goy et al., 1992;Machare and Ortlieb, 1992;Ortlieb et al., 1996;Saillard, 2008;Regard et al., 2010;Saillard et al., 2011;Regard et al., 2017) or uplifted Neogene formations (e.g., DeVries, 1998;Macharé et al., 1988;Pena et al., 2004;Vega and Marocco, 2004;Hall et al., 2008;Rodriguez et al., 2013;Alván and von Eynatten, 2014). ...
... These works emphasize that it is hard to date precisely landforms older than 0.5 Ma. Other Neogene chronological milestones are given by: (i) the remnants of fan deltas (Pena et al., 2004;Alván and von Eynatten, 2014), (ii) the geomorphological evolution in relation to supergene mineralization (Quang et al., 2005), and (iii) the canyon incision history (Kober et al., 2006;Schildgen et al., 2007Schildgen et al., , 2009a; Thouret et al., 2007;Wipf et al., 2008). These milestones are indirect and are therefore not accurate indicators for uplift; consequently, the Plio-Pleistocene uplift history of the forearc is still uncertain, which limits our ability to link potential uplift variations with tectonic controls. ...
... The main surface of the pampa was later incised, producing deep canyons where rivers flow. This entrenchment was interpreted as the consequence of the forearc uplift (Kober et al., 2006;Schildgen et al., 2007Schildgen et al., , 2009aGarcia et al., 2011;Evenstar et al., 2017). Sediments from the Camana formation indicate that the rapid uplift initiated around 12-9 Ma (Alván and von Eynatten, 2014), compared to the 11-8 Ma period for the canyon incision initiation obtained using low temperature thermochronology by Schildgen et al. (2009a). ...
Conference Paper
Full-text available
We explore coastal morphology along a 500 km long uplifted coastal segment in the central Andes, between the cities of Chala (Peru) and Arica (Chile). We use accurate DEM and field studies to extract uplifted shoreline sequences along the study area. In addition, we consider continental pediment surfaces that constrain both the geographic and vertical extent of marine landforms. We establish a chronology based on published dates for marine landforms and pediment surfaces. We extend this corpus with new 10Be data on uplifted shoreline platforms. We find that the last 12 Ma is marked by three periods of coastal stability or subsidence dated at ~12-11 Ma, ~8-7 Ma, and ~5-2.5 Ma. Uplift that accumulated between these periods of stability has been ~1000 m since 11 Ma; its rate is up to 0.25 mm/a (m/ka). For the last period of uplift only, during the last 800 ka, the forearc uplift has been accurately recorded by the cutting of numerous coastal sequences. Within these sequences, we correlated marine terraces with high sea levels (interglacial stages and substages) up to MIS 19 (790 ka), i.e., with a resolution of ~100 ka. The uplift rate for this latter uplift period increases westward from 0.18 mm/a at the Peru-Chile border to ~0.25 mm/a in the center of the study area. It further increases northwestward to 0.45 mm/a due to the influence of the Nazca Ridge. In this study, we document unusual cyclic forearc uplift with cycles of ~4 Ma in duration. This periodicity is consistent with predictions made by Menant et al. (2020) based on numerical models, and may be related to episodic tectonic uplift (subduction slab detachment) beneath the coastal forearc zone.
... Andean uplift and the shift to hyperarid climate conditions have commonly been considered as the main extrinsic control on the Neogene large-scale landscape evolution of the Atacama Desert (Isacks, 1988;Rech et al., 2006Rech et al., , 2010Rech et al., , 2019Hoke et al., 2007;Jordan et al., 2010Jordan et al., , 2014Amundson et al., 2012). The timing of Andean uplift is controversial, although rapid Late Miocene surface uplift has been classically associated with the formation of the current deep canyons that cross the western Andean slope (e.g., Thouret et al., 2007;Schildgen et al., 2007;Fox et al., 2015;Farías et al., 2005;García et al., 2011;Barnes and Ehlers, 2009;Hoke et al., 2007). In the southern Atacama Desert, scarce geochronological constraints about the ages of canyon formations are reported. ...
... Valleys formed at that time were as deep as ~2 km below the highest neighboring summits, indicating that the Precordillera was already uplifted and reached altitudes of at least 2000 m (Riquelme et al., 2007). Nonetheless, a rapid Late Miocene surface uplift, like the one attributed to the formation of the canyons in northern Chile (Thouret et al., 2007;Schildgen et al., 2007;Farías et al., 2005;García et al., 2011;Barnes and Ehlers, 2009), can be deduced from the shift of the alluvial backfilling deposition from the Precordillera and Eastern-CD to the Western-CD. ...
Article
Pediplains are classically identified as flat landscape surfaces in arid regions linked to tectonic quiescence, whereas deep incision of a pediplain is attributed to tectonic uplift. In the Atacama Desert, pediplains are generally used as morphotectonic markers to define the chronology of episodes of Late Cenozoic Andean uplift from their erosion and incision patterns and timings. The Atacama Pediplain (AP) extends over >12,000 km² (26° to 28°S Lat) through the Central Depression and Precordillera of the southern Atacama Desert. In this study we perform geomorphologic and stratigraphic observations on the AP in the Salado Canyon area, combining new geochronological results derived from ⁴⁰Ar/³⁹Ar biotite ages from volcanic layers interbedded within the alluvial deposits, and ²¹Ne exposure ages on quartz-clasts on alluvial plains, to determine the chronology of the AP evolution. Results show that the evolution of the AP is a long-term and continuous process (from >20 to ~2.3 Ma) of alluvial deposition and subsequent alluvial plain formation developed by interplay between the climate variability of the Atacama Desert and Andean uplift. The AP alluvial deposition occurred in two spatially and temporally separated episodes of alluvial backfilling: 1) shortly before ~20.14 Ma and to prior to ~9.4 Ma, a timespan that allows for the drainage capture of the eastern Precordillera, and considerable landscape rearrangement; 2) post ~9.4 Ma, with a re-positioning of alluvial backfilling from the Precordillera towards the Central Depression. This occurs after the Mid Miocene onset of the hyperarid conditions in the Central Depression and is more likely due to late Miocene surface uplift. Exposure ages reveal the cessation of basin-scale deposition and the abandonment of the alluvial plains during ~5.24 to ~3.8 Ma linked to the incision of Salado Canyon. Subsequent climate conditions modulated the surfaces by the development of lag deposits until ~2.69 to ~2.3 Ma when hyperarid conditions reach a threshold that limits surficial activity. Additionally, the drainage capture of the Precordillera by headwards erosion of the Salado Canyon explains marked deep incision depth of this canyon without the need for an increase in surface uplift or a change in climate conditions. The AP is not a general marker of a single climatic or tectonic event/period but a composite paleosurface formed by a complex concatenation of extrinsic and intrinsic geomorphic processes over more than ~17 myr.
... Mountainous environments in the dry Andean region are intrinsically prone to erosion processes, i.e. gullies or landslides, due to their accentuated topography and infrequent, but large, rainfall events (Molina et al., 2008;Clark et al., 2016;Morera et al., 2017). The rugged western Peruvian Andes (4°S-18°S) is located between the Pacific coastal plain and the central Andes and is characterised by elevations of up to 5000 m and up-to 3000 m deep, and incised, braided, river valleys (Schildgen et al., 2007). The alluvial plains are formed by the accumulation of sediments sourced from the Andean region and delivered to the stream network by ephemeral channels and gullies that are activated during seasonal extreme rainfall events (Morera et al., 2017). ...
Article
The western Peruvian region is prone to erosion and geomorphic change. Extreme precipitation events lead to rapid change in river channel and floodplain morphology due to bank erosion and debris flows delivering detrital material to the fluvial system. Monitoring geomorphic events and their associated topographic changes at high spatial and temporal resolutions remains a challenge. Here, we used an Uncrewed Aerial Vehicle - Post-Processing Kinematic - Structure from Motion (UAV-PPK-SfM) approach that includes co-registration of point clouds by using relative Ground Control Points (GCPs). This workflow adjusts each elevation model to a reference model using invariant features that did not change their position or form over time. We applied this technique to monitor landscape change (2019-2021) in an area of 0.3 km2 located in the Cañete River basin. Our results showed that a minimum observable elevation change of 0.56 m (95% confidence interval) can be achieved using this workflow, beyond which an actual elevation change can be separated from systematic error. Using object-based classification techniques on the aerial images, we separated geomorphic dynamics from land cover changes. This allowed us to isolate the effect of geomorphic processes, and quantify rates related to gully erosion, river scouring, bank erosion, and sediment deposition. Within the study area, a hotspot of geomorphic change corresponded to an ephemeral tributary channel. The gully channel incising an alluvial fan is highly dynamic, showing bank erosion of 0.75 to 3.2 m and net export of 37 m3 of sediment in the 25-month study period. Given that the monitoring period did not include high intensity rainfall events, the study illustrates how geomorphic activity in ungauged Andean river basins, such as the Cañete valley, may be considerably underestimated in literature.
... In general, low-temperature thermochronometric data provide cooling or exhumation information of the rock that the river is incising into (Reiners and Brandon, 2006). This exhumation can only constrain river incision if differential exhumation between the interfluves on the river valley flanks and the valley bottom can be resolved (e.g., Schildgen et al., 2007;Flowers et al., 2015). ...
... Convergence and subduction of the Nazca Plate beneath the South American Plate began in the late Paleozoic (Mamani et al., 2010) and the margin became a well-organized compressional system in the Late Cretaceous (Horton, 2018a). The onset of orogenesis is marked by Late Cretaceous magmatism and exhumation in the Western Cordillera (Vicente, 1990;Mamani et al., 2010), supported by an extensive thermochronological dataset spanning southern Peru-northern Chile showing widespread exhumation starting in the Late Cretaceous (Andriessen and Reutter, 1994;Maksaev and Zentilli, 1999;Schildgen et al., 2007;Wipf et al., 2008;Ruiz et al., 2009;Schildgen et al., 2009b;Gunnell et al., 2010;Juez-Larré et al., 2010;Reiners et al., 2015;Avdievitch et al., 2018;Henriquez et al., 2019). Coeval with the onset of Andean orogenesis, deposition in the Altiplano region transitioned from marine to non-marine during Late Cretaceous-Paleocene retroarc foreland basin evolution (Sempere, 1995). ...
Article
Full-text available
Highlights • Detrital zircon (DZ) geochronology documents chronostratigraphy based on maximum depositional ages and sediment provenance. • DZ data show a Paleogene unconformity/condensed section followed by rapid sediment accumulation and provenance changes. • Southward migrating basin reorganization tracks initiation of Eastern Cordilleran exhumation and an inboard magmatic sweep. • Upper plate systems responded to southward migrating, late Paleocene–early Miocene flat slab subduction. • Slab flattening was driven by sequential subduction of a Manihiki Plateau Conjugate and the Juan Fernandez Ridge.
... Thirty-one paleoelevation studies in the Central Andes utilising paleoclimate and paleoaltimetry proxies (e.g., Ghosh et al., 2006;Scott et al., 2018; for full list see supplementary material) geomorphic analysis of erosion surfaces (e.g., Victor et al., 2004;Barke and Lamb, 2006; for full list see supplementary material) and reconstruction of fluvial incision rates (e.g. Schildgen et al., 2007;Evenstar et al., 2020; for full list see supplementary material) are compiled for the Western Cordillera, Altiplano and Eastern Cordillera in Figs. 5 and 6. The latter two, erosional surfaces and fluvial incision rates, can be used as proxies for the timing of surface uplift in regions which are highly active with large rivers (Demoulin et al., 2017 and references within). ...
Article
The rise of large mountain ranges is considered to be driven by tectonics potentially coupled with climate driven-erosion, although the role of this coupling remains uncertain. The arid climate of the Central Andes allows us to strengthen our understanding of the relative roles of these processes in mountain range development globally. Here we compile estimates of exhumation, sedimentation, aridity and surface uplift across the Central Andes for the last 50 Ma. We aim to place constraints on the relative timing of rock uplift (displacement of rocks with respect to the geoid), exhumation (displacement of rocks with respect to the surface) and surface uplift (displacement of the earth's surface with respect to the geoid). We show that initial rock uplift of the Andes extends back at least 50 Myr. This rock uplift generated orographically driven precipitation on windward slopes leading to increased exhumation but limited preservation of surface uplift. Eastward propagation of the mountain range resulted in increasingly extreme orographic effects on the leeward side amplifying aridity, reducing exhumation and increasing preservation of surface uplift. Essentially, surface uplift shows a ∼5-10 Myr lag behind initial rock uplift as the Andes grow asymmetrically through time. We suggest that an eastward propagating pattern of exhumation, aridity and surface uplift with time, reconciles previous contradictory models of Andean uplift. One Sentence Summary Uplift of the Central Andes is reconstructed over the last 50 Myr and the precise relationship between roles of tectonics and climate established.
... In the western Altiplano, volcanic and sedimentary Cenozoic rocks are dominant ( Fig. 1) and the geologic evolution indicates major deformation and uplift during the Eocene Incaic orogeny and Miocene Quechua orogeny (Charrier et al., 2013;García et al., 2017). In the Western Cordillera, older rocks are locally exposed and their thermochronological record is poor, whereas in the forearc (from the Coastal Cordillera to Chilean Precordillera), the basement is Jurassic-Paleocene in age and their thermochronological data are largely older than 24 Ma (Fig. 1) and indicate very slow exhumation since the Cretaceous (Schlunegger et al., 2006;Schildgen et al., 2007;Avdievitch et al., 2018). ...
Article
Full-text available
In the Western Cordillera of northern Chile, the Proterozoic-Paleozoic Belén Metamorphic Complex is covered by late Oligocene-early Miocene (25-18 Ma) rocks, and both units are involved in west-vergent contractional deformation, which results in exhumation. A Miocene age (18 to 6 Ma) for deformation has been previously constrained by stratigraphy and cross-cutting relationships. To understand the youngest exhumation event and reverse faulting, we obtained 21 (U-Th)/He ages from two samples of the metamorphic rocks and the associate inverse thermal modeling. Five zircon (U-Th)/He ages from one sample are 113 to 226 Ma, very scattered, while five zircon ages from the other, are 20 to 49 Ma. The high dispersion of zircon (U-Th)/He data prevents the geological interpretation of results. Apatite grains from both samples yielded 11 (U-Th)/He ages between 10.4 and 18.7 Ma, with 9 values from 12.0 to 15.5 Ma. A slight positive correlation between apatite single-grain dates and effective uranium for 4 crystals of one sample suggests relatively slow cooling. The T-t model including these 4 apatite ages shows continuous cooling from 15 to 0 Ma with a relatively more marked cooling period at 11-7 Ma. The middle-late Miocene thermal signal agrees with the geologic evolution of the region and would permit to date the last activity of the Chapiquiña-Belén reverse fault, which uplifted and exhumed the metamorphic rocks. This signal is relatively similar to that the eastern Altiplano, but differs considerably from that the forearc.
Article
Full-text available
This study assesses the impact of fold-thrust belt driven deformation on the topographic evolution, bedrock exhumation and basin formation in the southeastern Peruvian Andes. We do this through a flexural and thermokinematically modelled balanced cross-section. In addition, published thermochronology samples from low-elevation (river canyons) and high-elevation (interfluves) and Cenozoic sedimentary basin datasets along the balanced cross-section were used to evaluate the age, location, and geometry of fault-driven uplift, as well as potential relationships to the timing of ∼2 km of canyon incision. The integrated structural, thermochronologic, and basin data were used to test the sensitivity of model results to various shortening rates and durations, a range of thermophysical parameters, and different magnitudes and timing of canyon incision. Results indicate that young apatite (U-Th)/He (AHe) canyon samples from ∼2 km in elevation or lower are consistent with river incision occurring between ∼8–2 Ma and are independent of the timing of ramp-driven uplift and accompanying erosion. In contrast, replicating the young AHe canyon samples located at >2.7 km elevation requires ongoing ramp-driven uplift. Replicating older interfluve cooling ages concurrent with young canyon ages necessitates slow shortening rates (0.25–0.6 mm/y) from ∼10 Ma to Present, potentially reflecting a decrease in upper plate compression during slab steepening. The best-fit model that reproduces basin ages and depositional contacts requires a background shortening rate of 3–4 mm/y with a marked decrease in rates to ≤0.5 mm/y at ∼10 Ma. Canyon incision occurred during this period of slow shortening, potentially enhanced by Pliocene climate change.
Article
Full-text available
Thermochronological data are essential to constrain thermal and exhumation histories in active mountain ranges. In the Central Andes, bedrock outcrops are rare, being blanketed by widespread late Palaeogene–Neogene and younger volcanic formations. For this reason, the exhumation history of the Western Cordillera (WC) in the Peruvian Andes has only been investigated locally along the mountain range. Dense thermochronological data are only available in canyons of the Arequipa (16° S) and Cordillera Negra regions (10° S). We present new apatite (U-Th)/He and fission-track data from the 1 km deep Cañete Canyon (13° S), where the Oligo-Miocene deposits are preserved lying conformably on an Eocene palaeo-topographic surface. Thermal modelling of thermochronological data indicate that the 30–20 Ma ignimbrite deposits overlying the bedrock were thick enough to cause burial reheating. We demonstrate that burial associated with thick volcanic formations should be taken into account when interpreting thermochronological data from the WC or in similar volcanic-arc settings.
Article
Full-text available
K-Ar and Ar/Ar ages from magmatic rocks of northern Chile (18-22°S) describe duration and extent of the Tertiary and Quaternary magmatic evolution and date major tectonic events in northernmost Chile. This paper summarizes new K-Ar and Ar-Ar mineral and whole rock ages for intrusive rocks from the Precordillera, Tertiary ignimbrites and andesitic stratovolcanoes from the Western Andean Escarpment at 18°S (WARP) and the volcanic front. Intrusive rocks of the Precordillera (Quebrada Paguana, Quebrada Blanca, Quebrada Choja, Quebrada Guatacondo, Cerro Chandacolla) represent the Cretaceous to Eocene magmatic arc system and gave between 45 and 35 Ma. Younger ages on intrusive rocks are invariably caused by deuteric alteration. Ignimbrites of the Putani and Oxaya formations gave Ar-Ar sanidine ages around 24.2 to 24.8 Ma and 22.8 to 19.4 Ma, respectively. Andesitic stratovolcanoes, which directly overlie Oxaya ignimbrites east of the Western Cordillera gave ages of 20.3 Ma (Cordon Quevilque) to 9.0 Ma (Cerro Margarita). Samples from the Miocene to Pleistocene arc system on the Chilean Altiplano underlying the volcanoes of the active volcanic front have been dated between 10.5 to Ο3 Ma. A widespread ignimbrite can be correlated from the Lauca basin to the Pacific coast and to the east to occurrences of near Pérez. Repeated Ar-Ar sanidine dating of the Lauca-Pérez-ignimbrite resulted in highly concordant ages of 2.71±0.25 Ma, 2.72 Ma±0.01 Ma, and 2.73±0.11 Ma. Rocks from the active chain (Volcanic Cordillera) gave ages younger than 0.9 Ma (Volcán Irruputuncu, Volcán Olca, Volcán Aucanquilcha, Volcán Ollagüe, Volcán Porun̄ita). These new data are used to constrain Miocene stratigraphy and tectonic movements as well as the timing of uplift and sedimentary response at the Western Andean Escarpment within the framework of the tectonic evolution of the Central Andes.
Article
Full-text available
The enigma of continental plateaus formed in the absence of continental collision is embodied by the Altiplano-Puna, which stretches for 1800 km along the Central Andes and attains a width of 350–400 km. The plateau correlates spatially and temporally with Andean arc magmatism, but it was uplifted primarily because of crustal thickening produced by horizontal shortening of a thermally softened lithosphere. Nonetheless, known shortening at the surface accounts for only 70– 80% of the observed crustal thickening, suggesting that magmatic addition and other processes such as lithospheric thinning, upper mantle hydration, or tectonic underplating may contribute significantly to thickening. Uplift in the region of the Altiplano began around 25 Ma, coincident with increased convergence rate and inferred shallowing of subduction; uplift in the Puna commenced 5–10 million years later.
Article
Full-text available
We analyze the west vergent thrust system (WTS) along the western flank of the Altiplano in northern Chile (18°S–21°S). In our study area (19°20′S–19°50′S), the WTS consists of three thrust propagation monocline folds (flexures) developing growth strata. The relative uplift accommodated by the flexures is rapid between 26 and 8 Ma (0.1 mm/yr), diminishing to 0.02 mm/yr after 8 Ma. Approximately 2000 m of relative surface uplift was accommodated by the flexures since the late Oligocene. Sedimentological and geomorphological analysis shows that westward tilting of the forearc occurred after 10 Ma, coeval with the shifting of deformation from the Altiplano to the sub-Andean zone, where the underthrusting of the Brazilian Craton would have resulted in crustal thickening, surface uplift in the orogen, and westward ductile subcrustal flow. Forearc tilting is accommodated by east vergent thrusts (ETS) issued from the Benioff zone beneath the Central Depression emerging into the Western Cordillera, contributing 500–1400 m of surface uplift. The WTS connects the ETS in the brittle-ductile crustal transition (∼25 km depth), continuing farther east as the Altiplano low-velocity zone, configuring the western Altiplano as a crustal-scale fault bend fold. Forearc tilting would be caused by westward ductile flow in the lower crust pushing the rigid forearc in the ETS. Meanwhile, between 19°S and 21°S, the WTS accommodates dextral strike slip, and ∼3 km of N-S shortening occurred in the Coastal Cordillera. Transcurrence and strain partitioning are probably the result of slight plate convergence obliquity, strong coupling within the interplate zone, westward continental concavity, and high elevation opposing horizontal contraction.
Article
Full-text available
Sedimentological data indicate that a semi-arid/arid climate prevailed across the Central Andes from 15 Ma to 4 Ma. Between 4 and 3 Ma a switch to hyperaridity occurred along the western margin of South America. Palaeoaltitude data suggest that a substantial proto-Central Andean mountain range was in place between 15 and 9 Ma. These data support the idea that the Andean rain shadow existed by 15 Ma and that it reinforced the preexisting climatic regime rather than changing it. The change to hyperaridity in western South America is attributed to a combination of global climate cooling and enhanced upwelling of the Humboldt current generated by closure of the Central American Seaway between 3.5 and 3 Ma, and not to the Andean rain shadow.
Article
The topographic data combined with information on structure, magmatism, seismicity, and paleomagnetism support a simple kinematic model for the late Cenozoic evolution of the central Andes. The model does not require collisional effects or enormous volumes of intrusive additions to the crust but instead calls upon plausible amounts of crustal shortening and lithospheric thinning. The model interrelates Andean uplift, a changing geometry of the subducted Nazca plate, and a changing outline (in map view) of the leading edge of the S American plate.-from Author
Article
Lithospheric delamination is the foundering of dense lithosphere into less dense asthenosphere. The causes for this density inversion are thermal, compositional, and due to phase changes. For delamination to occur in the specific, and probably common, case where lithospheric mantle is intrinsically less dense than underlying asthenosphere due to composition differences, a critical amount of shortening is required for the densifying effect of cooler temperature to counterbalance the effect of composition. Crustal thickening that results from shortening may result in a crustal root that, due to phase changes, becomes denser than the underlying mantle lithosphere and should delaminate with it: most of the negative buoyancy resides at the top of the mantle and the bottom of the crust. In most cases composition is not known well enough to calculate the driving energy of delamination from densities of equilibrium mineral assemblages in a lithospheric column Poorly known kinetics of phase changes contribute additional uncertainties. In all cases however, the effects of delamination under a region are readily recognizable: rapid uplift and stress change, and profound changes in crustal and mantle-derived magmatism (a reflection of changes in thermal and compositional structure). Characteristics of delamination magmatism are exhibited in the Southern Puna Plateau, central Andes. The consequences of delamination for theories of crustal and mantle evolution remain speculative, but could be important. Recognition of delamination-related magmas in older (including Archean) orogens may be the best way to recognize past delamination events, because the magmas are among the most indelible and least ambiguous of delamination indicators.
Article
We report 35 new and 9 revised terrestrial heat flow measurements overlying the Andean subduction zone in Bolivia and Peru. The measurement sites are distributed in the Andean Cordillera, the Altiplano, the sub-Andean ranges, and the sedimentary platform and basins to the east of the Andes. They fall in the distance range 75-900 km form the Peru-Chile trench. Sites in the Peruvian Cordillera have a mean heat flow of 41 mW m-2, whereas those in the Bolivian Cordillera and Altiplano average 84 mW m-2. The sub-Andean ranges and the adjacent sedimentary platform have a mean heat flow of 50 mW m-2. The higher heat flow of the Bolivian Cordillera and Altiplano lies to the east of Quaternary volcanoes along the Bolivia-Chile border and in southernmost Peru and thus can be recognized as a ``back arc'' heat flow high. Neither Quaternary volcanism nor high heat flow are present in central and northern Peru. This contrasting along-strike pattern correlates with the variable angle of subduction of the Nazca plate beneath the region. Beneath Bolivia the subduction is at about 30°-35°, whereas beneath Peru it is subhorizontal and may be providing a cold underplate to the overlying Peruvian lithosphere. Extensive Miocene volcanism in Peru suggests, however, that heat flow in Peru 10 Ma ago was likely similar to that in Bolivia in the present day and implies a chage in subduction and a rapid diminution of heat flow in Peru over the past 10 Ma. Imbrication of cold oceanic lithosphere beneath Peru, analogous to that beneath western British Columbia, may provide a mechanism for rapid reduction of the heat flow.
Article
High-precision stepped-heating experiments were performed to better characterize helium diffusion from apatite using Durango fluorapatite as a model system. At temperatures below 265°C, helium diffusion from this apatite is a simple, thermally activated process that is independent of the cumulative fraction of helium released and also of the heating schedule used. Across a factor of ~4 in grain size, helium diffusivity scales with the inverse square of grain radius, implying that the physical grain is the diffusion domain. Measurements on crystallographically oriented thick sections indicate that helium diffusivity in Durango apatite is nearly isotropic. The best estimate of the activation energy for He diffusion from this apatite is Ea=33+/-0.5kcal/mol, with log(D0)=1.5+/-0.6cm2/s. The implied He closure temperature for a grain of 100 mum radius is 68°C assuming a 10°C/Myr cooling rate; this figure varies by +/-5°C for grains ranging from 50 to 150 mum radius. When this apatite is heated to temperatures from 265 to 400°C, a progressive and irreversible change in He diffusion behavior occurs: Both the activation energy and frequency factor are reduced. This transition in behavior coincides closely with progressive annealing of radiation damage in Durango apatite, suggesting that defects and defect annealing play a role in the diffusivity of helium through apatite.
Article
Newly dated Tertiary strata in the Bolivian central Andean plateau and synthesis of the Tertiary record in the adjacent Subandean fold-thrust belt constrain the age of deformation in both regions. Age of deformation within the plateau is determined by dated crosscutting relations associated with a regionally extensive high-level surface known as the San Juan del Oro surface. New 40Ar-39Ar dates on undeformed strata above the high-level surface preclude significant upper-crustal shortening within the Eastern Cordillera after 10 Ma. Tertiary strata within the adjacent Subandean region demonstrate that formation of the fold-thrust belt occurred after 10 Ma. On the basis of these data, we propose a two-stage model of late Cenozoic Andean growth that links plateau uplift to the development of the fold-thrust belt. In the first stage, early plateau uplift occurred in response to widespread compressional deformation of the plateau (Eastern Cordillera and Altiplano). During the second stage, beginning after 10 Ma, upper-crustal deformation within the plateau terminated, and the Subandean fold-thrust belt developed. Crustal-scale eastward thrusting along the eastern margin of the Eastern Cordillera drove Subandean folding and thrusting; the Eastern Cordillera served as the "bulldozer" for the deforming Subandean wedge.